Mathematical Cardiac Electrophysiology: Piero Colli Franzone Luca F. Pavarino Simone Scacchi
Mathematical Cardiac Electrophysiology: Piero Colli Franzone Luca F. Pavarino Simone Scacchi
Mathematical Cardiac
Electrophysiology
MS&A
Modeling, Simulation & Applications
ABC
MS&A
Volume 13
Editor-in-Chief
A. Quarteroni
Series Editors
T. Hou
C. Le Bris
A.T. Patera
E. Zuazua
More information about this series at
http://www.springer.com/series/8377
Piero Colli Franzone • Luca F. Pavarino •
Simone Scacchi
Mathematical Cardiac
Electrophysiology
123
Piero Colli Franzone Luca F. Pavarino
Dipartimento di Matematica Dipartimento di Matematica
UniversitJa degli Studi di Pavia UniversitJa degli Studi di Milano
Pavia Milano
Italy Italy
Simone Scacchi
Dipartimento di Matematica
UniversitJa degli Studi di Milano
Milano
Italy
The goal of this book is to present some of the most successful and advanced
mathematical and numerical models used in the field of cardiac electrophysiology.
The bioelectric activity of the heart is the subject of a vast and still growing
interdisciplinary literature in medicine, physiology, bioengineering, mathematical
biology, chemistry, physics and bioinformatics. The long history and diversity of
this field of research is shown e.g. by the earlier monographs by Jack, Noble and
Tsien [250], Peskin [383], Nelson and Gezelowitz [346], Pilkington and Plonsey
[386], by the reference works by Pilkington et al. [387], Panfilov and Holden [369],
Keener and Sneyd [273], Gulrajani [215], Plonsey and Barr [393], Efimov et al.
[158], and by some of the periodic review books by Zipes and Jalife [588–590].
In the last decade, more specific monographs on computational electrocardiology
were published by Sachse [446], Pullan at el. [406], Sundnes et al. [502]; see
also Macfarlane et al. [312]. Since the emphasis of this book is on mathematical
and numerical aspects of the models and algorithms presented, we hope that this
monograph will present new elements and will complement the works above.
Electrocardiology deals with the description of both intracardiac bioelectric
phenomena and the extracardiac electric field generated in the animal or human
body. The practice of modern medicine relies on noninvasive imaging technologies,
such as CT, MRI and PET, for diagnostic purposes and for driving therapeutic
procedures. Even though cardiac arrhythmias are among the major causes of death
and disability, a noninvasive imaging technique yielding an accurate and reliable
diagnosis of the electrophysiological state of the heart is not yet available. Clinic
Electrocardiography deals with the detection and interpretation of noninvasive
potential measurements collected from the time course of the usual electrocardio-
grams (ECG) at a few points on the body surface or from the evolution of body
surface maps, i.e. potential distribution maps on the body surface reconstructed
from measures at numerous electrodes (100 or more, see the surveys by [504, 511]).
Since the electrode location of the ECG is centimeters away from the heart
surface and the current conduction from heart to thorax yields a strong signal
v
vi Preface
attenuation and smoothing, the information content of ECGs and body maps is
limited and it is a difficult task to extract from these signals detailed information
on pathological heart states associated with ischemia or sudden death. Indeed, the
origin of arrhythmogenic activity or the existence of abnormal electrophysiological
substrates in many cases may not be easily inferred from the sequence of cardiac
excitation.
The scientific base of Electrocardiology is the so-called Forward Problem of
Electrocardiology, i.e. modeling the bioelectric cardiac sources and the conducting
media in order to derive the potential field. Of considerable applicative interest are
also the so-called Inverse Problems of Electrocardiography in terms of potentials
(see e.g. the review [216, 442] and [68, 69, 415]) or in terms of the cardiac sources
(see e.g. [110, 439]).
In the past few decades, experimental electrophysiology has been increasingly
supported by the mathematical and numerical models of computational electro-
cardiology. The formulation of models at both cellular and tissue levels provide
essential tools in order to integrate the increasing knowledge of the bioelectrochem-
ical phenomena occurring through cardiac cellular membranes. Detailed cellular
phenomena are described in microscopic membrane models and the latter are then
inserted in macroscopic tissue models in order to investigate their effects at tissue
level. These coupled models are then validated by comparing simulated results with
experimental in vitro and in vivo data, generating a feedback loop that may lead to
improved and more detailed models and/or the redesign of new experiments. As a
further step, these electrophysiological models are being increasingly coupled and
integrated with mechanical models of tissue deformation, hemodynamical models
of cardiac blood flow and more in general with models of the cardiovascular
system. This complex integrative effort is the current focus of several research
projects, for example such as the Physiome Project (www.physiome.org.nz) and the
EC-sponsored Virtual Physiological Human (VPH) Initiative (www.vph-noe.eu).
Ultimately, the integration of these models should provide new tools enabling the
biomedical community to link genetic and proteomic databases to anatomy and to
functions at the cellular, tissue and organ level.
From a macroscopic point of view, the Forward Problem of Electrocardiology
is described by the so-called Bidomain model for the evolution of the intra,
extracellular and extracardiac potential fields. The two main components of the
Bidomain model are: (a) the dynamics of the ionic current flow through the cardiac
cellular membrane, modeled by a system of ordinary differential equations and
(b) a macroscopic representation of the cardiac tissue modeled as a bidomain
superposition of the intra and extra cellular media characterized by anisotropic
conductivity tensors associated with the fiber architecture of the myocardium.
The Bidomain model is computationally expensive because of the involvement of
different space and time scales. In fact, meaningful portions of cardiac tissue have
sizes on the order of centimeters, while the steep potential gradient is localized in
a thin layer about 1 mm thick, requiring discretizations on the order of a tenth of
millimeter. Moreover, a normal heartbeat can last on the order of 1 s, while the
time constants of the rapid kinetics involved range from 0.1 to 500 ms, requiring in
Preface vii
some phases time steps on the order of the hundredths of milliseconds (or less when
currents or shocks are applied). Therefore, in realistic three-dimensional models it is
possible to have discrete problems with more than O.107 / unknowns at every time
step and simulations have to be run for many thousands of time steps.
A simplified cardiac tissue model is the anisotropic Monodomain system, i.e. a
parabolic reaction-diffusion equation describing the evolution of the transmembrane
potential coupled with an ionic membrane model. This model has been widely used
for three-dimensional simulations due to its reduced computational costs.
Current large-scale simulations of whole heartbeats using Bidomain and Mon-
odomain models require adaptive and parallel tools in order to reduce their high
computational cost. While both tools can in principle be applied to both space and
time, most studies employ adaptive methods in time and parallel solvers in space,
since the other alternatives are still the subject of current research even for simpler
model problems in two dimensions. Therefore in this book, we present the main
numerical techniques for efficiently simulating cardiac reaction-diffusion models.
In particular, we focus on scalable parallel Bidomain solvers that are capable of
efficiently scaling their performance for increasing processor counts in current and
future multicore parallel computers.
Among the important aspects of cardiac modeling not covered in this book are
cardiac mechanics, blood flow, electro-mechanical and fluid-mechanical coupling,
and cardiac imaging. Research in these fields is also growing tremendously and a
separate book would be necessary to properly present the main mathematical and
numerical models available. For an overview of these related fields, we refer to e.g.
the monographs [47, 140, 242, 369, 446], the works [138, 232, 246, 384, 385, 491]
with the references therein, and the recent proceedings of the conferences FIMH
(Functional Imaging and Modeling of the Heart) [21, 180, 266, 316, 327, 366, 447],
CINC (Computing in Cardiology, http://cinc.org/archives/2013/, http://cinc.org/
archives/2012/), STACOM (Statistical Atlases and Computational Models of the
Heart) [74–76].
The book is structured in the following chapters.
In Chap. 1, we give a brief review of the basic physiology and anatomy
of the heart, including the specialized cells of the cardiac conduction system,
working cardiomyocytes, fibroblasts, extracellular matrix, collagen, gap junctions,
connexin, cardiac stem cells and the fiber and laminar architecture of the ventricular
myocardium. We then present the main phases of a cardiac action potential, its
spatial and temporal heterogeneity, and continue by describing the main features
of an electrocardiogram (ECG), with its leads, deflections, intervals and main
alterations. The chapter concludes with a review of the main cardiac imaging
techniques currently available.
Chapter 2, introduces the fundamental tools for modeling the bioelectric activity
of excitable cells: the Nernst – Planck equation, the Goldman-Hodgkin-Katz (GHK)
current-voltage relation, and the Nernst equilibrium potential, together with its ther-
modynamical derivation. Next, the Poisson-Nernst-Planck (PNP) electrodiffusion
model is derived and two classical current – voltage relations are obtained in the
short and long channel limits. With these tools, we can define the basic electrical
viii Preface
circuit model of the cellular membrane, where the transmembrane current, modeled
as the sum of the capacitive and ionic currents through the membrane, must balance
the given applied current. The ionic currents are then described by using the classical
ion channel gating models, allowing us to build cardiac action potential models. We
start with the celebrated Hodgkin-Huxley (H-H) model and briefly review some of
the historical ventricular models based on the H-H formalism, such as the Beeler-
Reuter, Luo-Rudy I and Luo-Rudy dynamic models, examining also how these ionic
models satisfy the principle of charge conservation and how to derive the so-called
restitution curve for the action potential duration of a given ionic model. Reduced
models, such as the minimal FitzHugh-Nagumo model are also presented, together
with their phase-plane analysis, bifurcation and frequency diagrams.
Chapter 3 presents mathematical models of periodic cardiac cells arrangements,
beginning with one-dimensional fibers, deriving the cable equation, showing a one-
dimensional homogenization technique and the main results on one-dimensional
traveling waves, namely traveling fronts for the bistable equation and traveling
pulses for the FitzHugh-Nagumo system with a diffusion term. We then move to
models of cardiac tissue in more dimensions, illustrating a two-scale homogeniza-
tion technique that allow us to derive an averaged Bidomain model, proving both
well-posedness results for the cellular and the averaged models and convergence
results based on -convergence techniques. We then present an heuristic derivation
of the anisotropic Bidomain model in both parabolic-parabolic and parabolic-elliptic
forms. Well-posedness results are derived using different techniques, such as time
semi-discretization, Faedo-Galerkin techniques, and fixed point arguments.
Chapter 4 presents the main reduced macroscopic cardiac models: the linear
anisotropic Monodomain model, Eikonal models (both Eikonal-curvature and
Eikonal-diffusion models), and the relaxed non-linear anisotropic Monodomain
model. These model are then given in dimensional form and some well-posedness
results are summarized. The chapter is concluded by a numerical comparison
between activation time maps computed with these reduced models and the full
Bidomain model.
Chapter 5 is devoted to the modeling of anisotropic cardiac sources, presenting
both the differential and integral formulations of the potential field. Approximate
representations of cardiac sources such as the heart surface and oblique dipole
source models are given, as well as the cardiac sources splitting into axial and
conormal components in both axially symmetric and orthotropic media. The
chapter concludes with a numerical example illustrating this source splitting and
its comparison with experimental results.
Chapter 6 briefly reviews the Inverse problem of Electrocardiology, in terms of
cardiac sources, in terms of wavefront and in terms of potential alone, presenting
the mathematical models of the cardiac electric sources and their numerical
approximations.
Chapter 7 presents the main numerical techniques employed in the space
and time discretizations of the Monodomain and Bidomain cardiac models. In
particular, we apply the finite element method in space and finite difference
methods in time. The latter can be fully implicit or semi-implicit, and can employ
Preface ix
decoupling techniques and operator splitting methods between the ordinary and
partial differential equation components of the cardiac models. The chapter is
concluded by a review of numerical methods for the eikonal–diffusion equation.
Chapter 8 is devoted to the construction and analysis of parallel solvers for the
discrete Bidomain systems arising at each time step of an implicit or semi-implicit
time discretization. Our parallel solvers are based on domain decomposition meth-
ods, more specifically on overlapping Schwarz methods, which provide scalable
preconditioners accelerated with a Krylov space iterative method such as PCG or
GMRES. After recalling the main results of the abstract Schwarz theory, we derive
scalable convergence rate bounds for two-level and multilevel additive Schwarz
preconditioners for the Bidomain system. We then present the results of several
numerical tests with these Schwarz preconditioners in additive, multiplicative and
hybrid form, showing their scalability on different parallel machines and inves-
tigating their performance with respect to the different discretization parameters.
We also present how these Schwarz preconditioners can be combined with block-
diagonal and block-factorized preconditioners suggested by the 22 block structure
of discrete Bidomain systems.
Chapter 9 illustrates how to apply the Bidomain and Monodomain solvers
developed in the previous chapters to simulate and study some of the most
important phenomena in cardiac electrophysiology. More precisely, we present
detailed simulations of: (1) the genesis of cardiac excitation and virtual electrode
phenomena, in particular anode/cathode make/break and strength-interval (S-I)
curves; (2) the anisotropic propagation of excitation and recovery fronts in three
dimensional domains; (3) the effects of cardiac heterogeneities (transmural and
apico-basal) on fronts propagation and APD distribution; (4) the morphology of
electrograms, in particular of the QRS complex and the T wave; (5) the computation
of excitation and repolarization time markers; (6) the presence of ischemic regions
and effects such as ST-segment depression and elevation; (7) the simulation of
cardiac reentry phenomena.
An Appendix lists some of the main cardiac simulation research projects,
software libraries, some related monographs and tables of physical units and
constants used in the book.
The authors would like to thank Prof. Bruno Taccardi for introducing them
to the field of Mathematical Physiology and for many stimulating discussions,
Prof. Alfio Quarteroni for his encouragement throughout this project, Dr. Fabrizio
del Bianco and Dr. Lara Charawi for their help in proofreading the manuscript.
xi
xii Contents
References .. .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 367
Index . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 395
Chapter 1
Basic Cardiac Anatomy and Electrocardiology
In this chapter, we briefly review the basic physiology and anatomy of the heart; a
more complete treatment with many more details can be found in standard textbooks
such as [189,267]. The heart is a double pump consisting of four chambers, two atria
in the upper part, separated by the interatrial septum and two ventricles in the lower
part, separated by the interventricular septum. Atria and ventricles are separated by
the atrioventricular septum, which contains the tricuspid valve in the right heart and
the mitral valve in the left heart. The right ventricle is connected to the pulmonary
artery via the pulmonary valve and the left ventricle is connected to the aorta via
the aortic valve, see Fig. 1.1 for a schematic view and [189] for more details. The
ventricular walls are externally lined by a thin connective layer called epicardium
and internally by the endocardium, with the cardiac muscle called myocardium in
between. The left ventricular wall is about three times thicker than the right one,
while the atrial walls are considerably thinner. The right heart functions as a pump
driving blood through the pulmonary circulation, to the lungs and then back to the
heart, while the left heart functions as another pump driving the oxygenated blood
through the systemic circulation around the body. These two main cardiac functions,
mechanical and haemodynamical, are driven and coordinated by the electrical
activity of the heart, which is the main focus of this monograph. In this chapter,
we will briefly review the essential features of the cardiac bioelectrical activity. The
main events in a normal cardiac cycle are often described by a Wiggers diagram,
see Fig. 1.2, showing the simultaneous time course of blood pressure (aortic, atrial,
ventricular), ventricular volume, electrocardiogram, and phonocardiogram during a
normal cardiac cycle.
cells and they can modulate both beat-to-beat and long-term variations of the
cardiac frequency. In normal conditions, SAN cells generate an action potential
that propagates throughout the right atrium and through Bachmann’s bundle to the
left atrium, stimulating the myocardium of both atria to contract. The activation
front reaches the atrioventricular node (AVN) located at the base of the atria
4 1 Basic Cardiac Anatomy and Electrocardiology
where the interatrial septum and interventricular septum meet. AVN cells have a
relatively slow conduction velocity and they are responsible for the major part of
normal conduction delays between atrial and ventricular contractions. Such delays
are properly timed in order to optimize the atrial pump activity and to protect
the ventricles from too early stimulation. The AVN conducts the action potential
through the nonexcitable atrioventricular septum, activating the specialized fibers
of the bundle of His and the Purkinje network that spread as a tree-like left and right
bundle branches ending on the endocardial surface of the ventricles. These Purkinje
terminations transmit the action potentials to the ventricular walls and cardiac
excitation then propagates throughout the ventricles. The electrical activation of the
left ventricle, the largest cardiac chamber, normally starts at the interventricular
septum, propagates towards the anteroapical region and reaches the posterobasal
region. The electrical activation of the right ventricle starts shortly after the left
ventricle activation; see e.g. [462].
Cardiac muscle presents specific features different from the other two main types
of muscle, skeletal and smooth muscles. The two fundamental features of cardiac
muscle are its cyclical activity of contraction/relaxation, where the individual
myocytes contract and relax according to a precise sequence and its homeostatic
regulation in order to maintain a relatively constant circulatory pressure in spite
of large flow variations. While skeletal myocytes are functionally separated and
individually innervated, cardiac myocytes are electrically connected in order to
form a single functional unit, known as cardiac syncytium. While skeletal muscle
strength is modulated mostly by recruiting a variable number of motor units, in the
cardiac syncytium recruitment is not possible and strength modulation relies on the
myocytes strength modulation. Moreover, while in skeletal muscle cells a sequence
of action potentials can generate an increasing force just by a temporal sum of single
pulses, in cardiac myocytes each cycle of contraction/relaxation is activated by a
single action potential that has a duration about 100 times longer.
In cardiac tissue are present three fundamental cell types with different structural
and functional characteristics: working, nodal, and conduction cells.
Working cardiomyocytes. These form the mass of atrial and ventricular tissues
and are responsible for force development. Their morphology is tubular (with
possible branches), with a length of about 50–150 m and a diameter of about
10–20 m. Myocytes volume and shape can be complex and variable, according
to the tissue region, species, developmental stage and disease processes. They are
enclosed by a lipid membrane, the sarcolemma, and contain one or more nuclei,
mitochondria, myofibrils, the sarcoplasmic reticulum, sarcomeres, the cytoskele-
ton anchoring the different organelles and an aqueous solution, the sarcoplasm,
filling the intracellular space. The sarcolemma is a semi-permeable barrier and
contains ion channels, pumps and exchangers that allow the inward and outward
1.2 Cardiac Tissue Organization 5
currents involved in the action potential, as well as other proteins involved in cell
adhesion and signalling. Cardiomyocytes exhibit a periodic structure with cross
striations formed by alternating segments of thick and thin protein filaments. Their
sarcolemma presents deep invagination along the Z-discs known as T-tubules, that
allow depolarization of the membrane to quickly penetrate to the interior of the cell
and play a critical role in the excitation-contraction coupling. Cardiomyocytes are
mechanically coupled at specialized region of the membrane known as intercalated
disks, which also include gap junction channels providing electrical coupling (see
below).
Nodal cells. These cells constitute the SA and AV nodes and are able to
autonomously activate due to the presence of specific ionic channels. Compared
with working myocytes, they are smaller, more tortuous, have limited contractile
activity and lack T-tubules. Nodal cells are richly innervated by parasympathetic
and sympathetic nervous system fibers, that make them susceptible to autonomic
influences. The SA and AV nodes also contain some transitional cells with
intermediate features between nodal and atrial myocytes.
Conduction cells. These are also called Purkinje cells and have a morphology
similar to working myocytes but with a larger diameter (up to 80 m), reduced
contractile proteins and a few T-tubules. They also have the ability of automaticity,
but at a slower rate than SA or AV node cells.
Table 1.1 reports the average values of diastolic potential, upstroke velocity,
overshoot, action potential duration (APD) and propagation velocity for different
mammalian cardiomyocytes. For detailed measurements of conduction velocities in
canine cardiac tissue, see [586, Table 16-1, p. 148].
Fibroblasts. Cardiac tissue is composed by myocytes that are arranged in
a complex three-dimensional fiber and laminar architecture, described below. In
areas between and surrounding cardiac myocytes there are cardiac fibroblasts, that
constitute the major non-myocyte cell population in the ventricles [206]. Although
cardiomyocytes occupy the largest volume fraction of the normal heart, they are
outnumbered by the smaller fibroblasts. These are responsible for the synthesis of
extracellular matrix proteins such as different types of collagen. Fibroblasts can
Table 1.1 Average values of diastolic potential, upstroke velocity, overshoot, action potential
duration (APD) and propagation velocity for different mammalian cardiomyocytes
Diastolic Upstroke Propagation
potential velocity Overshoot APD velocity
Cell types (mV) (mV/ms) (mV) (ms) (mm/ms)
SAN 50, 60 1–10 30 100–200 0.03–0.05
Atria 80 100–200 30 100–200 0.3–0.7
AVN 60, 70 5–15 20 100–300 0.1
His budle/Purkinje 80, 90 800 40 200–500 2–4
Ventricles 80 100–200 40 200–500 0.5–0.8 along
0.15–0.25 across
6 1 Basic Cardiac Anatomy and Electrocardiology
develop into myofibroblasts and both cell types are involved in the development of
fibrosis in injured cardiac tissue. The structural arrangement of fibroblasts is still not
well understood, but some authors have suggested that they are organized in sheets
that follow closely the myocardial sheets, with some fibroblasts forming strands that
bridge cleavage planes between sheets [206]. Cardiac fibroblasts are electrically
non-excitable cells, but they are efficient mechano-electrical transducers. The
possibility that fibroblasts may actively contribute to cardiac electrophysiology has
been considered only recently, see [77, 192, 330, 424].
Extracellular matrix and collagen. The extracellular matrix (ECM) is an
extensive and highly organized network of fibrous proteins, collagen and elastin,
and is the major determinant of passive mechanical properties of cardiac tissue. The
ECM plays important roles in muscle development, myoblast differentiation and
in maintaining functional integrity of the myocardium [576]. Collagen and elastin
have relative densities that can vary with tissue type, species and age. The ECM
structure differs at different cardiac locations. In the endomysium, the collagen
network surrounds single myocytes and interconnects neighboring myocytes. In
the perimysium, bundles of collagen fibers envelope groups of adjacent myocytes,
providing the laminar cardiac structure and long perimysal collagenous tendons can
link adjacent laminae. In the epimysium, layers of collagen and elastin are found at
the epicardium and endocardium.
Gap junctions and connexin. Cardiac myocytes are mechanically connected
by junctions anchored to the cytoskeleton, known as intercalated disks, and are
electrically connected by electrical synapses, known as gap junctions, connect-
ing myocytes mostly end-to-end (longitudinal gap junctions) but also laterally
(transverse gap junctions) [423]. Gap junctions are small ionic channels with low
selectivity, with a diameter of about 2 nm and length of about 2–12 nm. They are
composed of two adjoint connexons, located in the membranes of the two coupled
cells, and each connexon is formed by six channel proteins known as connexins
(Cx40, Cx43, Cx45). Cx43 is the main cardiac connexin and is mostly present in
ventricular tissue. Cx40 is present in atrial tissue and in the conduction system,
while Cx45 is present only in the conduction system. Longitudinal gap junctions
are prevalent in ventricular myocardium, yielding a macroscopic electrical coupling
that is anisotropic.
Cardiac stem cells. The human heart has been traditionally viewed as a
terminally differentiated postmitotic organ, with cardiomyocytes established in
number at birth and persisting throughout the lifespan of the organism. This view has
been recently changed by the discovery that cardiac stem cells live in the heart and
differentiate into the various cardiac cell lineages; see e.g. [41,46,190,289,296] and
the references therein. Cardiac stem cells regulate myocyte turnover and condition
myocardial recovery after injury. The initial enthusiasm triggered by the discovery
of endogenous cardiac stem cells was then followed by some controversy over
the nature of myocyte renewal in the human heart. Intense research studies are
underway to explore this new field of cardiac regeneration. Several categories of
cardiac progenitors have been described and diverse experimental strategies to
remuscularize the injured heart using adult stem cells and pluripotent stem cells,
1.3 Fiber and Laminar Architecture of Ventricular Myocardium 7
cellular reprogramming and tissue engineering are in progress. The novel results that
are rapidly appearing in the literature impose a reconsideration of the mechanisms
involved in myocardial aging and the progression of cardiac hypertrophy to heart
failure. Even if controversial experimental issues are still present, cardiac regener-
ation studies have great potentials and could soon lead to new understanding and
treatments of infarcted hearts and to restoration of myocardial mass and ventricular
function; see e.g. [190].
Fig. 1.4 Left: schematic representation of ventricular fiber architecture with orthonormal triplets
of vectors along fiber (red), and two orthogonal vectors (blue and black) on intramural, epi-, mid-,
endocardial sections. Right: schematic view of a few laminae (blue) on selected sections (outlined
in green) of the ventricular wall (outlined in red)
Cardiac cells are excitable cells, some autonomously (SAN pacemaker cells) and
some after a proper electrical stimulus. The excitation of a cardiac cell causes a
rapid variation of its potential difference across the cell membrane, the so-called
transmembrane potential. If the stimulus is below a certain threshold value, the
transmembrane potential quickly returns to its resting value after the stimulus ends.
On the other hand, if the stimulus is above threshold, the cell membrane depolarizes
and the transmembrane potential increases from the negative resting value to a value
around or above zero, remains around this value for a certain interval and then
returns to the resting value after the membrane repolarizes. This event is called an
action potential.
The main phases of a typical ventricular action potential are displayed in Fig. 1.5.
Phase 0. In this phase, myocytes undergo a rapid depolarization due to the
opening of the fast NaC channels. This causes a rapid increase in the membrane
1.4 Cardiac Action Potentials 9
0 + K+ (out)
Na (in)
IKs, IKr, IK1
I 3
Na
0
4
4
−84 mV
0 300 msec
Fig. 1.5 Schematic plot of a cardiac action potential with its phases 0 (depolarization or upstroke),
1 (peak or notch), 2 (plateau), 3 (repolarization or recovery), 4 (resting) and main ions and currents
involved in each phase
conductance and thus a rapid influx of NaC ions into the cell, the INa current,
through sodium channels formed by alpha-subunits with channel proteins of the
NAV 1:# family encoded by SCN genes (mostly NaV 1:5 channel protein encoded by
the SCN5A gene associated with Long QT and Brugada syndrome channelopathies)
and by beta-subunits with NaV ˇ# channel proteins encoded by SCN#B genes.
Phase 1 occurs with the inactivation of the fast NaC channels. The transient net
outward current causing the small downward deflection of the action potential is due
to K C and Cl ions carried by the Ito1 and Ito2 currents, respectively.
Phase 2 is the plateau phase sustained by a balance between inward movement
of Ca2C , the ICa current through L-type calcium channels (CaV 1:2 channel protein
encoded by CACNA1C gene) and outward movement of potassium ions K C , the IKs
current through the slow delayed rectifier potassium channels (KvLQT1 channel
protein encoded by the KCNQ1 gene). The sodium-calcium exchanger current
INa;Ca and the sodium/potassium pump current INa;K also play minor roles during
phase 2.
Phase 3. This is the “rapid repolarization” phase of the action potential, when
the L-type Ca2C channels close while the slow delayed rectifier (IKs ) K C channels
are still open. This ensures a net outward current, corresponding to negative change
in membrane potential, thus allowing more types of K C channels to open. These
are primarily the rapid delayed rectifier K C channels (IKr current, KV 11:1 channel
protein encoded by hERG or KCNH2 genes) and the inwardly rectifying K C current
IK1 . This net outward, positive current causes the cell to repolarize. The delayed
10 1 Basic Cardiac Anatomy and Electrocardiology
rectifier K C channels close when the membrane potential is restored to about 80 to
85 mV, while IK1 is maintained throughout phase 4, contributing to set the resting
membrane potential.
Phase 4 is the resting phase, during which the transmembrane potential remains
at the resting value of about 84 mV until it is stimulated by an external electrical
stimulus.
Cardiac action potentials differ significantly in different regions of the heart. Indeed,
action potentials from atrial, ventricular and specialized conduction myocytes
exhibit different waveforms, see Fig. 1.6. Moreover, isolated myocytes with differ-
ent action potential morphologies have been observed in the epicardial, midmyocar-
dial (M-cells), and endocardial regions (transmural heterogeneity) and also along
the apex-to-base direction of the ventricle (apex-to-base heterogeneity). When these
myocytes are embedded in the ventricular wall, their heterogeneous properties affect
the sequence of repolarization and the actual APD distribution in the entire wall
(spatial dispersion of APD).
Right and left ventricle heterogeneity. Right ventricular myocytes from dog
hearts show action potentials with shorter APD and deeper notches than in left
ventricular myocytes. Some authors have related these features to larger Ito and
IKs currents in the right ventricle and have suggested that arrhythmogenic gradients
could arise at the interventricular septum.
Fig. 1.6 Heterogeneity of cardiac action potentials from different regions of the heart
1.5 The Electrocardiogram (ECG) 11
The ECG is the registration of the extracellular potential on the body surface due to
the propagation of cardiac action potentials. The registrating electrodes are usually
placed at standard positions described below.
12 1 Basic Cardiac Anatomy and Electrocardiology
Einthoven leads. Given the potentials at the left arm ˚L , at the right arm ˚R and at
the left foot ˚F , Einthoven standard limb leads are defined as
VI D ˚L ˚R
VII D ˚F ˚R
VIII D ˚F ˚L :
By Kirchhoff’s law, we have VI C VIII D VII . Einthoven assumed that the heart
is located at the center of a homogeneous spherical conductor representing the
torso, and that the three vectors associated with these three leads form an equilateral
triangle in the vertical frontal plane of the body, see Fig. 1.7.
Augmented leads. In addition to the three Einthoven leads, three additional
leads, denoted by aV R ; aV L ; aV F and known as augmented leads, are defined on
the frontal plane in the directions that bisect the sectors associated with each pair of
Einthoven leads. For example, the augmented aV F lead is defined by
˚L C ˚R
VaV F D ˚F :
2
Fig. 1.7 Schematic view of Einthoven standard limb leads (left) and augmented leads (right)
1.5 The Electrocardiogram (ECG) 13
Fig. 1.8 The six precordial leads V1 ; : : : ; V6 in a frontal view (left) and horizontal cross-section
(right) of the thorax
In this way, we obtain six leads that divide the frontal plane into equal 30ı sectors,
see Fig. 1.7. In order to measure potentials close to the heart, six additional leads are
defined on a horizontal plane by using the Wilson central terminal
˚R C ˚L C ˚F
˚CT D :
3
Precordial leads. The six precordial (or chest) leads, denoted by V1 ; : : : ; V6 ,
are obtained by comparing the Wilson central terminal with unipolar electrode
readings taken from six different locations on the chest, see Fig. 1.8 and [444]
for more details. V1 and V2 are located at the fourth intercostal space on the
right and left side of the sternum, and the other four proceed around the left
chest just below the fourth rib, ending with V6 under the armpit. The 12 leads
I; II; III; aV R ; aV L ; aV F ; V1 ; : : : ; V6 form the most commonly used clinical ECG
system.
QRS complex. The subsequent QRS complex, usually the largest deflection
in the ECG, is associated with the complex activation sequence of the ventricles,
resulting in upward and downward deflections and multiple peaks in most of the
leads. The Q wave corresponds to the activation of the interventricular septum, the R
wave corresponds to the septum activation completion and the apex activation, the S
wave corresponds to the activation of the ventricular free walls and the basal region.
These three waves are not always detectable and can sometimes merge, yielding a
quite variable morphology of the QRS complex.
ST segment. The QRS complex in normal ventricular myocytes is followed by a
rather long period where all cells are depolarized during the plateau of the action
potential. This isoelectric interval, known as ST segment, lasts a few hundreds
1.5 The Electrocardiogram (ECG) 15
milliseconds, starting at the end of the QRS complex (the so-called J point) and
ending at the beginning of the T wave.
T wave. The T wave signals ventricular repolarization and its morphology
depends on the repolarization sequence through the ventricles. In normal hearts, the
start of the T wave is thought to correlate with the plateau ending of the epicardial
cells, the T wave peak with the full repolarization of the epicardium, the end of the
T wave with the repolarization of the midmyocardial M-cells.
QT interval. The QT interval is measured from the beginning of the QRS
complex to the end of the T wave. Since the QT interval is strongly frequency
dependent, a corrected QT interval
p QTc is sometimes defined, for example by
normalizing it as QTc D QT= RR.
U wave. After the T wave, it is sometimes present a smaller diastolic deflection
known as U wave. Some authors explain the U wave as due to repolarization of
the papillary muscles or Purkinje fibers, while others explain it as due to late
repolarization of M-cells or yet to afterpotentials caused by ventricular stretch.
Other important cardiac events too weak to be detected on the ECG include atrial
repolarization, SA nodal activation, AV nodal conduction and Purkinje network
propagation. An example of a normal 12-lead ECG recording is shown in Fig. 1.10.
We briefly review here some of the main ECG alterations. For a more complete
treatment see [189] and [317].
Normal sinus rhythm. The normal sinus rhythm of a healthy normal heart has
an ECG with all three P, QRS, T complexes normal, distinguishable and with a
frequency between 60 and 100 bpm (beats per minute).
Sinus bradycardia and tachycardia. A sinus rhythm of less than 60 bpm
is called sinus bradycardia and may be a consequence of increased vagal or
parasympathetic tone. A sinus rhythm of higher than 100 bpm is called sinus
tachycardia; most often it is due to physiological response to physical exercise or
psychical stress, but may also result from congestive heart failure.
Sinus arrhythmia. If the sinus rhythm is irregular such that the longest PP or RR
interval exceeds the shortest interval by 0.16 s, then it is called sinus arrhythmia, a
very common situation in all age groups and not considered a heart disease in young
people.
Nonsinus atrial rhythm and wandering pacemaker. If the origin of atrial
contraction is located somewhere else in the atria other than the sinus node (nonsinus
atrial rhythm) or wanders in the atria (wandering pacemaker), then the P waves can
have variable polarity and the PQ interval variable duration.
Paroxysmal atrial tachycardia (PAT). Paroxysmal atrial tachycardia describes
the condition when the P waves are a result of a reentrant activation front (circus
movement) in the atria, usually involving the AV node. This leads to a high rate of
activation, usually between 160 and 220 bpm. In the ECG the P wave is regularly
16 1 Basic Cardiac Anatomy and Electrocardiology
followed by the QRS complex. The isoelectric baseline may be seen between the T
wave and the next P wave.
Atrial flutter. When the heart rate is sufficiently elevated so that the isoelectric
interval between the T wave ending and the P wave beginning disappears, the
arrhythmia is called atrial flutter. The origin is also believed to involve a reentrant
atrial pathway. The frequency of these fluctuations is between 220 and 300 bpm.
The AV node and, thereafter, the ventricles are generally activated by every second
or every third atrial impulse (2:1 or 3:1 heart block).
Atrial fibrillation. The activation in the atria may also be fully irregular and
chaotic, producing irregular fluctuations in the baseline called atrial fibrillation.
A consequence is that the ventricular rate is rapid and irregular, though the QRS
contour is usually normal.
Ventricular tachycardia. Slower conduction in ischemic ventricular muscle that
leads to circular activation (reentry) causes activation of the ventricular muscle at a
high rate (over 120 bpm) and generates rapid, bizarre, and wide QRS complexes.
This arrythmia is called ventricular tachycardia and is often a consequence of
ischemia and myocardial infarction.
Ventricular fibrillation. When ventricular depolarization occurs chaotically, the
situation is called ventricular fibrillation. The ECG has coarse irregular deflections
without QRS-complexes. The cause of fibrillation is the establishment of multiple
re-entry loops usually involving diseased heart muscle. The contraction of the
ventricular muscle is irregular and is ineffective at pumping blood, with a result
of almost immediate loss of consciousness and death within minutes.
Atrioventricular blocks. In normal sinus rhythm, the P waves always precede
the QRS-complex with a PR-interval of 0.12–0.2 s. When the P wave always
precedes the QRS complex but the PR interval is prolonged over 0.2 s, first-degree
atrioventricular block is diagnosed. If the PQ interval is longer than normal and
the QRS complex sometimes does not follow the P wave, the atrioventricular block
is of second-degree. If the PR interval progressively lengthens, leading finally to
the dropout of a QRS complex, the second degree block is called a Wenkebach
phenomenon. Complete lack of synchronism between the P wave and the QRS
complex is diagnosed as third-degree (or total) atrioventricular block.
Bundle-branch blocks. In right bundle-branch block (RBBB), the right bundle-
branch is defective so that the electrical impulse cannot travel through it to the
right ventricle, but activation reaches the right ventricle by proceeding from the
left ventricle. This can be seen in the ECG as a QRS complex wider than 0.1 s, or
a broad terminal S wave in lead I, or a double R wave in lead V1. The situation in
left bundle-branch block (LBBB) is similar, but activation proceeds in a direction
opposite to RBBB. The ECG has a broad and tall R wave, usually in leads I, aVL,
V5 or V6.
Wolff-Parkinson-White (WPW) syndrome. In this syndrome, the QRS com-
plex initially exhibits an early upstroke called the delta wave. The interval from the
P wave to the R spike is normal, but the early ventricular excitation forming the
delta wave shortens the PQ time. The cause of the WPW syndrome is the passage of
activation from the atrium directly to the ventricular muscle via an abnormal route,
1.6 Cardiac Imaging 17
called the bundle of Kent, which bypasses the AV junctions. This activates part of
the ventricular muscle before normal activation reaches it via the conduction system
(after a delay in the AV junction), a process called pre-excitation.
Atrial hypertrophy. Right atrial hypertrophy is a consequence of right atrial
overload, due to tricuspid valve disease (stenosis or insufficiency), pulmonary valve
disease, or pulmonary hypertension. An unusually large (i.e. 0.25 mV) P wave is
seen in leads aVF and III. Left atrial hypertrophy is a consequence of left atrial
overload, due to mitral valve disease (stenosis or insufficiency), aortic valve disease,
or hypertension in the systemic circulation. The P wave exhibits two phases with the
same polarities in lead I but opposite polarities in V1. This typical P wave form is
called the mitral P wave.
Ventricular hypertrophy. Right ventricular hypertrophy is a consequence of
right ventricular overload, due to pulmonary valve stenosis, tricuspid insufficiency,
pulmonary hypertension or many congenital cardiac abnormalities, such as a
ventricular septal defect. The ECG shows a tall R wave of 0.7 mV in lead V1. Left
ventricular hypertrophy is a consequence of left ventricular overload, due to mitral
valve disease, aortic valve disease, or systemic hypertension. The ECG has a tall R
wave in leads I, III, V5, V6 and has a tall S wave in leads III, V1 and V2.
Myocardial ischemia and infarction. The decreased transport of oxygen to the
cardiac muscle due to occlusion of a coronary artery is called ischemia. Ischemia
causes changes in the resting potential and in the repolarization of the muscle cells,
which are seen as changes in the T wave. If the oxygen transport is terminated in a
certain area, the heart muscle dies in that region. This is called an infarction.
As part of the rapidly evolving fields of medical and biological imaging, also the
field of cardiac imaging is currently experiencing tremendous progresses that are
producing better imaging techniques seeking to reveal, diagnose, or examine both
cardiac diseases and the normal cardiac anatomy and physiology. Among the many
techniques available, we mention the following ones and refer the interested readers
to some of the recent monographs [83,291,475,554] and reviews [2,83] in the field.
Angiocardiography is a technique for radiographic examination of the heart
chambers and thoracic veins and arteries. A liquid radiocontrast agent, typically
containing iodine, is injected into the bloodstream, then the tissues are examined
using X-rays. To avoid dilution, the radiopaque material is typically introduced with
a catheter, a process known as selective angiocardiography. The X-ray image is
normally captured on high speed serial media that allow the motion to be observed,
such as a 35 mm film. The process requires fasting before the test, with a sedative
and an antihistamine being administered before the test. Angiocardiography can be
used to detect and diagnose congenital defects in the heart and adjacent vessels. The
use of angiocardiography has declined with the introduction of echocardiography.
18 1 Basic Cardiac Anatomy and Electrocardiology
In this chapter, we briefly review the main mathematical models used in cellular
electrophysiology, and we refer to [212, 273] for a more complete treatment and
additional topics.
In general, the flux across the membrane of the generic ion K with valence z is
the sum of two contributions, the diffusion flux Jdiff and the electric flux Jelect , all
measured in mol cm2 s1 . The constitutive law describing this total flux is known
as the Nernst-Planck equation.
+
[K ]i [K+]
e
intracellular extracellular
media J media
K,diff
J
K,elect
membrane
Due to the ion motion by diffusion, down the concentration gradient, the
diffusion flux Jdiff satisfies Fick law:
Jdiff D Drc;
where D is the diffusion coefficient (cm2 s1 ) and c is the concentration (mol cm3).
Due to the ion motion by electric field, the electric flux Jelect satisfies the Planck
equation:
z
Jelect D cru;
jzj
where u is the electric potential (V ), isthe mobility of the ion (cm2 V1 s1 ), z is
C1 for positive ions
jzj D
z
the valence of the ion, so that
1 for negative ions:
We note that if the concentration gradient is zero then Jdiff D 0, while if the
potential gradient is zero then Jelect D 0.
In both the electric and the diffusion fluxes, the ions collide with the solvent
and therefore the diffusion coefficient D and the ion mobility should be related.
Indeed, Einstein [159] found that this relationship is given by
RT
DD ; (2.1)
jzjF
where R is the gas constant .8:314 J=.K mol//, T the absolute temperature .K/,
F D e NA the Faraday’s constant .9:6485e C 4 C=mol/, e the elementary charge
and NA the Avogadro’s number, see Tables A.1 and A.2. Then it follows that the
total ionic flux J is given by
z
J D Jdiff C Jelect D Drc cru;
jzj
2.3 The Goldman-Hodgkin-Katz (GHK) Current-Voltage Relation 23
Let us assume now that the ion flux J and the electric potential u are transverse to
the cell membrane, which extends from x D 0 (inside) to x D L (outside), and the
diffusion coefficient D is constant. Moreover, if we are at the steady state @t c D 0,
thus the conservation law
@t c C @x J D 0
implies that J is constant. Hence, we can reduce Eq. (2.2) to the one-dimensional
problem
dc zF du J
.x/ C .x/c.x/ C D 0: (2.3)
dx RT dx D
Equation (2.3) is clearly a linear differential equation in c.x/, whose solution can
be obtained in terms of u.x/ and J by multiplying by
Z
zF x du
exp .s/ds
RT 0 ds
and integrating on .0; x/, yielding
Z x
zF J zF
c.x/ D exp .u.x/ u.0// c.0/ exp .u.s/ u.0// ds :
RT D 0 RT
(2.4)
Using the standard notation with indexes i; e denoting intra- and extracellular
quantities, respectively, we then obtain the Nernst equation for the equilibrium
(reversal) potential
i
RT c
vK WD ui ue D log e :
zF c
2.5 Thermodynamical Derivation of the Nernst Potential 25
This derivation is based on the constitutive Nernst-Planck law for the total flux.
For a thermodynamical derivation from electrochemical potentials that shows the
universal character of the Nernst potential, see the next Section.
For ventricular myocytes, the typical intra- and extracellular concentrations and
Nernst potential values are given in Table 2.1.
We recall that the first law of Thermodynamics states that for a closed system the
change dU of internal energy equals the difference between the amount of heat
dQ supplied to the system and the amount of external work dW performed by the
system,
dU D dQ dW:
Since the work performed by the system can be written in terms of the increase dV
in its volume V and the external pressure p as dW D pdV, the first law can also be
written as
dU D dQ pdV:
The second law of Thermodynamics states that the change dS of the entropy S of
the system is
dS D dSrev C dSirr ;
where dSrev is the change due to the reversible interaction of the system with its
environment and dSirr is the irreversible internal change within the system that is
always 0. Since dSrev D dQ=T with T the absolute temperature, this law can also
be written as
dSirr D dS dQ=T:
26 2 Mathematical Models of Cellular Bioelectrical Activity
G D U C pV TS;
i.e. as the difference between the enthalpy H D U C pV and the product TS.
Differentiating and using the first and second laws of Thermodynamics, we obtain
For reversible processes, dSirr D 0, hence the second law implies dQ D TdS and
the first law
dU D TdS pdV;
where S; V and the other components nj , with j ¤ i , are kept constant. Hence,
the first law of Thermodynamics becomes
X
dU D TdS pdV C i dni ;
i
@U
where i D is the chemical potential of the i -th component, and the
@ni S;V;nj
Gibbs free energy becomes
X
dG D Vdp SdT C i dni : (2.6)
i
This last equation shows that the chemical potential for the i -th component can also
@G
be defined as i D .
@ni T;p;nj
For charged ions with valence zi , the electrochemical potential is defined as
Q i D i C zi Fu; (2.7)
The chemical potential g for an ideal gas can be derived using the equation
of state pV D nRT, relating the pressure p, volume V , number of moles n, gas
constant R, temperature T . If we consider n D 1 mol of ideal gas expanding at
constant temperature, then Eq. (2.6) gives us
dp
dG D Vdp D RT ;
p
and by integration
p
G D G0 C RT log ;
p0
where p0 is a reference pressure and G0 the associated Gibbs free energy. Since for
a mole of pure substance the chemical potential is given by the Gibbs free energy,
the last equation can be rewritten as
p
g D 0g C RT log ; (2.8)
p0
with 0g the chemical potential of the gas at the reference pressure p0 . If we take the
reference pressure equal to 1 atmosphere, Eq. (2.8) becomes
For an ion species in solution, Eq. (2.10) generalizes to the case of an electrochemi-
cal potential Q s as in (2.7)
The Nernst equilibrium potential can be derived from this equation by considering
an ion species s in equilibrium across a membrane and denoting by csi ; cse the
intracellular and extracellular concentrations, respectively. Since at equilibrium we
must have Q is D Q es , Eq. (2.11) yields
so that
csi
zs F .usi use / D RT log ;
cse
RT ci
i.e. the Nernst equilibrium potential vs D usi use D log se :
zs F cs
We now suppose to have two types of ions, a cation S1 and an anion S2 , with
concentrations c1 and c2 , valences z1 D z > 0 and z2 D z < 0. We consider
the flux through a channel ˝ connecting the intra- and extracellular neutral spaces.
The Nernst-Planck equations for the fluxes J1 ; J2 of S1 and S2 state that
zF
J1 D D1 rc1 C c1 ru ;
RT
zF
J2 D D2 rc2 c2 ru ;
RT
@t c1 C div J1 D 0 in ˝;
@t c2 C div J2 D 0 in ˝:
div D D in ˝;
Fz
u D .c1 c2 /:
a
2.6 Electrodiffusion Models: The Poisson-Nernst-Planck (PNP) Equation 29
We denote by S the lateral surface of the channel ˝ and by @˝i ; @˝e the two
channel openings in contact with the intra and extracellular media, respectively.
Then, we make the following additional assumptions, yielding boundary conditions
for the ionic concentrations, fluxes and the electric potential:
(a) The cation and anion concentrations on both sides of the membrane are
constant, so that the intra- and extracellular spaces are in electrical equilibrium
(b) Since the dielectric constants a of the channel’s solution and m of the
membrane satisfy m a , we assume that the channel wall is electrically
insulated,
(c) The channel wall prevents the movement of ions toward the lipidic regions,
nT Ji jS D 0; nT Je jS D 0I
(d) The electric potentials at the channel openings in contact with the extracellular
and intracellular media are constant and equal to ui and ue , respectively, but,
since the potential is given up to an additive constant, we set
uj@˝i D ui ue D v; uj@˝e D 0:
We now consider the steady-state of the 1-D model, assuming that the ion
concentrations are constant across the channel section. Collecting the previous
equations and boundary conditions, we find the one-dimensional Poisson-Nernst-
Planck (PNP) system:
8
ˆ zF
ˆ
ˆ @t c1 C @x J1 D 0 with J1 D D1 @x c1 C c1 @x u
ˆ
ˆ RT
ˆ
ˆ
ˆ
ˆ zF
ˆ
ˆ @t c2 C @x J2 D 0 with J2 D D2 @x c2 c2 @x u
ˆ
ˆ RT
ˆ
ˆ
ˆ
< @2 u D zF .c1 c2 /
xx
a
ˆ
ˆ with boundary conditions:
ˆ
ˆ
ˆ
ˆ D t/ D c i c1 .L; t/ D c2 .L; t/ D c e
ˆ
ˆ
c1 .0; t/ c2 .0;
ˆ
ˆ u.0; t/ D v u.L; t/ D 0
ˆ
ˆ
ˆ
ˆ
ˆ and initial conditions:
:̂
u.x; 0/ D 0 c1 .x; 0/ D c2 .x; 0/ D 0:
c1 c2
y D x=L; D uzF=RT; pD ; nD i ;
Cc
ci e c C ce
L L c i;e vzF
Jp D J1 ; Jn D J2 i ; O
c i;e
D ; vO D ;
.c i C c e /D1 .c C c e /D2 ci C ce RT
2.6 Electrodiffusion Models: The Poisson-Nernst-Planck (PNP) Equation 31
so that
d2 zFL2 d 2 u z2 F 2 L2 i
D D .c C c e /.p n/;
dy2 RT dx2 RTa
dp d
Jp D Cp ;
dy dy
dn d
Jn D n :
dy dy
dJ p dJ n
The first two equations of (2.12) require that D D 0, i.e. that Jp and Jn
dy dy
are constant. Therefore, the dimensionless one-dimensional stationary PNP system
becomes:
find fp.x/; n.x/; .x/g and constants fJp ; Jn g such that
8
ˆ dp d
ˆ
ˆ Jp D Cp ;
ˆ
ˆ dy dy
ˆ
ˆ
ˆ
ˆ dn d
ˆ
< Jn D n
dy dy
d 2 (2.13)
ˆ
ˆ D 2 .p n/;
ˆ
ˆ
ˆ
ˆ dy2
ˆ
ˆ
ˆ with boundary conditions:
:̂
p.0/ D n.0/ D cOi ; p.1/ D n.1/ D cOe ; .0/ D vO ; .1/ D 0;
z2 F 2 L2 i
with 2
D .c C c e /.
a RT
Since no analytical solutions of this system are known, we will look for
approximate solutions obtained by perturbation techniques with respect to the
parameter .
d2
D0
dy2
32 2 Mathematical Models of Cellular Bioelectrical Activity
d
implies that the electric field E D is constant and the potential .y/ is linear,
dy
so from the boundary conditions we obtain
d
.y/ D vO vO y; D Ov:
dy
dp
Jp D vO p:
dy
Integrating between 0 and y and using the left boundary condition p.0/ D cOi , we
obtain
1 e vOy
p.y/ D cOi e vOy C Jp :
vO
dn
Jn D C vO n;
dy
1 e Ovy
n.y/ D cOi e Ovy Jn ;
vO
cOi cOe e vO
Jn D Ov :
1 e vO
Returning to the original variables, we then determine the dimensional fluxes J1 ; J2 ,
hence the ionic density currents
zF
z2 F 2 D1 c i c e e RT v
I1 D zFJ 1 D v zF
;
RT L 1 e RT v zF
z2 F 2 D2 c c e
i e RT v
I2 D zFJ 2 D .v/ zF :
RT L 1 e RT v
The reduced problem obtained in the limit case for D 0 implies that p.y/ D n.y/,
so by adding or subtracting the equations for Jp and Jn in (2.13) we have
8
ˆ
ˆ dp dn dp
ˆ
ˆ Jp Jn D C D2 ;
< dy dy dy
d d d (2.14)
ˆ
ˆ Jp C Jn D p Cn D 2p
ˆ dy dy dy
:̂
p.0/ D n.0/ D cOi ; p.1/ D n.1/ D cOe ; .0/ D vO ; .1/ D 0:
dp
The first equation requires to be constant and from the boundary conditions we
dy
obtain p.y/ D cO C .cO cO /y, therefore system (2.14) becomes
i e i
8
ˆ
ˆ Jp C Jn
< D .cOe cOi /
2 (2.15)
ˆd JO JO
:̂ D D i ;
dy p.y/ cO C .cOe cOi /y
JO cOe
Then the boundary condition .1/ D 0 gives us the condition vO C log D
cOe cOi cOi
Ov
0, that determines JO D .cOe cOi / , which coupled with (2.15) implies
log ccOOi
e
8
ˆ
ˆ Ov
ˆ Jp D .cOe cOi /.
< e C 1/;
log ccOOi
ˆ Jn D .cOe cOi /. Ov e 1/:
ˆ
:̂
log ccOOi
34 2 Mathematical Models of Cellular Bioelectrical Activity
ci C ce ci C ce
Returning to the dimensional fluxes J1 D D1 Jp ; J2 D D2 Jn , we
L L
c Cc
i e
c C ce
i
find that the dimensional currents I1 D zF D1 Jp ; I2 D zF D2 Jn
are L L
!
zF e v zFD1 c e c i
I1 D .c c i /D1
RT ce
1 D .v v1 /;
L zF log c i
L v1
! (2.16)
zF e v zFD2 c e c i
I2 D .c c /D2 RT
i
c e 1 D .v v2 /;
L zF log c i
L v2
RT ce
where vk D log i ; k D 1; 2 are the Nernst potentials of the two ions
zk F c
considered.
These linear current-voltage relations are derived using the approximation of a
constant diffusive gradient.
RT i
PNa cNa C PK cK
i
vr D log e :
F e
PNa cNa C PK cK
GNa vNa C GK vK
vr D :
GNa C GK
Analogously, if we consider the three ions NaC ; K C ; Cl with GHK channels, we
have
RT i
PNa cNa C PK cK
i
C PCl cCl
i
vr D log e ;
F e
PNa cNa C PK cK
e
C PCl cCl
Since the cell membrane separates charges that accumulate at its intra- and
extracellular surfaces, it can be viewed as a capacitor. The capacitance is defined
as the ratio between the charge Q across the capacitor and the voltage potential
drop v necessary to hold the charge
Q
Cm D :
v
dQ dv
Icap D D Cm :
dt dt
The cellular membrane is modeled as a capacitor in parallel with a resistor (ionic
current), as shown in Fig. 2.1, so by the current conservation law, the transmembrane
current given by the sum of the capacitative and ionic currents, must be equal to the
applied current Iapp
dv
Cm C Iion D Iapp : (2.17)
dt
The structure of the total ionic current will be described by the specific ionic
membrane model adopted. In order to discuss these models, we need to introduce
the formalism for modeling ion channel gating. For more properties about ionic
channels of excitable membranes see [228].
36 2 Mathematical Models of Cellular Bioelectrical Activity
In general, for a given transmembrane potential drop, the ionic current through a
population of ionic channels is given in a unit area of membrane surface and can be
modeled as a product
where g.v; t/ is the proportion of open channels in a unit area of the membrane
surface and .v/ is the current-voltage (I-V) relation of a single open channel. In
the previous sections on electrodiffusion models, we have seen different models for
.v/. For example, in the long channel limit (2.16), we have .v/ D gc .v vr /,
where gc and vr are the channel conductance and resting potential, respectively.
The proportion of open channels in a unit area of the membrane surface can be
N Ntot N
written as g.v; t/ D D , where N is the number of open channels
S S Ntot
on the membrane surface, Ntot is the total number of membrane channels, S is
the membrane surface area. Hence, in the long channel limit we have Iion D
Ntot gc
G c w.v vr /, where G c D is the maximal channel conductance per unit
S
2.8 Ion Channel Gating 37
N
area of the membrane surface and w D is the percentage of open channels. We
Ntot
now show how to model this percentage w, also known as gating variable, starting
with the simplest case of a single-unit two-state channel.
One unit protein with two states. Consider a membrane portion of unit area
containing a given type of ionic channels that behave independently and that assume
two states S0 (close) or S1 (open), with transition rate constants ˛; ˇ (in general
dependent on the transmembrane potential v) as indicated in the first order kinetic
diagram
˛
close state D S0 S1 D open state :
ˇ
If S0 .t/; S1 .t/ are the average numbers of channels that at time t are in the state
S0 ; S1 , respectively, then
8
< dS1 .t/
D ˛ S0 .t/ ˇ S1 .t/
: S dt
1 .t/ C S0 .t/ D S:
dw
D ˛ .1 w/ ˇ w :
dt
By setting w1 D ˛
˛Cˇ
and w D 1
˛Cˇ
, this equation can be rewritten as
dw w1 w
D ; (2.18)
dt w
where w1 and w are the equilibrium state and the time constant, respectively.
By voltage-clamp or patch-clamp techniques, it is possible to apply an external
current that balances the membrane current in such a way that the transmembrane
potential v is kept at a fixed value in time. In this case, the capacitative current
38 2 Mathematical Models of Cellular Bioelectrical Activity
Icap D Cm dv
dt
is zero and in the circuit model equation (2.17) in the long-channel
limit we have Iion .t; v/ D G w.t; v/ .v vr / D Iapp ; hence the membrane
conductance can be obtained as
Iapp .t; v/
G w.t; v/ D :
v vr
For each fixed value of v, these experimental curves can be plotted in time and
compared with the theoretical curves predicted by solving the gating equation (2.18)
in the voltage-clamp setting: if at time t D 0 v is increased from 0 to v0 and held
fixed at such value, the time behavior is given by
see Fig. 2.2. If the experimental conductance curves do not agree with these
exponential curves, then the single-unit two-state channel model is not adequate and
more complex models must be considered. Such models are introduced in the next
sections and indeed they will be needed to correctly describe the ionic currents in the
membrane models developed in the literature, starting with the classical Hodgkin-
Huxley model presented in Sect. 2.9.1.
0.9 v = +15
0
0.8
v = −25
0
0.7
0.6 v = −45
0
W
0.5
0.4
0.3 v0 = −65
0.2
0.1
0
0 20 40 60 80 100
TIME
Fig. 2.2 Plots of voltage-clamp solutions (2.19) on the interval Œ0 T and (2.20) on ŒT 100, with
T D 50 ms, for v0 D 65; 45; 25; 15 mV
2.8 Ion Channel Gating 39
Two subunits with two states. Analogously, if we consider ionic channels with
two equal and independent subunits, which can open and close according to the
following kinetic diagram
2˛ ˛
close state S0 S1 S2 D open state ;
ˇ 2ˇ
we have
8 ds
ˆ
ˆ
0
D ˇs1 2˛s0
ˆ
ˆ dt
ˆ
ˆ
ˆ
ˆ
<
ds1
D ˇs1 C 2˛s0 ˛s1 C 2ˇs2
ˆ
ˆ dt
ˆ
ˆ
ˆ
ˆ
ˆ ds2
:̂ D ˛s1 2ˇs2 :
dt
Two different subunits. We consider now two subunits of different type, that we
will denote by m and h. The transition rate constants are ˛; ˇ for m and ; ı for h,
as indicated in the following kinetic diagram
2α α
S S10 S
00 20
β 2β
γ δ γ δ γ δ
α α
S S S
01 11 21
β 2β
dm dh
D ˛.1 m/ ˇm; D .1 h/ ıh: (2.21)
dt dt
The first mathematical model that describes accurately the action potential wave-
form and the associated permeability changes was proposed by Hodgkin and Huxley
in [229]. Although the Hodgkin-Huxley (HH) model was developed specifically for
nerve action potentials, the ideas and the mathematical formalism have been used
later in a number of models describing cardiac action potentials, see the survey
papers [356, 443] and also [141, 215, 282].
2.9 Cardiac Action Potential Models 41
8
ˆ dv
ˆ
ˆ Cm C Iion .v; m; h; n/ D Iapp
ˆ
ˆ dt
ˆ
ˆ dm m1 .v/ m
ˆ
ˆ D
<
dt m .v/
ˆ dh h1 .v/ h
ˆ
ˆ D
ˆ
ˆ dt h .v/
ˆ
ˆ
ˆ dn n1 .v/ n
:̂ D
dt n .v/
where the ionic current is the sum of sodium, potassium and leakage currents
Iion D INa C IK C IL :
Each of these currents has a linear current-voltage structure as in (2.16), with three
equal activation and one inactivation independent subunits for the sodium channel
INa D GN Na m3 h .v vNa /;
IK D GN K n4 .v vK /;
IL D GN L .v vL /:
Here GN Na ; GN K ; GN L are the maximal conductances and vNa ; vK ; vL are the Nernst
potentials of each channel type, respectively.
We recall that the coefficients of the gating variables equations are given by
˛m 1
m1 D ; m D ;
˛m C ˇm ˛m C ˇm
˛h 1
h1 D ; h D ;
˛h C ˇh ˛h C ˇh
˛n 1
n1 D ; n D ;
˛n C ˇn ˛n C ˇn
42 2 Mathematical Models of Cellular Bioelectrical Activity
with
0:1.25 v/ v
˛m D ; ˇm D 4 exp ;
vv/ 1
expŒ0:1.25 18
1
˛h D 0:07 exp ; ˇh D ;
20 expŒ0:1.30 v/ C 1
0:01.10 v/ v
˛n D ; ˇn D 0:125 exp :
expŒ0:1.10 v/ 1 80
The ionic current through channels of the membrane is modulated by the transmem-
brane potential v D ui ue , by gating variables w WD .w1 ; : : : ; wM / and by ionic
intracellular concentration variables c WD .c1 ; : : : ; cS /. In the membrane models
considered in the remainder of this chapter, the ionic current has the following
general structure
X
N Y
M
pj
Iion .v; w; c/ D Gk .v; c/ wj k .v vk .c// C In .v; w; c/;
kD1 j D1
where Icj is the sum of ionic currents carrying ion cj , Acap is the capacitive
membrane area, Vcj is the volume of the compartment where cj is updated, zcj
is the valence of ion cj and F is the Faraday constant.
The transmembrane current Im is the sum of the capacitive current, associated
with the membrane lipidic bilayer and of the ionic current Iion . Since Im must
balance the applied current Iapp (see e.g. [250]), then the evolution of the trans-
membrane potential of a single myocyte is given by the following system of ODEs
8
ˆ dv
ˆ
< Cm dt C Iion .v; w; c/ D Im D Iapp ;
ˆ
dw dc (2.24)
ˆ dt R.v; w/ D 0;
ˆ
dt
S.v; w; c/ D 0;
:̂
v.0/ D v0 ; w.0/ D w0 ; c.0/ D c0 ;
where Cm , Iion , and Iapp are the surface capacitance, the ionic current of the mem-
brane and the applied current per unit area of the membrane surface, respectively.
In this work, we will focus on cardiac ionic models of ventricular cells. Among the
many other cardiac ionic models, we mention the Punkinje fiber models by Noble
[355], McAllister-Noble-Tsien [323], DiFrancesco-Noble [149], the Sinoatrial node
(SAN) models by Zhang et al. [585], Severi et al. [472], and the references therein,
the atrial models by Nygren et al. [360], Kneller et al. [280] and the references
therein. For many additional models, we refer e.g. to Pullan et al. [406].
The first ventricular membrane model was formulated by Beeler and Reuter in
1977 [33]. This model, as the McAllister-Noble-Tsien model [323] developed in
1975 for Purkinje fibers, relied on the Hodgkin-Huxley formalism and, similarly
to the Hodgkin-Huxley model, assumed that sodium and potassium intracellu-
lar concentrations (ŒNaC i , ŒK C i ) remain constant during the action potential.
However, in cardiac myocytes entry of Ca2C through the L-type calcium channel
ICa.L/ produces a significant change in the calcium intracellular concentration
ŒCa2C i , mainly by triggering Ca2C release from the sarcoplasmic reticulum (SR)
via the calcium-induced calcium-release (CICR) process. A first attempt to take into
account the ŒCa2C i dynamics necessary to reproduce action potential morphology
was introduced in the earliest ventricular cell models, such as the Beeler-Reuter
44 2 Mathematical Models of Cellular Bioelectrical Activity
[33] and the Luo-Rudy phase I [308], but only at a phenomenological level. More
biophysically detailed ŒCa2C i models were introduced in later models, see e.g.
[254, 306, 405, 473].
Changes in ŒNaC i and ŒK C i can also influence the action potential morphology
when cells are paced at a fast rate. The first model incorporating detailed information
regarding ŒNaC i and ŒK C i dynamic concentration changes was the DiFrancesco-
Noble model [149] of the Purkinje fibers. Rasmusson et al. [418] developed a
similar model for a bullfrog atrial cell. The Luo-Rudy dynamic (LRd) model
[309, 310] of the guinea pig ventricular action potential formulated these processes
for the ventricular myocytes. These models were founding members in a new
class of second generation models that account for dynamic ion concentration
changes. Several such models for ventricular myocytes in different species have
been published since that time (see Table 2.2) and in particular Rudy’s group has
recently published a detailed ionic model for human ventricular myocytes, see [361].
Table 2.2 Some of the most used ventricular ionic models for different species and dimensions
of the associated dynamical systems of differential equations
Reference Species Equations
Beeler and Reuter [33] Mammalian 8
Ebihara and Johnson [154] Chick embrio
Luo and Rudy [308] Guinea pig 8
Noble et al. [357] Guinea pig 17
Luo and Rudy [309, 310] Guinea pig 19
Fenton and Karma [170] Mammalian 3
Jafri, Rice and Winslow [254] Guinea pig 31
Noble et al. [358] Guinea pig 22
Priebe and Beuckelmann [402] Human 22
Winslow et al. [563] Canine 33
Clancy and Rudy [93, 94] Guinea pig
Pandit et al. [367] Rat 26
Puglisi et al. [405] Rabbit
Bondarenko et al. [53] Mouse 41
ten Tusscher et al. [521] Human 17
Hund and Rudy [243] Canine 29
Shannon et al. [473] Rabbit 46
Cortassa et al. [136] Mammalian 50
Livshitz and Rudy [306] Canine 18
Niederer and Smith [351] Rat
Bueno-Orovio, Cherry and Fenton [65] Human 4
Mahajan et al. [315] Rabbit 26
Li et al. [299] Mouse 36
O’Hara and Rudy [361] Human 41
2.9 Cardiac Action Potential Models 45
More recently, detailed ionic models have been coupled with detailed signaling
pathways models; see for example [224] for a simulation study of beta-adrenergic
signaling and its whole-cell effects in ventricular myocytes by incorporating
receptor isoforms, multiple pathways and phosphorylation.
In 1977, Beeler and Reuter developed the first mathematical model of a mammalian
ventricular muscle cell (see [33]) based on the Hodgkin-Huxley formalism. The
ionic current Iion is given by the sum of N D 4 currents
The fast inward sodium current INa is primarily responsible for the rapid upstroke
of the action potential, while the other currents determine the configuration of the
plateau and repolarization phases. The slow inward current Is , primarily carried by
calcium ions, influences the duration of the action potential. The time-dependent and
time-independent outward potassium currents Ix1 and IK1 are instead responsible
for the repolarization phase. It was also discovered that the intracellular calcium
concentration changes significantly during the course of an action potential, hence
this was included in the model as an additional dynamic variable.
Fast inward sodium current INa . In addition to the activation gate m and the
inactivation gate h of the Hodgkin-Huxley model, Beeler and Reuter added a slow
inactivation gate j . All the three gating variables follow the dynamics governed by
Eq. (2.22). The expression of the sodium current is given by
where ENa is the sodium reversal potential (50 mV), gN Na is the maximal sodium con-
ductance (0:04 mS=mm2) and gN NaC is the constant background sodium conductance
(3 105 mS=mm2).
Slow inward current Is . This current is controlled by an activation gate d and
an inactivation gate f , both following the dynamics of (2.22). The magnitude of the
current is given by
Is D gN s df .v Es /;
d ŒCa2C i
D 0:01Is C 0:07.104 ŒCa2C i /:
dt
Time-dependent outward potassium current Ix1 . As in the Hodgkin-Huxley
model, this current is controlled by a single gating variable x1 , following (2.22).
The expression of the current is
exp.0:04.v C 77// 1
Ix1 D 8 103 x1 :
exp.0:04.v C 35/
Time-independent outward potassium current IK1 . The experimental evi-
dence suggested the presence of a time-independent background potassium current
with magnitude given by
4.exp.0:04.v C 85// 1/
IK1 D 0:0035 C
exp.0:08.v C 53// C exp.0:04.v
C 53//
0:2.v C 23/
0:035 :
1 exp.0:04.v C 23//
In 1991, Luo and Rudy developed a new mammalian ventricular membrane model
[308], based on the Beeler-Reuter model, adjusting some parameters to reproduce
more recent experimental results and including additional potassium currents. In the
LR1 model, the ionic current Iion is given by the sum of N D 6 currents
two inwards (INa , Isi ) and four outwards (IK , IK1 , IKp , Ib ). The first three currents
depend on six gating variables and one ion (intracellular calcium) concentration,
while the last three are time-independent. The reversal potentials are set to the
Nernst potentials for the ions involved, allowing the reversal potentials to vary
according to ionic concentrations gradients. Figure 2.3 reports the time evolution
of the transmembrane potential v, the gating variables m, h, j , d , f , X and the
intracellular calcium concentration ŒCa2C i .
Fast inward sodium current INa . The expression of this current is the same as
in the Beeler-Reuter model with the omission of the constant background sodium
conductance, thus
0 1
m
V
0.5
−90 0
0 100 200 300 400 0 100 200 300 400
0 1
INa
0.5
h
−200
−400 0
0 100 200 300 400 0 100 200 300 400
0 1
Isi
−2 0.5
j
−4 0
0 100 200 300 400 0 100 200 300 400
1 1
IK
0 0.5
d
−1 0
0 100 200 300 400 0 100 200 300 400
4 1
2
IK1
0.5
f
0 0
0 100 200 300 400 0 100 200 300 400
2 1
IKp
0.5
0
−1 0
0 100 200 300 400 0 100 200 300 400
−3
x 10
5
4
Cai
Ib
2
0
0
0 100 200 300 400 0 100 200 300 400
time (msec) time (msec)
Fig. 2.3 LR1 model: Transmembrane potential v, ionic currents INa ; Isi ; IK ; IK1 ; IKp ; Ib , gating
variables m, h, j , d , f , X and intracellular calcium concentration ŒCa2C i as functions of time
Slow inward current Isi . The general form of this current is the same as the one
of the Is current in the Beeler-Reuter model
Isi D gN si df .v Esi /;
The dynamics of the gating variable d , f and the intracellular calcium concentration
ŒCa2C i is the same as in the Beeler-Reuter model.
Time-dependent potassium current IK . Unlike in the Beeler-Reuter model,
this current is controlled not only by an activation gate X , but also by an inactivation
gate Xi . The magnitude of the current is given by
IK D gN K XX i .v EK /;
48 2 Mathematical Models of Cellular Bioelectrical Activity
where
1:02
˛K1 D ;
1 C exp.0:2385.v EK1 59:215//
IKp D gN Kp Kp .v EKp /;
plays a role during the plateau phase of the action potential, restoring the cell to
its resting state. The maximal current conductance is gN Kp D 1:83 104 mS=mm2 ,
the reversal potential is the same of the time-independent potassium current, thus
EKp D EK1 and
1
Kp D :
1 C exp..7:488 v/=5:98/
Ib D gN b .v Eb /;
In 1994, Luo and Rudy developed a new mammalian ventricular membrane model
[309, 310], mainly based on data from the guinea pig. The new model includes a
detailed phenomenological description of calcium-induced calcium-release, intra-
cellular calcium cycling and calcium buffering. The ionic current Iion is given by
the sum of N D 11 currents
Iion D INa C ICa.L/ C IK C IK1 C IKp C INaCa C INaK C InsCa C IpCa C IbCa C IbNa :
Figure 2.4 reports the time evolution of the transmembrane potential v, the gating
variables m, h, j , d , f , Xs1 , Xs2 , Xr and the intracellular calcium concentration
ŒCa2C i .
Fast inward sodium current INa . The formulation of this current is the same as
the LR1 model, but the maximum conductance was decreased to gN Na D 16 mS=cm2.
L-type calcium current ICa.L/ . This current, equivalent to the slow inward
current Isi from the LR1 model, is the sum of the currents through the L-type
calcium channel that is permeable to Ca2C , NaC and K C . The name is due to its
activity, which is long lasting in comparison to other calcium currents, such as the
T-type transient calcium current. The magnitude of the current is given by
100 1.5
1
m
0 0.5
v
0
−100 −0.5
1.5 1.5
1 1
h
0.5 0.5
j
0 0
−0.5 −0.5
1.5 1.5
1 1
d
0.5 0.5
f
0 0
−0.5 −0.5
0.4 0.2
s1
Xs2 0.1
0.2
X
0
0
−0.1 −4
x 10
1.5 4
[Ca2+]i
1
r
0.5 2
X
0
−0.5 0
0 100 200 300 400 0 100 200 300 400
Fig. 2.4 LRd model: Transmembrane potential v, gating variables m, h, j , d , f , Xs1 , Xs2 , Xr and
the intracellular calcium concentration ŒCa2C i as functions of time
where
The INCa.L/Ca , INCa.L/Na and INCa.L/K terms represent the maximum currents through
the channel for each ion species. Because there is only one physical channel, each
of these equations is governed by the same gating variables. The activation gate d
and the inactivation gate f follow the Hodgkin-Huxley formalism (2.22), while the
further inactivation gate fCa is governed by the following equation
1
fCa D 2 ;
ŒCa2C i
1C KmCa
2.9 Cardiac Action Potential Models 51
where KmCa is the half activation concentration for Ca2C set to 0:6 103 mM. The
three fully activated current terms follow the short channel limit, thus
vF 2
Si ŒS i exp.zS vF=RT/ So ŒS o
INS D PS z2S ;
RT exp.zS vF=RT/ 1
IK D gN K X 2 Xi .v EK /;
where
1
Xi D :
1 C exp v56:26
32:1
Plateau potassium current IKp . The formulation of this current is the same as
the one in the LR1 model.
Sodium-calcium exchanger INaCa .
1 1 1
INaCa D kNaCa 3 C 3
Km;Na C ŒNa o Km;Ca C ŒCa o
2C 1 C k exp.. 1/vF=RT/
C 3
sat
C 3
exp. vF=RT/ŒNa i ŒCa o exp.. 1/vF=RT/ŒNa o ŒCa2C i ;
2C
where kNaCa D 2 000 A=F, Km;Na D 87:5 mM, Km;Ca D 1:38 mM, ksat D 0:1
and D 0:35.
Sodium-potassium pump INaK .
1 ŒK C o
INaK D INNaK fNaK C
C ;
1 C .Km;Nai =ŒNa i / 1:5 ŒK o C Km;Ko
52 2 Mathematical Models of Cellular Bioelectrical Activity
where KmnsCa D 1:2 M and INnsNa ; INnsK are computed according to the short
channel limit formula.
Sarcolemmal calcium pump IpCa .
ŒCa2C i
IpCa D INpCa ;
Km;p.Ca/ C ŒCa2C i
d ŒK C i Acap
D .IK;t 2INaK /; (2.25)
dt Vmyo F
where Acap is the capacitative area of the membrane, Vmyo is the volume of the
myoplasm, F is the Faraday constant and IK;t is the total K C current through all
ion channels, i.e.
d ŒNaC i Acap
D .INa;t C 3INaK C 3INaCa /;
dt Vmyo F
where INa;t is the total NaC current through all ion channels, i.e.
where ICa;t is the total Ca2C current through all ion channels, i.e.
The development of the LRd model continued after 1994 and the latest version is
described in the 2007 work [306] by Livshitz and Rudy. Among the major updates,
we recall those introduced in 1995 by [583], in particular the original formulation
of the T-type transient calcium current ICa.T / and the introduction of the rapid and
slow potassium currents IKr , IKs , replacing the obsolete time-dependent potassium
current IK .
T-type calcium current ICa.T / . The magnitude of this current is given by
with
p
gKr D 0:02614 ŒK C o =5:4; EKr D .RT=F / log.ŒK C o =ŒK C i /;
1
rD
1 C exp..v C 9/=22:4/
where
0:6
gKs D 0:433 1 C ;
1 C .0:000038=ŒCa2C i /1:4
C
ŒK o C PNa;K ŒNaC o
EKs D .RT=F / log :
ŒK C i C PNa;K ŒNaC i
and
respectively.
Membrane models have proven to be a useful and widely accepted tool for studying
the electrophysiology and contractility of excitable cells. The second-generation of
membrane ionic models incorporates the dynamics of the cytoplasmic concentra-
tions of NaC , K C and Ca2C and in the literature concerns regarding their behavior
in response to prolonged periods of rapid pacing have been reported, see e.g.
[163, 213, 244, 307, 417, 573]. In particular, drift of intracellular ion concentrations
and transmembrane potential v and the existence of an infinite number of steady
56 2 Mathematical Models of Cellular Bioelectrical Activity
states, are often cited as problems with such models, see [213, 573]. In order to
address these concerns, models based on the principle of charge conservation have
been formulated. In this formulation, the differential equation computing v from
the transmembrane current, i.e. the first equation in (2.24), is replaced with an
algebraic equation relating v to intracellular ion concentrations, see [163, 539]. This
algebraic approach has proven to be stable with respect to drift in the computed ion
concentrations and v. We present here the algebraic approach applied to the LRd
model, as done in [244].
Introducing the total intracellular calcium concentration ŒCa2C i;t , given by the
sum of troponin (trpn) and calmodulin (cmdn) bound and free Ca2C , i.e.
ŒtrpnŒCa2C i ŒcmdnŒCa2C i
ŒCa2C i;t D ŒCa2C i C C ;
ŒCa2C i C Km;trpn ŒCa2C i C Km;cmdn
and the total JSR calcium concentration ŒCa2C jsr;t , given by the sum of calse-
questrin (csqn) bound and free JSR Ca2C , i.e.
ŒcsqnŒCa2C jsr
ŒCa2C jsr;t D ŒCa2C jsr C ;
ŒCa2C jsr C Km;csqn
we have that
d ŒCa2C i;t 2C
1 d ŒCa i d ŒCa2C jsr;t 1 d ŒCa
2C
jsr
D ˇmyo ; D ˇjsr : (2.29)
dt dt dt dt
Substituting (2.29) into Eqs. (2.26) and (2.27), we obtain
d ŒCa2C jsr;t
D Itr Irel : (2.31)
dt
Combining (2.28), (2.30) and (2.31), we have
dv
Cm D Iion
dt
D .IK;t C INa;t C ICa;t C INaK C INaCa /
Vmyo F d ŒK C i d ŒNaC i d ŒCa2C i;t (2.34)
D C C2
Acap dt dt dt
Vnsr d ŒCa2C nsr Vjsr d ŒCa2C jsr;t
C2 C2 ;
Vmyo dt Vmyo dt
hence it holds
V F C C Vjsr
Cm v Amyo
cap
ŒK i CŒNa i C2ŒCa 2C
i;t C2 Vnsr
Vmyo
ŒCa 2C
nsr C2 Vmyo
ŒCa 2C
jsr;t DC0 ;
Vmyo F C
C0 D e D ŒK e C ŒNaC e C 2ŒCa2C e ;
Acap
As a consequence, it is clear from this conservation law that the initial values of v
and of the ion concentrations can not be selected arbitrarily.
Summarizing, we have two equivalent mathematical formulations:
• The common full-differential approach, composed of the differential system for
the state variables transmembrane potential, gating variables and concentrations
of ŒNaC i , ŒK C i , ŒCa2C i , ŒCa2C nsr , ŒCa2C jsr , where due to the latent conserva-
tion relationship, the state variables are not linearly independent, requiring that
the initial condition must satisfy the conservation law;
• The differential-algebraic approach, composed of the differential system for
the concentrations of ŒNaC i , ŒK C i , ŒCa2C i , ŒCa2C nsr , ŒCa2C jsr and gating
variables, that is coupled with the algebraic relationship for the transmembrane
potential.
We remark that in the former approach, as a consequence of the linear dependence of
the state variables, the resulting jacobian matrix is singular. Therefore, the study of
stability of equilibrium points and cycles as functions of relevant parameters should
be performed using the differential-algebraic formulation.
When an external current is applied, Eq. (2.33) becomes
dv
Cm C Iion D Iapp :
dt
To satisfy the algebraic conservation relation between the transmembrane poten-
tial and ionic concentrations, it is required that charges carried by a stimulation
current are accounted for a specific ionic species. Usually, in many simulation
studies of periodical pacing, K C is chosen as a charge ion for the stimulus current.
Therefore Eq. (2.25) is modified as follows:
d ŒK C i Acap
D .IK;t 2INaK Iapp /:
dt Vmyo F
would be added at the right hand side of Eq. (2.35), which might alterate the
membrane dynamics for long-term pacing, see [244].
2.9 Cardiac Action Potential Models 59
In order to investigate phenomena on larger spatial and temporal scales, and also for
theoretical purposes, several reduced membrane models have been developed that
do not seek to model sub-cellular processes, but only to provide an action potential
180
160
(ms)
140
90
APD
120
100
Fig. 2.5 Action potential duration restitution curve computed with the LRd model
60 2 Mathematical Models of Cellular Bioelectrical Activity
Here 0 < a < 1; k; ; ; in ; out ; open ; close ; 0 < vgate < 1 are given
constants;
• Morris-Lecar model [337]:
where
The parameters of the Morris-Lecar model are: Cm D 20 F=cm2, vCa D 120 mV,
vK D 84 mV, gK D 8 mS=cm2, vL D 60 mV, gL D 2 mS=cm2, v1 D 1:2 mV,
v2 D 18 mV. For the remaining parameters, we consider three different calibrations
denoted by ML1, ML2, ML3 in Table 2.4 and used later in the section on
bifurcation diagrams.
where H is the Heaviside function centered at zero and the ionic current
f .v; u; w/ D Ifi Iso Isi is the sum of the three currents
The values of the parameters in the previous expressions can be found in Table I
of [170].
62 2 Mathematical Models of Cellular Bioelectrical Activity
From ic D C du
dt , vR D Ri (Ohm’s law), vL D L dt (Faraday’s law), we have
di
8
ˆ du
<C D i F .u/ C Ia ;
dt
:̂ L di D u Ri v0 :
dt
In order to write this system in dimensionless form, we scale the time as D
Rm t=L, where Rm D 1=F 0 .0/ and we define the dimensionless variables v. / D
u.L =Rm /=vm ; w. / D Rm i.L =Rm/=vm , where vm is the maximal value of u.
Then
8
ˆ
ˆ
dv
D
L du
D
L L
.i F .u/ C Ia /D 2 .
Rm i Rm F .u/ Rm
C
< Ia /;
d vm Rm dt vm Rm C Rm C vm vm vm
ˆ dw Rm L di L u R Rm i v0
:̂ D D .v Ri v0 / D :
d vm Rm dt vm L vm Rm vm vm
0.5
Ω−, g<0
g
0.4
Ω+g, g>0
Ω−f , f<0
0.3 g(v,w)=0
Ω+, f>0
0.2 f
w
f(v,w)=0
0.1
wmax
h h h+
− 0
0
w
min
−0.1
−0.2
−0.4 −0.2 0 0.2 0.4 0.6 0.8 1 1.2
v
Fig. 2.6 v; w-phase plane, nullclines and notations for the FHN model
We assume a situation as the one in Fig. 2.6, where the nullcline f .v; w/ D 0
is a cubic-like function in v with 3 distinct zeros, local maximum value wmax , local
minimum value wmin and the nullcline g.v; w/ D 0 has only one intersection point
.v ; w / with f D 0 (for an example with more than one intersection, see Fig. 2.9).
More precisely, we assume that it exists an interval .wmin ; wmax / such that:
– For w > wmax 9Šv W f .v; w/ D 0, defining a function v D h .w/;
– For w < wmin 9Šv W f .v; w/ D 0, defining a function v D hC .w/;
– For wmin < w < wmax there are three real roots h .w/ < h0 .w/ < hC .w/ of
f .v; w/ D 0.
The cubic-like nullcline f .v; w/ D 0 partitions the v; w-plane into two regions,
˝f D f.v; w/ W w > w; where f .v; w/ D 0g and ˝fC D f.v; w/ W
w < w; where f .v; w/ D 0g. We assume that f < 0 in ˝f and f > 0 in
˝fC . Analogously, the other nullcline g.v; w/ D 0 partitions the plane into two
regions ˝g D f.v; w/ W w > w; where g.v; w/ D 0g and ˝gC D f.v; w/ W
w < w; where g.v; w/ D 0g, and we assume that g < 0 in ˝g and g > 0
in ˝gC . These assumptions guarantee that the partial derivatives of f and g are
nonpositive or nonnegative but we will make the stronger assumptions @w f .v; w/ <
0; @v f .v; w/ > 0; and @v g.v; w/ > 0; @w g.v; w/ < 0.
64 2 Mathematical Models of Cellular Bioelectrical Activity
According to the chosen initial conditions .v0 ; w0 / and the applied current IOa , the
generalized FitzHugh-Nagumo system (2.37) admits solutions with quite different
dynamics. The intersection point .v ; w / between the nullclines f D 0, g D 0 is
an equilibrium point and its stability is determined by the eigenvalues 1 ; 2 of the
Jacobian matrix evaluated at .v ; w /
@v f @w f
J D W
@v g @w g
– We have stability if 1 ; 2 are both real and negative (stable node or sink) or if
they are a complex conjugate pair with negative real part (stable spiral);
– We have instability if 1 ; 2 are both real positive (unstable node or source)
or they have opposite signs (saddle) or they are a complex conjugate pair with
positive real part (unstable spiral).
Equivalently, the stability of a planar dynamical system is often described in terms
of the trace tr.J / D 1 C 2 and determinant det.J / D 1 2 of the Jacobian matrix.
For our generalized FitzHugh-Nagumo system (2.37), we have stability if
.v ; w / lies on the lower branch h .w/ or upper branch hC .w/ of the nullcline
f D 0, while we have instability if .v ; w / lies on its middle branch h0 .w/. In
Fig. 2.7, we consider first some examples without applied current, IOa D 0. We will
then vary IOa in Figs. 2.8 and 2.9 we will consider a case with three equilibrium
points. In each figure, the left panel shows the nullclines (dashed lines) and the
trajectory (continuous line) in the phase plane v; w starting from the initial point
.v0 ; w0 / (marked with ˘); the equilibrium point .v ; w / is marked with ?.
(a) FHN dynamics with zero applied current IOa (Fig. 2.7).
(a1) Stable equilibrium: subthreshold response. Consider first w0 2 .wmin ; wmax /
and v0 < h0 .w0 /. Then v returns to the equilibrium point v (Fig. 2.7 top,
stable spiral), because in the region where f < 0 if the initial point .v0 ; w0 /
is above the nullcline g D 0, then here g < 0 and .v; w/ returns to the
equilibrium point .v ; w /, and if .v0 ; w0 / is below the nullcline g D 0,
then here g > 0 so w initially increases until the trajectory crosses g D 0
into the region where g < 0 and .v; w/ again returns to .v ; w /. Analogous
considerations apply when w0 2 .wmin ; wmax / and v0 < h0 .w0 / in the region
where f > 0.
(a2) Stable equilibrium: action potential (Fig. 2.7 middle, stable spiral). If instead
v0 > h0 .w0 / (for simplicity we still assume w0 2 .wmin ; wmax /), then we
have an action potential. First, we have an excitation phase during a time
D O./, where v tends to the upper branch hC .w/ of the nullcline f D 0,
2.9 Cardiac Action Potential Models 65
V
0
0.2
−1
0 1 2 3
t
W
0.1
0.2
W
0
V =0 eig 1 =−5.25+8.79986i
W =0 eig 2 =−5.25−8.79986i −0.2
−0.1 0 1 2 3
−0.5 0 0.5 1 t
V
Phase Plane and Nullclines
0.3 1
V
0
0.2
−1
0 1 2 3
t
W
0.1
0.2
0
W
0
V =0 eig 1 =−5.25+8.79986i
W =0 eig 2 =−5.25−8.79986i −0.2
−0.1 0 1 2 3
−0.5 0 0.5 1 t
V
Phase Plane and Nullclines
0.3 1
V
0
0.2
−1
0 1 2 3 4 5
t
W
0.1
0.5
0
W
0
V =0 eig 1 =4.75+8.511i
W =0 eig 2 =4.75−8.511i −0.5
−0.1 0 1 2 3 4 5
−0.5 0 0.5 1 t
V
Fig. 2.7 FHN model (2.37) with IOa D 0; D 0:01. In top and middle panels f .v; w/ D v.v
0:1/.1 v/ w; g.v; w/ D v w=2, while in bottom panels f .v; w/ D v.v C 0:1/.1 v/
w; g.v; w/ D v w=2. Phase plane trajectories, nullclines, coordinates of the marked equilibrium
point ? and associated eigenvalues (left), solutions v; w as functions of time (right)
66 2 Mathematical Models of Cellular Bioelectrical Activity
0.4
V
0
0.3 −1
0 1 2 3 4 5
t
W
0.2
0.2
0.1
W
0
0
V =0 eig 1 =−6+9.16515i
W =0 eig 2 =−6−9.16515i −0.2
−0.1 0 1 2 3 4 5
−0.5 0 0.5 1 t
V
Phase Plane and Nullclines
0.5 1
0.4
V
0
0.3 −1
0 1 2 3 4 5
t
W
0.2
0.4
0.1
W
0.2
0 V = 0.5 eig 1 =20.5692
W =0.25 eig 2 =2.43082
0
−0.1 0 1 2 3 4 5
−0.5 0 0.5 1 t
V
Phase Plane and Nullclines
0.5 1
V =0.81538 eig 1 =−25.8826
W =0.40769 eig 2 =−6.18714
0.4
V
0.3 −1
0 1 2 3 4 5
t
W
0.2
0.5
0.1
W
0
0
−0.1 0 1 2 3 4 5
−0.5 0 0.5 1 t
V
Fig. 2.8 FHN model (2.37) with varying IOa , with f .v; w/ D v.v 0:1/.1 v/ w; g.v; w/ D
v 2w; D 0:01. IOa D 0 (top), 0:15 (middle), 0:3 (bottom). Phase plane trajectories, nullclines,
coordinates of the marked equilibrium point ? and associated eigenvalues (left), solutions v; w as
functions of time (right)
2.9 Cardiac Action Potential Models 67
V
W =0.15403 eig 2 =−10.2438 0
0.2
−1
0 1 2 3 4 5
t
W
W
0
V =0 eig 1 =2.5+6.6144i
W =0 eig 2 =2.5−6.6144i −0.2
−0.1 0 1 2 3 4 5
−0.5 0 0.5 1
t
V
Fig. 2.9 FHN model (2.37) with nullclines intersecting in three points: f .v; w/ D v.v C 0:1/.1
v/ w; g.v; w/ D v 5w; D 0:01; IOa D 0. Phase plane trajectories, nullclines, coordinates
of the marked equilibrium points ? and associated eigenvalues (left), solutions v; w as functions of
time (right)
0.4 1
v
0.3 0
−1
0.2 0 1 2 3
t
w
0.1 0.2
0.1
w
0 0
−0.1
−0.1 0 1 2 3
−1 −0.5 0 0.5 1
t
v
Fig. 2.10 Cathode make due to current pulse Ia D 0:2 lasting 0.02 ms for FHN model (2.37) with
f .v; w/ D v.v 0:1/.1 v/ w; g.v; w/ D v w=2; D 0:01. Phase plane trajectories and
nullclines (left), solutions v; w as functions of time (right)
2.9 Cardiac Action Potential Models 69
1
0.2
v
0
0.1
−1
0 1 2 3
0 t
w
0.2
−0.1 0.1
w
0
−0.2 −0.1
0 1 2 3
−1 −0.5 0 0.5 1 t
v
1
0.2
v
0
0.1
−1
0 1 2 3
0 t
w
0.2
−0.1 0.1
w
0
−0.2 −0.1
0 1 2 3
−1 −0.5 0 0.5 1 t
v
Fig. 2.11 Anode break due to current pulse Ia D 0:2 lasting 0.1 ms (top) or 0.6 ms (bottom) for
FHN model (2.37) with f .v; w/ D v.v 0:1/.1 v/ w; g.v; w/ D v w=2; D 0:01. Phase
plane trajectories and nullclines (left), solutions v; w as functions of time (right)
and the variables v; w start a trajectory toward the closest nullcline branch.
When the pulse is turned off at time Tstim , the f nullcline returns to the original
curve f D 0. By this time, v and w have reached a point in phase-space that, for
proper choices of pulse amplitude and interval duration, belongs to the excitable
region, so that an action potential is generated. Figure 2.10 shows such an
action potential generated by a pulse with Ia D 0:2, Tstim D 0:02, known
as cathode make mechanism. Instead Fig. 2.11 shows two cases with action
potentials generated by pulses with Ia D 0:2, Tstim D 0:1 (top) or Ia D 0:2,
Tstim D 0:6 (bottom), known as anode break mechanism. Additional activation
mechanisms (known as cathode break and anode make) are present in space-
time cardiac domains, in particular in three-dimensional domains, where more
complex virtual electrode polarization (VEP) activation patterns are present, see
Chap. 9.1.
70 2 Mathematical Models of Cellular Bioelectrical Activity
There are several mechanisms from which limit cycles arise in models of excitable
cells. A periodic behavior of the transmembrane potential can be thought of as a
transition back and forth between hyperpolarized and depolarized states. We review
here four main mechanisms that govern the onset of oscillations in some of the
previous ionic models. For more details we refer to e.g. [288, 419].
A possible motivation for studying the dynamical response of an ionic model
to a sustained injection of a constant current, i.e. a bias current, is for example
the fact that during ischemia there is an accumulation of extracellular potassium in
the ischemic zone, leading to an increase of the membrane potential. This creates
a current flow between the injured and healthy zone. We study here this effect
by modeling the injury current as a time- and voltage-independent bias current,
and considering simple membrane models. We illustrate the first two mechanisms
considering the simplest FHN model.
The dynamics of the FHN system as a function of an applied bias current IOa ,
i.e. a continuous constant current, can be summarized by a bifurcation diagram with
respect to IOa . We present now some bifurcation and frequency diagrams computed
by the XPPAUT software [165].
(a) Choosing D 0:1, i.e. modifying the time scale of the v variable, we
find in Fig. 2.12 that by increasing IOa , a periodic solution appears through
a supercritical Hopf-Bifurcation (HB) point near IOa D 0:075, leading to a
branch of small-amplitude periodic orbits of nonzero frequencies. By increasing
IOa further, we find that the periodic orbits have decreasing amplitudes and
then disappear at another HB point near IOa D 0:17. This type of dynamical
phenomenon, with oscillations arising with arbitrary small amplitude via a
supercritical HB point and with nonzero oscillation frequency, is also called
soft excitation by some authors.
V Frequency
1
0.0395
0.8
0.039
0.6
0.0385
0.4 0.038
0.0375
0.2
0.037
0
0.0365
-0.2 0.036
0.0355
-0.4
0.035
0 0.05 0.1 0.15 0.2 0.25 0 0.05 0.1 0.15 0.2 0.25
Ia Ia
Fig. 2.12 Left: bifurcation diagram for FHN model with f .v; w/ D v.v 0:1/.1 v/ w C
IOa ; g.v; w/ D v 2w; D 0:1 with respect to IOa . Right: associated frequency diagram
2.9 Cardiac Action Potential Models 71
(b) Choosing D 0:01, i.e. shortening the time constant of v, we find in Fig. 2.13
that for all values of IOa 0 there is a unique equilibrium point (EP), which
loses its asymptotical stability for values of IOa between two subcritical Hopf
bifurcation points, HB1 and HB2 and we see also the presence of a stable
limit cycle for values of IOa ranging approximately between 0.03 and 0.22.
Near the subcritical HB1 point there is a large-amplitude stable limit cycles and
bistability between stable oscillations and a stable equilibrium point, separated
by an unstable small-amplitude periodic orbit. The unstable limit cycle turns
back and coalesces at a turning bifurcation point with the stable limit cycle at a
value of IOa that is smaller than the one associated to the subcritical HB point. A
zoom of the left part of this interval is shown in Fig. 2.13, right panel, revealing a
subcritical HB1 near IOa D 0:0319. The branch of unstable periodic orbits comes
out near HB1 and lies on the side of HB1 where there is a stable branch of EP.
Decreasing IOa from the HB1 value, a saddle-node periodic bifurcation (SNPB)
point appears, connecting a stable and an unstable branch of limit cycles. Above
it two stable structures appear, a stable EP and a stable limit cycle, as well as an
unstable limit cycle.
The frequency of the limit cycle as a function of IOa , also computed with
XPPAUT, is shown in Fig. 2.14, left panel, together with a zoom near the Hopf
bifurcation point shown in the right panel.
This type of onset of oscillations arising abruptly with stable large-amplitude
oscillations and with relatively high frequencies it is also called hard excitation
mechanism or also saddle-node bifurcation of limit cycles (SNBLC), since for
increasing IOa there is the creation of a pair of limit cycles, one stable and the
other unstable with non zero frequencies.
An analogous reflected situation happens at the right part of the bifurcation
interval (zoom not shown), where unstable small-amplitude periodic orbits now
appear on the stable equilibrium branch on the right side of the right HB point.
V V
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.0315 0.0316 0.0317 0.0318 0.0319
Ia Ia
Fig. 2.13 Left: bifurcation diagram for FHN model with f .v; w/ D v.v 0:1/.1 v/ w C
IOa ; g.v; w/ D v 2w; D 0:01 with respect to IOa . Right: zoom of the left bifurcation point
72 2 Mathematical Models of Cellular Bioelectrical Activity
Frequency Frequency
0.02 0.02
0.015 0.015
0.01 0.01
0.005 0.005
0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.0315 0.0316 0.0317 0.0318 0.0319
Ia Ia
Fig. 2.14 Left: frequency diagram for FHN model with f .v; w/ D v.v0:1/.1v/w; g.v; w/ D
v 2w; D 0:01. Right: zoom of the left interval
V Frequency
40 0.025
20 0.02
0 0.015
-20 0.01
-40 0.005
-60 0
60 80 100 120 140 160 180 200 220 240 60 80 100 120 140 160 180 200 220 240
Ia Ia
Fig. 2.15 Left: bifurcation diagram for the Morris-Lecar model (2.36) with respect to IOa with the
parameter calibration ML1 in Table 2.4. Right: frequency diagram
The previous mechanisms (a) and (b) of onset of oscillations are associated
with local bifurcations that are local phenomena where limit cycles originate
with small amplitude when a single equilibrium state changes from stable to
unstable or viceversa. A bifurcation diagram of the same local type (b) is also
obtained for the Morris-Lecar model using the parameter calibration (ML1)
reported in Table 2.4; see Fig. 2.15.
Moreover, the bifurcation diagram of the Hodgkin-Huxley model, displayed
in Fig. 2.16, shows the same bistable behavior between a SNPB and a HB1.
Differently from the other models, a second supercritical Hopf bifurcation point
appears, where stable periodic orbits of small amplitude come out lying near the
unstable branch. See Fig. 2.17 for the associated frequency diagram.
2.9 Cardiac Action Potential Models 73
V V
40 40
20 20
0 0
-20 -20
-40 -40
-60 -60
-80 -80
0 20 40 60 80 100 120 140 160 6 8 10 12 14
Ia Ia
Fig. 2.16 Left: bifurcation diagram for HH model with respect to IOa . Right: zoom of the left
bifurcation point
Frequency
0.2
0.15
0.1
0.05
0
0 20 40 60 80 100 120 140 160
Ia
Fig. 2.17 Left: frequency diagram for HH model with respect to IOa
(c) We now look at two other mechanisms of onset of oscillations that are very
different from the previous mechanisms (a) and (b) since they are associated
with global bifurcations and show that it is possible to create or destroy spiking
with arbitrary low frequencies when the amplitude of the sustained injected
current is increased. A type of global bifurcation is related to the appearance
of homoclinic bifurcation. We recall that an homoclinic orbit is a closed curve
that has a single equilibrium point located along the curve and the period of the
orbit is infinite. Figure 2.18 displays the bifurcation diagram with respect to the
injected current IOa for the Morris-Lecar model using the calibration parameter
74 2 Mathematical Models of Cellular Bioelectrical Activity
V Frequency
0.05
40 0.045
0.04
20
0.035
0.03
0
0.025
0.02
-20
0.015
-40 0.01
0.005
-60 0
-20 0 20 40 60 80 100 120 140 20 40 60 80 100 120
Ia Ia
Fig. 2.18 Transmembrane potential (left) and frequency (right) bifurcation diagrams with respect
to IOa for Morris-Lecar model (2.36) with the parameter calibration ML2 in Table 2.4
(ML2) reported in Table 2.4, still computed with XPPAUT. The diagram starts
on the left with a stable node branch but by increasing IOa two unstable branches
of equilibrium points appear, the upper one associated with unstable spirals,
and the intermediate with saddle points. By increasing IOa further, the unstable
saddle and stable nodes coalesce at a HB bifurcation point and subsequently
disappear. A limit cycle then emerges as a homoclinic orbit that has a finite
(large) amplitude with zero frequency as shown in Fig. 2.18, right panel. This
type of dynamical behavior associated with saddle-node on a limit cycle is
sometimes called a saddle-node homoclinic orbit. By increasing IOa further, the
stable limit cycle disappears through a saddle-node bifurcation of limit cycles.
(d) We now increase the constant from 0.066 to 0.23, yielding a decrease of the
time constant of the potassium gating variable w, thus increasing its changing
rate. Figure 2.19 displays the bifurcation diagram for the Morris-Lecar model
using the parameter calibration (ML3) reported in Table 2.4. After an initial
branch of stable node, we have again two saddle node bifurcation points
connecting two unstable branches and a subcritical HB appears on the upper
branch, but now with a value between the two saddle node bifurcation points.
As in the previous case (c), the stable branch of limit cycles terminates again
with an homoclinic orbit which is now on the middle branch before the right
saddle node bifurcation point. For IOa ranging between the subcritical HB values
and the right saddle node bifurcation there are two stable equilibrium points and
a stable limit cycle, yielding a twistable dynamics.
This type of global bifurcation is called saddle-homoclinic bifurcation. The
frequency of the stable limit cycles approaches the zero frequency of the homoclinic
orbit more rapidly than in the previous SNBLC diagram.
2.9 Cardiac Action Potential Models 75
V Frequency
20 0.07
10
0.06
0
0.05
-10
0.04
-20
-30 0.03
-40 0.02
-50
0.01
-60
0
-20 -10 0 10 20 30 40 50 60 30 32 34 36 38 40 42 44
Ia Ia
Fig. 2.19 Transmembrane potential (left) and frequency (right) bifurcation diagrams with respect
to IOa for Morris-Lecar model (2.36) with the parameter calibration ML3 in Table 2.4
V V
20 20
0 0
-20 -20
-40 -40
-60 -60
-80 -80
-100 -100
0 0.5 1 1.5 2 2.5 3 0 1 2 3 4 5
iapp
iapp
Fig. 2.20 Bifurcation diagram for LR1 model with the original calibration parameters (left) and
the with the IK current scaled by a factor 3 (right)
We have also computed the bifurcation diagrams for the LR1 model with both
the original calibration parameters and with the IK current scaled by a factor
3. The results obtained with the XPPAUT software display bifurcation diagrams
of Fig. 2.20, left and right panels, similar to the diagram of Figs. 2.19 and 2.18,
respectively, but with different stability properties.
For bifurcation diagrams related to second generation models, we refer e.g.
to [57].
Chapter 3
Mathematical Models of Cardiac Cells
Arrangements: The Bidomain Model
We model an excitable cell as a long cylindrical cable and assume the so-called core
conduction assumption, stating that the potential along the cable depends only on
the length variable and not on radial or angular variables, so that the cable model is
one-dimensional. Then we assume that the cable is composed of a number of short
sections of length dx with isopotential membrane, and that in each section we have
only the transmembrane and axial currents, as illustrated in the following diagram
Je(x) ue(x) ue(x+Δ x)
i
reΔ x m
r Δx
i
The relationship between the intra- and extracellular current fluxes Ji;e (A cm2 )
and the intra- and extracellular potentials ui;e (V ) is given by Ohm’s law
1
Ji D @x ui D i @x ui ;
ri
1
Je D @x ue D e @x ue ;
re
where i;e are the intra- and extracellular conductivities (1 cm1 ) and ri;e are the
intra- and extracellular resistances per unit space ( cm).
We denote by im the current across the membrane; it is an outward current, i.e. it
is considered positive when directed toward the extracellular space. By Kirchhoff’s
law of conservation of electric charges, we have
Z xCx
Ji .x/ D Ji .x C x/ C im .s/ds;
x
Z xCx
Je .x C x/ D Je .x/ C im .s/ds;
x
Je(x) Je(x+Δ x)
im Δ x
Ji(x) Ji(x+Δ x)
Δx
@x Ji D im ;
@x Je D im ;
which imply
We consider a fiber of length L, cut and healed at the two ends, leading to insulating
boundary conditions for the intracellular medium
Ji D 0 at x D 0; x D L:
Je D Iapp .t/ at x D 0; x D L;
3.1 Models of Cardiac Fibers 79
Ji .0/ C Je .0/ D Ji .L/ C Je .L/ D Iapp .t/; i.e. Ji .x; t/ C Je .x; t/ D Iapp .t/:
The membrane current per unit area of the membrane surface is given by
Im D Cm @t v C Iion :
This can be expressed as a current im per unit volume by multiplying by the ratio
between the membrane surface area for a given length of fiber and the volume
enclosed by this surface, i.e.
im D Im D .Cm @t v C Iion /:
Collecting together this last equation with Kirchhoff’s and Ohm’s laws, we have the
system
8
ˆ
ˆ v D ui ue
<
i @x ui e @x ue D Iapp .t/
ˆ @x .i @x ui / D im
:̂
im D .Cm @t v C Iion /:
1
i @x ui e @x ui C e @x v D Iapp .t/ ) @x ui D .e @x v Iapp .t//;
i C e
and from the third and fourth equation, we finally obtain the cable equation
i
@x .@x v/ @x Iapp .t/ D im D .Cm @t v C Iion /; (3.1)
i C e
where
i e
D
i C e
3.1.2 Homogenization
dui d
Ji .x/ D i ; Ji .x/ D im .x/; (3.2)
dx dx
where im is the membrane current. Since the extracellular medium is assumed
equipotential, the extracellular flux vanishes and, from v D ui ue , we have
@x D @x . Setting D i and combining the two equations in (3.2), we get
@v @ui
d dv
D im : (3.3)
dx dx
The junctions are more resistive than the cytoplasmic fluid, therefore we assume
that varies periodically with period equal to the distance between two subsequent
junctions. Hence, D . x / is periodic with period , or, defining D x , D ./
is periodic with period 1.
In general, if the cells present some kinds of heterogeneities, we suppose that
then it holds
Z 1 Z 1
v .x/ D V .x; / d D V0 .x; / d :
0 0
d d x @ x @ x d
v .x/ D V x; D V x; C V x;
dx dx @x @ dx
@ x @ x 1
D V x; C V x; ;
@x @
thus
d @ 1 @
D C
dx @x @
and
d dv d @V 1 @V
D C
dx dx dx @x @
@ @V 1 @V 1 @ @V 1 @V
D C C C
@x @x
@ @
@x @
@ @V 1 @ @V 1 @ @V 1 @ @V
D C C C 2 :
@x @x @x @ @ @x @ @
1 1
If we equal the terms of order , ,
2
1, we get the following differential problems
8
< @ @
.x; / V0 .x; / D 0 for 2 .0; 1/
.P 0/ @ @ (3.6)
:
V0 .x; 0/ D V0 .x; 1/;
8
ˆ
ˆ
@ @
D
@ @
C
ˆ
ˆ .x; / V 1 .x; / .x; / V 0 .x; /
ˆ
ˆ @ @ @x @
ˆ
ˆ @ @
< .x; / V0 .x; / for 2 .0; 1/
.P1/ @ @x
ˆ
ˆ
ˆ
ˆ V .x; 0/ D V1 .x; 1/
ˆ
ˆ Z1
ˆ 1
:̂ V1 .x; / d D 0;
0
8 (3.7)
ˆ
ˆ
@ @
.x; / V2 .x; / D im .x; /
@ @
.x; / V0 .x; / C
ˆ
ˆ
ˆ
ˆ @ @ @x @x
ˆ
ˆ
ˆ
ˆ
@ @
.x; / V1 .x; / C
ˆ
ˆ
< @x @
.P 2/ @ @
ˆ
ˆ .x; / V 1 .x; / for 2 .0; 1/
ˆ
ˆ @ @x
ˆ
ˆ
ˆ
ˆ V2 .x; 0/ D V2 .x; 1/
ˆZ 1
ˆ
ˆ
:̂ V2 .x; / d D 0:
0
If v is a solution of (3.8), then integrating both the sides of the equation we get
Z 1 Z 1
@ @ @ @
v d D .x; 1/ v.x; 1/ .x; 0/ v.x; 0/ D S.x; / d :
0 @ @ @ @ 0
@ @
.x; 1/ v.x; 1/ .x; 0/ v.x; 0/ D 0;
@ @
It can be proved that (3.9) is also a sufficient condition for the solvability of
(3.8),Ri.e.:
1
if 0 S.x; / d D 0, then 9Š solution v.x; / periodic in with period 1 of (3.8).
v is unique up to a constant, that can be fixed by imposing zero mean value for
2 .0; 1/,
Z 1
v.x; / d D 0:
0
Problem (P0). If we multiply both the sides of (3.6) times V0 and integrate on
.0; 1/, we get
Z Z 1
1
@ @V0 @V0 2
0D V0 d D d
0 @ @ 0 @
and
Z 1 2 Z 1 2
@V0 @V0
0D d 0 d ;
0 @ 0 @
hence
@V0
D 0:
@
@ @V1 @ @V0 @ @V1 @V0
D ) C D 0;
@ @ @ @x @ @ @x
implying that
@V1 @V0
C D .x/
@ @x
@ 1
V0 .x/ D .x/ ;
@x N
.x/
84 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
1
N
.x/ D R1 :
1
0 .x;/ d
Setting
@
V1 .x; / D w.x; / V0 .x/;
@x
it follows
@V1 @w @V0 N @V0
D D 1 ;
@ @ @x @x
and then
@w N
D 1:
@
In order to have
Z 1
w.x; / d D 0;
0
3.1 Models of Cardiac Fibers 85
Therefore
@
V1 .x; / D w.x; / V0 .x/
@x
with
Z Z 1 Z
1 1 1
w.x; / D .x/
N ds N .x/ ds d C :
0 .x; s/ 0 0 .x; s/ 2
Since
@ N .x/
w.x; / D 1
@ .x; /
We recall that
Z 1 Z 1 Z 1
V .x; / d D V0 .x; / d D V0 .x/ d D V0 .x/;
0 0 0
The cable equation (3.1) has been extensively studied in the special version known
as the bistable equation
where
• f .v/ has three zeros at v D 0; ˛; 1, with 0 < ˛ < 1;
• f 0 .0/ < 0 and f 0 .1/ < 0.
The values v D 0 and v D 1 are stable steady solutions of the ODE @t v D f .v/. For
existence, uniqueness and stability of traveling wavefront solutions and long time
behavior of the solution of (3.11), see [173–175].
An example of f often used in this context is the cubic polynomial
for a suitable value of c. When written as a function of the new traveling wave
variable , the wave appears stationary. Because we use D x C ct, a solution with
a positive c corresponds to a wave moving from right to left, while a solution with
a negative c moves from left to right.
By imposing that v.x; t/ given in (3.12) satisfies the bistable equation (3.11), one
obtains that any traveling wave solution solves the following second order ODE
V cV C f .V / D 0: (3.13)
From (3.13), it follows that a traveling wave solution of the bistable equation is a
solution of the first order ODE system
V D W
W D cW f .V /;
connecting the steady states .V; W / D .0; 0/ and .V; W / D .1; 0/ in the .V; W /
phase plane. Such a trajectory, connecting two different steady states, is called a
heteroclinic trajectory and it approaches .0; 0/ as ! 1 and .1; 0/ as ! C1.
Beyond the steady states .0; 0/ and .1; 0/, which are saddle points, the other steady
state is .˛; 0/, which is a node, if c is positive, or a spiral point, if c is negative.
Because .0; 0/ and .1; 0/ are saddle points, in order to find a traveling wave solution,
we determine whether a parameter c exists such that the trajectory leaving .0; 0/ at
D 1 reaches .1; 0/ at D C1.
We first determine the sign of the velocity c. Supposing a monotone increasing
(V > 0) connecting trajectory exists, multiplying (3.13) by V and integrating from
D 1 to D C1, one obtains the following relationship
Z C1 Z 1
c W ./2 d D f .u/du: (3.14)
1 0
R1
Theorem 3.1. Suppose 0 f .u/du > 0. Then there exists a unique velocity c > 0
with a corresponding traveling wave solution of the bistable equation.
Proof. To verify uniqueness, from
dW f .V /
Dc ;
dV W
we observe that the slope dW dV of trajectories in the .V; W / phase plane is a monotone
increasing function of the parameter c. Let c D c0 be a velocity value for which
a connecting trajectory exists. Suppose now that for a value c1 > c0 , another
connecting trajectory exists. The trajectory leaving the saddle point .0; 0/ for c1
must lie above the connecting trajectory for c0 . For the same reason, the trajectory
approaching the saddle point .1; 0/ for c1 must lie below the connecting trajectory
for c0 . A single trajectory cannot lie above and below another one, thus there cannot
be a connecting trajectory for c > c0 . Analogously, there cannot be a connecting
trajectory for c < c0 . Hence the connecting trajectory is unique.
To verify existence, we first suppose c large. Let K be the smallest positive
number for which f .u/ u K; 8u W 0 < u 1 and let be any fixed positive
number. On the line W D V the slope of trajectories satisfies
dW f .V / f .V / K
Dc Dc c :
dV W V
R1
which can occur only if 0 f .u/du 0. Since this is in contradiction with the
R1
hypothesis that 0 f .u/du > 0, it follows that the trajectory cannot reach V D 1 and
neither it can remain in the first quadrant, because W > 0 implies that V increases.
3.1 Models of Cardiac Fibers 89
W D BV.1 V /: (3.15)
.2B 2 C 1/V C .B 2 cB ˛/ D 0;
1 1
BDp c D p .1 2˛/;
2 2
A traveling pulse is a traveling wave solution that starts and ends at the same
steady state of the governing equations. In the previous section, we have seen that
a traveling front solution of the bistable equation corresponds to a heteroclinic
trajectory in the .V; W / phase plane, i.e. a trajectory that connects two different
steady states of the system. Analogously, a traveling pulse corresponds to a
trajectory that starts and ends at the same steady state in the .V; W / phase plane.
Such trajectories are called homoclinic orbits.
90 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
where is a small positive number and, without any loss of generality, the space
variable x has been scaled so that the diffusion coefficient is 2 , see e.g. [173–
175, 532]. This type of reaction-diffusion singular perturbation problem develops in
a short time a sharp propagating interface, i.e. a steep wavefront. In fact, in [85, 86],
it was shown that, after a time period of order O. log /, a sharp transition layer
of thickness of order O./ appears. Let us recall that the variable v represents the
transmembrane potential, while w represents a slow gating variable.
We now proceed further in understanding the structure of traveling wave
solutions of (3.16), by performing a singular perturbation analysis, see e.g. [79].
Starting from smooth initial data .v0 .x/; w0 .x//, the diffusion term @xx v and the
variation of w from its initial data can be neglected for a short time. In fact, setting
D t , vO .x; / D v.x; t/ and w.x;
O / D w.x; t/, then
@t vO D 2 @xx vO C f .Ov; w/
O
(3.17)
@t wO D g.Ov; w/: O
@t vO D f .Ov; w/
O
@t wO D 0:
O
Thus w.x/ D w0 .x/ and, using the notations introduced in Sect. 2.9.8 (see also
Fig. 2.6), in a short time (t D O./) vO ! h˙ .w0 /, if v0 .x/ ? h0 .w0 /. The region of
space where v D hC .w/ is called excited region, while the region where v D h .w/
is called recovered region. In these two regions, it holds
O D0
f .Ov; w/
(3.18)
@t wO D g.Ov; w/;
O
thus
vO D h˙ .w/
O
O D G˙ .w/
@t w O D g.h˙ .w/;
O w/:
O
However, there are regions of space, called interfaces, where diffusion is large and
(3.18) cannot be applied. To understand what happens when diffusion is large, we
rescale space and time by setting D t and D xy.t
/
, where y.t/ denotes the
position of the wavefront, i.e. the points
vO C y 0 . /Ov C f .Ov; w/
O D @ vO
y 0 . /wO D .g.Ov; w/
O @ w/;
O
vO C y 0 . /Ov C f .Ov; w/
O D0
0 (3.19)
y . /wO D 0:
lim f .Ov.; /; w. // D 0;
!˙1
i.e.
vO .˙1; / D h .w.
O //:
vO C c.w/O
O v C f .Ov; w/
O D0 (3.20)
has a heteroclinic orbit connecting two stable roots of f .Ov; w/ O D 0. This means
O is the unique value for which (3.20) has a solution with vO ! hC .w/
that c.w/ O as
! 1 and vO ! h .w/ O as ! C1.
Summarizing, in most of space, (3.18) holds. At any transition between the two
types of (3.18), a sharp transition in v occurs, with a wavefront traveling at the
speed y 0 .t/ D c.w/ if v D hC .w/ on the left and v D h .w/ on the right, or
y 0 .t/ D c.w/ if v D h .w/ on the left and v D hC .w/ on the right. We now
suppose that far to the right the medium is at rest, i.e. 9 wC W G .wC / D 0, and that
a wavefront of excitation moves from left to right with velocity y 0 .t/ D c.wC / > 0.
Following the same procedure used to derive (3.14), one obtains
R hC .w/
h .w/ f .v; w/dv
c.w/ D R C1 2 ;
1 v d
92 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
Z hC .w/
f .v; w/dv > 0: (3.21)
h .w/
If (3.21) fails to hold, then the medium is not sufficiently excitable to sustain
a traveling pulse.
0.07
0.065
CV (cm/ms)
0.06
0.055
0.05
20 40 60 80 100 120 140 160 180
DI (ms)
Fig. 3.1 Conduction velocity (CV) restitution curve computed with the cable equation coupled
with the LRd model
3.2 Models of Cardiac Tissue 93
reports the CV restitution curve calculated using the cable equation coupled with
the LRd model, i.e. the following system of PDEs
8
ˆ @v @2 v
ˆ
ˆ cm 2 C iion .v; w; c/ D iapp ;
ˆ
ˆ
< @t @x
dw
ˆ R.v; w/ D 0;
ˆ
ˆ dt
ˆ dc
:̂ S.v; w; c/ D 0;
dt
with cm D Cm , iion D Iion , D 1:2 103 1 cm1 and the functions Iion , S , R
given by the LRd model.
On the other hand, since the only active source elements lie on the membrane
m , each flux equals the membrane current per unit area Im , which consists of a
capacitive and an ionic term (see [250, 317]):
@v
nTi ˙i rui D nTe ˙e rue D Im D Cm C Iion : (3.22)
@t
Here Cm is the surface capacitance of the membrane per unit area and v WD ui j
m
ue j is the transmembrane potential evidencing that m is a discontinuity surface
m
for the potential (in the following, we will simply write v D ui ue ).
Denoting by Iis ; Ies the stimulation currents applied to the intra and extracellular
space, we have
Finally, the system (3.22), (3.23) must be supplemented by the initial conditions
v.; 0/ D v0 ; w.; 0/ D w0 ; on m :
For the electric potentials ui ; ue we can consider two characteristic length scales:
a micro scale related to typical cell dimensions fdc ; lc g and a macro scale
determined by a suitable length constant of the tissue. At the latter scale length,
i.e. at a macroscopic level, in spite of the discrete cellular structure, the cardiac
tissue can be represented by a continuous model. To identify this macroscopic
scale, following [348], we consider a suitable non-dimensional form of the cellular
mathematical model.
The cellular conductivity matrices ˙i .x/ and ˙e .x/ are symmetric positive
definite matrices; setting N D N i C N e , with N i , N e the average eigenvalues on
a cell element, we consider the dimensionless conductivity matrices
Cm Rm @v Rm
i Rm nTi rui D C Iion : (3.24)
N @t N
3.2 Models of Cardiac Tissue 95
xO D x=; tO D t= m:
Disregarding the presence of applied current terms, rescaling Eqs. (3.23) and (3.24)
in the intra and extracellular media, we obtain:
@v
nTi i rxO ui D nTe e rxO ue D " C Iion .v; w; c/ on m ;
@tO
where the dimensionless parameter is the ratio " D lc = between the micro and the
macro length constants.
The two-scale method of homogenization can be applied to the previous current
conservation equations. The microscopic space variable measured in unit cell
is defined by WD x= lc ; then with respect to the dimensionless macroscopic
coordinates, the micro and macro scales are related to each other by the scaling
parameter "
WD xO ="
For convenience, we will omit in the following the superscripts O of the dimension-
less variables.
Following the standard approach of homogenization theory, we are assuming
that the cells are distributed according to an ideal periodic organization similar to
a regular lattice of interconnected cylinders. Due to the longitudinal and transverse
intercellular interconnections, in the modeled periodic cellular aggregate the intra
and extracellular media are connected regions. If fe 1 ; e 2 ; e 3 g is an orthogonal basis
of R3 , we denote by
two reference open, connected and periodic subsets of R3 with Lipschitz boundary,
i.e. satisfying
Ei;e C e k D Ei;e ; k D 1; 2; 3:
nX
3 o
Y WD ˛k e k W 0 ˛k < 1; k D 1; 2; 3
kD1
96 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
is composed of the intra and extracellular volumes Yi;e D Y \ Ei;e and represents
a reference volume box containing a single cell Yi with cell membrane surface Sm D
m \ Yi .
The main geometrical assumption is that the physical intra or extracellular regions
are the "-dilation of the reference lattices Ei;e , defined as
˚ ˚
"Ei;e D " W 2 Ei;e ; with "m WD " W 2 m :
Y
Ωeε
Γmε
Yi Ωiε
Γmε
Ye Ωeε Ωeε
Sm Γmε
ξ= x Ωiε Ωiε ε
ε
ΩH ε
Fig. 3.2 Right: The ideal periodic geometry in a bidimensional section of the simplified 3–D
periodic network of interconnected cells. Left: Unit cell in the microscopic variable D x="
3.2 Models of Cardiac Tissue 97
Then the vector .u"i ; u"e ; w" ; c " /, with v" D u"i u"e satisfies the problem:
(
div i" .x/ru"i D 0 in ˝i"
(3.25)
div e" .x/ru"e D 0 in ˝e"
8
ˆ @v"
ˆ
ˆi" .x/ nT ru"i D " Œ C Iion .v" ; w" ; c " /
< @t
Im" D (3.26)
ˆ
ˆ "
:̂ " .x/ nT ru" D " Œ @v C I .v" ; w" ; c " /
e e ion
@t
@w" @c "
R.v" ; w" / D 0 on m" ; S.v" ; w" ; c " / D 0;
@t @t
98 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
v" .x; 0/ D v"0 .x/; w" .x; 0/ D w"0 .x/; c " .x; 0/ D c0" .x/ su m" :
The variables v" ; w" ; c " and Im" are defined only on the surface of the cellular
membrane m" .
We briefly indicate how to use the two–scale method (see [23, 44, 260, 362, 454])
and formal asymptotic expansions to convert the microscopic model of the periodic
cellular aggregate into an averaged continuum representation of the cardiac tissue,
neglecting the presence of stimulation currents. We seek a solution .u"i ; u"e ; w" ; c " /
where each component has an asymptotic form in powers of " of the following type:
ru D "1 r u C rx u;
div ru D "2 div r u C "1 div rx u C "1 divx r u C divx rx u;
substituting the asymptotic forms into the first equations of (3.25, 3.26) and equating
the coefficients of the powers 1; 0; 1; of " to zero, we obtain the following
equations for the functions uk .x; ; t/; k D 0; 1; 2 associated with u D u"i :
8
ˆ
ˆ
< Find uk Y –periodic in such that:
div i .x; /r uk D fk .x; / div i Fk .x; / in Ei (3.27)
ˆ
:̂nT .x; /r u D g .x; / C nT F .x; / on m ;
i k k i k
with
(
f0 D 0; F0 D 0 in Ei
(3.28)
g0 D 0 on m ;
3.2 Models of Cardiac Tissue 99
(
f1 D 2divx i r u0 ; F1 D 0 in Ei
(3.29)
g1 D nT i rx u0 on m ;
8
ˆ
ˆ f D divx i r u1 C divx i rx u0 ; F2 D i rx u1 ; in Ei
< 2
@v0
g D . C Iion .v0 ; w0 ; c0 //; on m
ˆ 2 @t
:̂
v0 D ui0 ue0 ; @t w0 R.v0 ; w0 / D 0; @t c0 S.v0 ; w0 ; c0 / D 0:
(3.30)
In problem (3.27), the variable x appears as a parameter. Let fk .x; /, Fk .x; /,
gk .x; / be Y –periodic functions in . Then the problems for k D 0; 1; 2 admit
a unique solution uk , apart from an additive constant (consequence of an easy
extension of the result in [23] Th. 2 or [362] Th. 6.1) if and only if
Z Z
fk d C gk ds D 0; k D 0; 1; 2: (3.31)
Yi Sm
From the first cellular problem (3.28) for k D 0, it follows that the Y –periodic
solution u0 is independent of and u0 .x/, depending only on the macroscopic vari-
R terms uk .x; ; t/ are
able x, represents a potential average over Yi if the subsequent
determined with zero mean value on Yi . Since f1 D 0 and Sm nT i rx u0 ds D 0
the solvability of problem (3.28) related to the data (3.29) R is assured and it is
easy to check that the solution with zero mean on Yi , i.e. Yi u1 d D 0, can be
represented as
(
div i;e .x; /r wek D 0; in Yi;e
nT i;e .x; /r wi;e
k D nk ; on S; k D 1; 2; 3:
The previous derivation based on the two-scale method is only formal, but the
averaged model can be rigorously justified in the framework of convergence
theory as a limit problem of the cellular model for " ! 0, see [4].
Find U " W Œ0; T ! V" W b " .@t U " ; UO /Ca" .U " ; UO /CF " .U " ; UO / D 0 8 UO 2 V" ;
(3.36)
where the parabolic b " and elliptic a" forms are degenerate, but their sum is coercive
on V" . The variational formulation is supplemented by the initial conditions:
v" .; 0/ D v"0 ; w" .; 0/ D w"0 ; c " .; 0/ D c0" ; on m " ;
102 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
and we introduce the following bilinear and nonlinear forms associated with the
averaged model (3.34): for U D .ui ; ue ; w; c/ 2 V; UO D .Oui ; uO e ; w;
O c/
O 2 V, and
v D ui ue ; vO D uO i uO e , we set
Z
O
b.U ; U / WD ˇ Œ@t v vO C @t w wO C @t c cO dx;
˝H
XZ
a.U ; UO / WD Di;e rui;e rui;e dx
i;e ˝i;e
Z
F .U ; UO / WD ˇ ŒIion .v; w; c/ vO R.v; w/ w
O S.v; w; c/ cO dx:
˝H
These tensors are symmetric and positive definite matrices, and the bilinear form
a.; / is coercive on V.
The variational formulation of the averaged problem P, related to the macro-
scopic model (3.34), is given by:
We now focus on the FitzHugh-Nagumo membrane model [178, 179]: the ionic
current is a cubic like function in v and is linear in the recovery variable w. In this
simplified model, the unknown is the vector .u"i ; u"e ; w" / and it was shown in [105]
that both the cellular and the averaged models share the same variational structure
and yield well posed problems. More precisely, introducing the quotient space as
in (3.35)
˚
V" D H 1 .˝i" / H 1 .˝e" / = f. ; / W 2 Rg L2 .m" /; (3.38)
Theorem 3.2. Assuming " is regular and supposing that the initial data satisfy
then there exists a unique solution U " D .u"i ; u"e ; w" / 2 C 0 .0; T I V" / of the
variational formulation (3.36) of Problem P" with
@v" @w"
; 2 L2 .0; T I L2 .m" //I
@t @t
u" WD .u" ; u" / solves the differential equations P " in the standard distributional
¯sense. i e
@v @w
; 2 L2 .0; T I L2 .˝H //I
@t @t
We now present a convergence result for the homogenization process related to the
Bidomain model with Nagumo membrane model (i.e. FHN without the recovery
variable); see [378] for details of the general FHN case.
104 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
The problem P " is not a standard parabolic homogenization problem and its
main difficulties are associated with the fact that b " depends explicitly on ", it is
degenerate and the boundaries of ˝i;e"
could be quite irregular. For z" 2 L2 . " /,
u D .ui ; ue / 2 H .˝i / H .˝e /, and z 2 L2 .˝H /, u D .ui ; ue / 2 .H 1 .˝H //2 ,
" " " 1 " 1 "
¯ define the energy-like functionals
we ¯
Z XZ
b " .z" / WD " jz" j2 d ; a" .u" / WD "
i;e ru"i;e ru"i;e dx;
m" ¯ i;e
"
˝i;e
Z XZ
b.z/ WD ˇ jzj2 dx; a.u/ WD Di;e rui;e rui;e dx;
˝H ¯ i;e ˝H
Z Z
G .z / WD "
" " "
G.z /d ; G .z/ WD ˇ G.z/dx; (3.39)
m" ˝H
with
Moreover, there exist extensions Ti " u"i ; Te" u"e of u"i ; u"e in the whole domain ˝H ,
solution of the cellular problem P" , which converge in L2 .0; T I Hloc 1
.˝H // to the
unique solution .ui ; ue / 2 V of the averaged model P.
The variational approach for the convergence of the evolution problem is based
on the introduction of the time-semidiscrete approximation by the implicit Euler
method, which reduces the evolution system to discrete families of stationary
problems. More precisely, given U "0 , we introduce the sequence of variational
problems:
Find U ";n 2 V; n D 1; : : : ; N such that
U ";n U ";n1
b". ; UO / C a" .U ";n ; UO / C F " .U ";n ; UO / D 0; 8UO 2 V:
n1 o
U ";n D arg min b " .V U ";n1 / C ˚ " .V / ; (3.40)
V 2V 2
The general strategy for passing to the limit (see [378]) can be summarized in the
following diagram
Stationary Problem
Evolution Problem → U ετ n+1 :
Pε : U ε minW 2τ b (W − Uτε n) + Φ ε (W )
1 ε
ε ↓0
Limit of the Evolution Limit of the Stationary Problem
0←τ
Problem P : U U τn+1 : minW 21τ b(W − Uτn) + Φ (W )
The macroscopic Bidomain representation of the cardiac tissue (3.34) has been
derived in the previous chapter using homogenization techniques.
We now present a heuristic derivation of the Bidomain model, based on the notion
of the interpenetrating domains introduced in [463], formalized by [199, 329, 531],
see also [225, 391]. We denote by ˝H the cardiac tissue volume that we assume
represented as the superposition of two anisotropic continuous media, the intra- (i)
and extra- (e) cellular media, coexisting at every point of the tissue and separated by
a distributed continuous cellular membrane.
Given a point x 2 ˝H , we denote by Ji .x; t/; Je .x; t/ the local average
current densities per unit area in the intra- and extracellular media and by im
the transmembrane current per unit volume flowing across the membrane surface.
Consider a volume V surrounding x and denote by jV j its volume, @V its smooth
surface, n the outward normal on @V . Then the current conservation law on the
volume V states that
Z Z Z
1 1 1
Je nd D Ji nd D cm im dx;
jV j @V jV j @V jV j V
i.e. the flux entering the extracellular volume must equal the flux exiting the
intra-cellular volume, and both must equal the transmembrane current across the
membrane (directed from the intra- to the extracellular media). Taking the limit as
jV j ! 0, it then follows
We recall that the cardiac tissue consists of an arrangement of fibers that rotate
counterclockwise from epi- to endocardium, and that have a laminar organization
modeled as a set of muscle sheets running radially from epi- to endocardium.
3.3 The Macroscopic Anisotropic Bidomain Model 107
The anisotropy of the intra- and extracellular media, related to the macroscopic
arrangement of the cardiac myocytes in the fiber structure, is described by the
anisotropic conductivity tensors Di .x/ and De .x/, defined as
Di;e .x/ D li;e al .x/aTl .x/ C ti;e at .x/aTt .x/ C ni;e an .x/aTn .x/ D
D li;e I C .ti;e li;e /at .x/aTt .x/ C .ni;e li;e /an .x/aTn .x/:
(3.41)
Here al .x/; at .x/; an .x/, is a triplet of orthonormal principal axes with al .x/
parallel to the local fiber direction, at .x/ and an .x/ tangent and orthogonal to the
radial laminae, respectively, and both being transversal to the fiber axis (see e.g.
LeGrice et al. [292]). Moreover, li;e ; ti;e ; ni;e are the conductivity coefficients
in the intra- and extracellular media measured along the corresponding directions
al ; at ; an . For axisymmetric anisotropic media ni;e D ti;e . Since the anisotropic
intra- and extracellular media are ohmic, currents and potentials are related by
and we obtain the following parabolic - parabolic (PP) formulation of the Bidomain
system.
PP-formulation. Given the applied intra- and extracellular currents per unit
s
volume Ii;e W ˝H .0; T / ! R, and initial conditions v0 W ˝H ! R, w0 W
˝H ! R , z0 W ˝H ! .0; C1/m , find the intra- and extracellular potentials ui;e W
k
where we have assumed an insulated cardiac boundary @˝H . The nonlinear reaction
term Iion and the ODE system for the gating variables w and the ionic concentrations
c are given by the ionic membrane model.
PE-formulation. System (3.42) can be equivalently rewritten in terms of the
transmembrane and extracellular potentials v.x; t/ and ue .x; t/, thus obtaining the
parabolic-elliptic (PE) formulation of the Bidomain model
8
ˆ @v
ˆ
ˆ cm div.Di rv/ div.Di rue / C Iion .v; w; z/ D Iis in ˝H .0; T /
ˆ
ˆ @t
ˆ
ˆ
ˆ div.Di rv/ div..Di C De /rue / D Iis C Ies
ˆ in ˝H .0; T /
ˆ
ˆ
ˆ
ˆ
ˆ
ˆ @w
ˆ
ˆ F.v; w/ D 0; in ˝H .0; T /
< @t
@z
ˆ
ˆ G.v; w; z/ D 0; in ˝H .0; T /
ˆ
ˆ @t
ˆ
ˆ
ˆ
ˆ nT Di r.v C ue / D 0 in @˝H .0; T /
ˆ
ˆ
ˆ
ˆ
ˆ
ˆ nT .Di C De /rue C nT Di rv D 0; in @˝H .0; T /
ˆ
ˆ
:̂
v.x; 0/ D v0 .x/; w.x; 0/ D w0 .x/; z.x; 0/ D z0 .x/; in ˝H :
(3.43)
We note that, in the previous models, we have not included many anatomical
components, such as:
• The specialized ventricular conduction system, composed of the bundle of His,
the left and right bundle branches, and an extensive Purkinje network, which
connects to the myocardium at discrete sites known as Purkinje-ventricular
junctions, see e.g. [22, 55, 425, 520];
• The presence of the coronary vasculature, see [51, 52, 490];
• The deformation of the cardiac tissue induced by the mechanical contraction and
relaxation process [232, 245, 246, 527].
Di .x/ and De .x/ are measurable and satisfy the uniform ellipticity condition
f .x/ f .y/
f .0/ D 0I 9 f 0W f ; 8x; y 2 R; with x ¤ y:
xy
(3.46)
We will also require that f has a cubic growth at infinity, i.e.
f .s/ f .s/
0 < lim inf 3
lim sup 3 < C1: (3.47)
jsj!C1 s jsj!C1 s
we seek
ui ; ue ; w W ˝H .0; T / ! R; with v WD ui ue ;
which solve
8
ˆ @v
ˆ
ˆ cm div.Di rui / C f .v/ C w D Iis in ˝H .0; T /
ˆ
ˆ @t
ˆ
ˆ
ˆ
ˆ @v
ˆ
ˆ cm div.De rue / f .v/ w D Ie
s
in ˝H .0; T /
< @t
@w (3.48)
ˆ
ˆ C w v D 0; in ˝H .0; T /
ˆ
ˆ @t
ˆ
ˆ
ˆ
ˆ nT Di;e rui;e D 0 in @˝H .0; T /
ˆ
ˆ
:̂
v.x; 0/ D v0 .x/; w.x; 0/ D w0 .x/; in ˝H ;
XZ Z
O WD
a.u; u/ Di;e rui;e r uO i;e dx C Œw.Oui uO e / .ui ue /wO dx;
i;e ˝H ˝H
(3.51)
the functionals on X
8 XZ
ˆ
ˆ O WD
< hL.t/; ui uO i;e dx;
s
Ii;e
˝
Z i;e H
(3.52)
ˆ 0
:̂ h` ; ui
O WD .cm v0 vO C w0 w/ O dx;
˝H
with domain
this means that for every u 2 D.F / there exists a constant C > 0 such that
Z
f .ui ue / dx C jjjjH 1.˝H / ; 8 2 H 1 .˝H / \ L1 .˝H /:
˝H
Again we observe that all these definitions are compatible with the quotient space V,
provided that
Z
.Iis C Ies / dx D 0 for a.e. t 2 .0; T /: (3.54)
˝H
O WD a.u; u/;
hAu; ui O O WD b.u; u/
hBu; ui O 8u; uO 2 V;
we seek u.t/ WD .ui .; t/; ue .; t/; w.; t//, with
satisfying
.Bu/.0/ D ` 0 :
It holds the following theorem, proved in [105] by applying the theory developed
in [457] for abstract evolution inequalities, based on semi-discretization in time,
a-priori error estimates and compactness properties.
Theorem 3.5. Assume that (3.54) and (3.55) are satisfied, together with
Then Problem (AM) admits a unique solution u. In particular there exist a couple
then
ui;e 2 C 0 .Œ0; T I H 1 .˝H //; @t v 2 L2 .˝H .0; T //; w 2 C 0 .Œ0; T I H 1 .˝H //:
112 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
In this section, we will review the well-posedness results for weak solutions of
the PE-formulation of the Bidomain model, established in [60] by using a Faedo-
Galerkin technique. See also [42, 59] for a similar approach, based on a parabolic
regularization technique.
The result in [60] holds for simple membrane models, which neglect the ionic
concentrations z. In the following, assume for sake of simplicity that the membrane
model takes into account only one gating variable, thus k D 1, w D w and F D F .
With the previous assumptions, the PE-formulation of the Bidomain model becomes
8
ˆ
ˆ @v
ˆ
ˆ cm div.Di rv/ div.Di rue / C Iion .v; w/ D Iis in ˝H .0; T /
ˆ
ˆ @t
ˆ
ˆ
ˆ
ˆ div.Di rv/ div..Di C De /rue / D Iis C Ies in ˝H .0; T /
ˆ
ˆ
< @w
F .v; w/ D 0; in ˝H .0; T /
ˆ @t
ˆ
ˆ
ˆ
ˆ nT Di r.v C ue / D 0 in @˝H .0; T /
ˆ
ˆ
ˆ
ˆ
ˆ n .Di C De /rue C n Di rv D 0; in @˝H .0; T /
T T
ˆ
:̂
v.x; 0/ D v0 .x/; w.x; 0/ D w0 .x/; in ˝H :
(3.56)
R by V D H .˝H /, H D L .˝H /, U D
1 2
Preliminary variational results. Denote
V =R and, for any u 2 V , Œu D u ˝H u 2 U . Consider now the following
variational problem:
find .u; ue / 2 U U such that
a.u; v/ D b..u; uQ e /; .v; 0//; < s; v >D< si ; v > ai .ue ; v/:
and ue D uQ e C ue .
We can now define the bilinear form a.; / on V V
8v 2 V a. i ; v/ D i . i ; v/:
1
where Iion 2
; Iion ; F1 W R ! R are continuous functions and F2 2 R;
(H3) There exist constants ci 0 (i D 1; : : : ; 6) such that for any v 2 R
jIion
1
.v/j c1 C c2 jvjp1
jIion .v/j c3 C c4 jvjp=21
2
(H4) There exist constants a; > 0, b; c 0 such that for any .v; w/ 2 R2
Remark 3.1. Three examples of membrane models that satisfy these assumptions
are the FitzHugh-Nagumo model [273], the Aliev-Panfilov model [1] and the Roger-
McCulloch model [422].
Existence for the initial value problem. Under the minimal regularity assumptions
s
on the data ˝H , Di;e , Ii;e and defining D.0; T / as the space of real functions C 1
on R with compact support in .0; T /, it is possible to write the
Definition 3.1. (Weak solutions). Consider > 0 and the three functions v W t 2
Œ0; / ! v.t/ 2 H , ue W t 2 Œ0; / ! ue .t/ 2 H , w W t 2 Œ0; / ! w.t/ 2 H .
Given .u0 ; w0 / 2 H , we say that .v; ue ; w/ is a weak solution of (3.56) if, for any
T 2 .0; /,
1. v W Œ0; T ! H and w W Œ0; T ! H are continuous, and u.0/ D u0 , w.0/ D w0
in H ;
2. For a.e. t 2 .0; /, we have v.t/ 2 V , ue .t/ 2 V =R, v 2 L2 .0; T I V / \ Lp .QT /
where QT D ˝H Œ0; T , and .v; ue ; w/ verify in D0 .0; T /
Z Z
d
.v.t/; vO /C Di r.v.t/Cue .t// r vO C Iion .v.t/; w.t//Ov D< Iis .t/; vO >
dt ZH
˝ ˝H
d
.w.t/; w/O C F .v.t/; w.t//wO D 0
dt ˝H
Lemma 3.1. The functions v; ue ; w are a weak solution of (3.56) if conditions 1–2
of Definition 3.1 hold and .u; w/ verify in D0 .0; T /
Z
d
.v.t/; vO / C a.v; vO / C Iion .v.t/; w.t//Ov D< s.t/; vO > 8Ov 2 V
dt Z ˝H
d
.w.t/; w/O C F .v.t/; w.t//wO D 0 8wO 2 H
dt ˝H
where a.; / and s 2 V 0 are defined in Theorem 3.6. The function ue is then
recovered from (3.60).
Theorem 3.8 (Global existence of a weak solution). Let ˝H , Di;e satisfy the
minimal regularity specified previously and suppose that the hypotheses (H1)–
(H4) on Iion and F hold forR some p 2. Let then be given u0 ; w0 2 H and
s
Ii;e 2 L2 .RC I V 0 / such that ˝H .Iis .x; t/ C Ies .x; t//dx D 0 for a.e. t > 0. Then
the system (3.56) has a weak solution .v; ue ; w/ in the sense of Definition 3.1 with
D C1.
Sketch of the proof. The proof consists of three steps:
• Construction of an approximate solution using the Faedo-Galerkin technique,
based on the special orthonormal Hilbert basis (in H ) f i gi 2N of eigenvectors
defined in Theorem 3.7;
• A priori estimates on the approximate solution;
• Compactness results and convergence of the approximate solution towards a
weak solution.
Uniqueness. Consider the function G W R2 ! R2 defined by
Iion .v; w/
G.v; w/ D 8.v; w/ 2 R2 ; for some > 0:
F .v; w/
Theorem 3.9 (Uniqueness). If the condition (U1) is satisfied, then the solution
obtained in Theorem 3.8 is unique.
A different proof based on a regularization argument has been given in [59] for
the PP-formulation of the Bidomain model. The proof holds true for the Bidomain
model coupled with the Laplace equation for the torso and, in addition to the
116 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
w 2 v
Iion .v; w/ D v .v 1/ ;
in out
1 close open
F .v; w/ D C h1 .v/ .v h1 /;
close close open
where
1 v vgate
h1 .v/ D 1 tanh
2 gate
and in ; out ; open < close ; vgate ; gate are positive constants.
The regularized finite dimensional problems used in the Faedo-Galerkin proce-
dure are constructed by adding the terms
1 @ui;n 1 @ue;n
and with n 2 N
n @t n @t
into the first and second equations of the PP-formulation of the Bidomain system,
respectively.
In this section, we will review the well-posedness result for the PP-formulation of
the Bidomain model, obtained in [541] using a fixed point argument. This result
covers a wide range of complex membrane models.
The ionic current. Assume that the ionic current
X
m
Iion .v; w; z/ WD Ji .v; w; log zi / C HQ .v; w; z/; (3.61)
i D1
3.6 Well-Posedness Results Based on Fixed Point Arguments 117
where, 8i D 1; : : : ; m,
Ji 2 C 1 .R Rk R/;
@
0 < G.w/ Ji .v; w; / G.w/;
ˇ ˇ
@ (3.62)
ˇ@ ˇ
ˇ Ji .v; w; 0/ˇ Lv .w/;
ˇ @v ˇ
Remark 3.2. The functions HQ (and Hi in the following) do not have a precise
physical meaning, but allow some currents which differ from Ji to be included in the
model, e.g. they could represent the Na-K pump and the non-specific Ca-activated
currents in [309].
The dynamics of the gating variables is described by the system of ODEs
@wj
D Fj .v; wj /; j D 1; : : : ; k: (3.64)
@t
Assume that
8j D 1; : : : ; k.
In the membrane models considered, Fj has the particular form
C1 exp. vv
C2
n
/ C C3 .v vn /
;
1 C C4 exp. vv n
C5 /
@zi
D Gi .v; w; z/ WD Ji .v; w; zi / C Hi .v; w; z/ i D 1; : : : ; m; (3.66)
@t
118 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
Integrating on ˝H and applying the divergence theorem and the Neumann boundary
conditions, we have the following compatibility condition for the system to be
solvable
Z
.Iis C Ies /dx D 0: (3.68)
˝H
Suppose then that Di .x/ and De .x/ are measurable and satisfy the uniform
ellipticity condition (3.45).
The condition on the initial datum. In view of the result of continuity for
the solution v of the Bidomain model, we must ask for the initial datum v0 to
be compatible, in a sense to specify, with the Neumann homogeneous boundary
conditions. Intuitively, if v0 D u0i u0e , then we should have
but fixing both ui .x; 0/ and ue .x; 0/, as initial data, may render the problem
unsolvable, because the time derivative involves only the difference ui ue . The
correct assumption may seem abstract, see [541, Sec. 3]: let v 2 H 1 .˝H / be given,
then the following minimization problem has a unique solution:
( )
XZ Z
min Di;e rui;e ui;e dx W ui;e 2 H .˝H /;
1
ue dx D 0; ui ue D v :
i;e ˝H ˝H
(3.70)
3.6 Well-Posedness Results Based on Fixed Point Arguments 119
Now, if Iis .0/ C Ies .0/ 2 L2 .˝H /, then the following elliptic problem admits
a unique solution u0b 2 H 2 .˝H /:
8
ˆ
ˆ div..Di C De /ru0b / D Iis .0/ C Ies .0/ in ˝H ;
<
Z i C De /rub / n D 0
0
..D on @˝H ;
ˆ
:̂ ub dx D 0:
0
˝H
We can now state the main result of [541], concerning the existence of a
variational solution for (3.42).
Theorem 3.10. Assume that
Let us also take as given the ionic currents satisfying (3.61)–(3.63), the dynamics of
the gating variables F.v; w/, satisfying (3.64) and (3.65), the dynamics of the ionic
concentrations G.v; w; z/, satisfying (3.66) and (3.67).
Then there exists a unique solution of Problem (3.42), given by k C m C 2
functions w1 ; : : : ; wk ; z1 ; : : : ; zm ; ui ; ue satisfying
Sketch of the proof. The proof of Theorem 3.10 is divided into three parts.
In a first step, one fixes v and solves the ODE systems of the gating and
ionic concentrations variables, obtaining suitable a priori estimates and qualitative
properties of the solution.
In the second step, a reduction technique is used in order to split the degenerate
parabolic system into an elliptic equation coupled with a non-degenerate parabolic
equation in L2 .˝H /, governed by the generator of an analytic semigroup. Consider-
ing Iion .v; w; z/ as a known function, we apply a result of maximal regularity in Lp ,
obtaining the existence and uniqueness of, and estimates for, the potentials ui ; ue
(and thus for v D ui ue ) in Lp .0; T I H 2 .˝H // \ W 1;p .0; T I L2 .˝H //.
These estimates, owing to classical interpolation techniques, provide a crucial
bound for v in L1 .Q/. Then, by choosing the correct functional spaces for w, z
and v, it is possible to establish the existence and uniqueness of a solution .v; w; z/
for Problem (3.42), using Banach’s Fixed Point Theorem.
The main difficulties in the parabolic equation reside in its degenerate structure,
which reflects the differences in the anisotropy of the intra- and extracellular tissues,
and in the lack of a maximum principle. Moreover, the concentration variables zi
appear as arguments of a logarithm, both in the dynamics of the concentrations and
in the ionic currents, and therefore it is necessary to bound z far from zero.
uτ (t)
U n U n1 O
b. ; U / C a.U n ; UO / C F .U n ; UO / D 0; 8UO 2 Vh : (3.72)
n
Considering the discrete solution U .t/ given by the continuous piecewise linear
function interpolating the values fU n gN
nD0 on the grid P, we have the following
error estimate
Theorem 3.12. For sufficiently regular initial data, the following a priori error
estimate between u and U , measured by the natural variational (semi)norms holds:
Z
T
e2 WD max b U .t/ U .t/ C a U .t/ U .t/ dt C 2
;
t 2.0;T / 0
or equivalently
p
k b.U U /kL1 .0;T / C kU U kL2 .0;T IV/ C ;
with C independent of :
We conclude presenting a result related to a posteriori error estimates. In [32],
resorting to the theory developed in [359] for evolution variational problems,
a posteriori error estimates were derived for general degenerate evolution equations.
122 3 Mathematical Models of Cardiac Cells Arrangements: The Bidomain Model
Theorem 3.13. For sufficiently regular initial data, let e be the error between
U and U measured by the natural variational (semi)norm
h Z
T i
e2 WD max max e 2 gt
b U .t/ U .t//; e 2 gt
a U .t/ U .t/ dt
t 2.0;T / 0
with g D infv2R g 0 .v/ .
Then applying the theory of [32] to the Bidomain model, the error e can be
estimated a posteriori by
X 2 X q
g
e2 n Dn
2
C 2
n b ıU n ;
n
2 n
where Dn D a U n U n1 C F U n ; U n U n1 F U n1 ; U n U n1 ;
U n U n1
ıU n D :
n
Chapter 4
Reduced Macroscopic Models:
The Monodomain and Eikonal Models
In the Bidomain model (3.42), the transmembrane potential v during the excitation
phase of the heartbeat exhibits a steep propagating layer spreading throughout the
myocardium with a thickness of about 0.5 mm. At every point of the cardiac domain,
this upstroke phase lasts about 1 ms. Therefore, the simulation of the excitation
process requires the numerical solution of problems with small space and time steps
(of the order of 0.1 mm and 0.01 ms). A further inconvenient of the Bidomain model
is its severe ill-conditioning, mainly due to the pure Neumann boundary conditions,
of the linear systems deriving from its discretization. The solution of such linear
systems with iterative methods such as the Conjugate Gradient, preconditioned
by standard cheap preconditioners (ILU or SSOR), is not effective, thus more
sophisticated computational demanding preconditioners are required (see Chaps. 7
and 8).
All these facts constraint 3–D simulations to limited blocks with dimensions of
a few cm, see [101, 102, 104, 226]. For large scale simulations involving the whole
ventricles, the computer memory and time requirements become excessive and less
demanding approximations have been developed, such as Monodomain and Eikonal
models.
In order to reduce further the computational load many large scale simulations have
been performed using the so-called Monodomain model; it is well known that if
the two media have the same anisotropy ratio then the Bidomain system reduces
to the Monodomain model. We remark that this is not a physiological case, as it
clearly follows from well established experimental evidence. We shall present an
interesting derivation of a reduced Bidomain model, which does not make such a
priori assumption (see also [99,269]) and that we will still call Monodomain model.
Denoting by Jtot D ji C je the total current flowing in the two media and by
D D Di CDe the conductivity of the bulk medium, since Jtot D Di rui De rue ,
substituting ui D v C ue , we obtain
Therefore, the second equation in the Bidomain system (3.42) can be written as
@v
cm C div.De D 1 Di rv/ C div.De D 1 Jtot / iion .v; w; c/ D Iapp
e
: (4.2)
@t
Since the conductivity tensors are given by (3.41), then
De D 1 D el I C .et el /at .x/aTt .x/ C .en el /an .x/aTn .x/; (4.3)
div.De D 1 Jtot / D el div Jtot C .et el / divŒat .x/aTt .x/Jtot
C.en el / divŒan .x/aTn .x/Jtot
(4.4)
D el .Iapp
i
C Iapp
e
/ C .et el / divŒat .x/aTt .x/Jtot
C.en el / divŒan .x/aTn .x/Jtot :
Using the split form (4.3), the first term on the right hand side can be written as
nT De D 1 Di rv D 0: (4.7)
We remark that for media having equal anisotropic ratio, i.e. le =li D te =ti D
ne =ni , we have el D et D en . Thus, the two additional terms in (4.4), related
to the projections of Jtot on the directions across fiber, disappear. Disregarding
these two additional source terms aTt Jtot and aTn Jtot , we have div.De D 1 Jtot /
4.2 Eikonal Models 125
el .Iapp
i
C Iapp
e
/. Substituting this approximation in (4.2) and considering the
boundary condition (4.7), we obtain an approximate model consisting in a single
parabolic reaction-diffusion equation for v and coupled with the same gating system
8
ˆ @v
ˆ
ˆ cm div.Dm rv/ C iion .v; w; c/ D Iapp
m
in ˝H .0; T /
ˆ
ˆ @t
< @w @c
R.v; w/ D 0; S.v; w; c/ D 0 in ˝H .0; T /
ˆ
ˆ @t @t
ˆ
ˆ n Dm rv D 0
T
in H .0; T /
:̂
v.x; 0/ D v0 .x/; w.x; 0/ D w0 .x/; c.x; 0/ D c0 .x/ in ˝H ;
(4.8)
i e
with the conductivity tensor Dm D De D 1 Di , Iapp
m
D Iapp l Iapp
e
li =.le C li /.
The evolution equation determines the distribution of v.x; t/ and then the
extracellular potential distribution ue is derived by solving the elliptic boundary
value problem
(
div.Drue / D div.Di rv/ C Iapp
i
C Iapp
e
in ˝H ;
(4.9)
n Drue D n Di rv on H :
T T
We refer to the system consisting of Eqs. (4.8) and (4.9) as the anisotropic
Monodomain model. We remark that the Bidomain and the Monodomain model
are described by a system of a parabolic equation coupled with an elliptic equation,
but in the latter the evolution equation is fully uncoupled with the elliptic one in the
case of an insulated domain ˝H .
Another route to avoid the high computational costs of the full Bidomain model is
based on the use of eikonal models for the evolution of the excitation wavefront
surface.
With these models the simulation of the activation sequence in large volumes of
cardiac tissue has become computationally practical but at the price of a loss of fine
details concerning the thin layer where the upstroke of the action potential occurs.
These numerical simulations are based on laws describing the macroscopic kinetic
mechanism of the spreading of the excitation wavefronts, and do not require a fine
spatial and temporal resolution.
We now outline the derivation of this kind of approximated models. The
FitzHugh–Nagumo approximation of the membrane kinetics is very useful for a
qualitative analysis of the non-linear dynamics of the R-D system. As we shall focus
only the excitation phase, we can neglect the recovery variable w hence iion D g.v/.
126 4 Reduced Macroscopic Models: The Monodomain and Eikonal Models
The resulting simplified ionic model is widely used for gaining general insight
into wave propagation in the cardiac excitable medium. Although this model is not
suitable in a quantitative detailed study, at a fine scale, of the upstroke of the action
potential v through the excitation wavefront, it is nevertheless appropriate if we are
interested in the large scale behavior of the front-like solution. We note that during
the excitation phase of the heart beat the main feature, at a macroscopic level, is
the excitation wavefront configuration and its motion. In order to investigate the
propagation of this wavefront we must analyze more deeply the internal layer of v
which affects the spreading.
Denoting by the membrane surface area per unit volume of the tissue, by Cm
the membrane capacitance per unit area of the membrane surface, and by Iion the
ionic current per unit area of the membrane surface, that is carried by the flow of
ions across the membrane, then the transmembrane current im per unit volume is
given by
@v
im D Cm C Iion :
@t
We recall that the so-called fast sodium current INa is the main current responsible
for the depolarization of cardiac cells during the excitation phase of the transmem-
brane action potential. Experimental findings in measuring INa in different cardiac
preparations justify the modeling of the sodium current mechanisms by means
of gating equations of the Hodgkin-Huxley type, see e.g. [193, 308] for guinea
pig, [367] for rat, [473] for rabbit, [209, 521] for human myocytes. Hence INa is
represented by
In the depolarization phase the time constant of the activation gate m for ventricular
cells is about 0.2 ms, whereas the inactivation time constants are of order 50
and 300 ms for h and j gates, respectively. Thus, we assume that the membrane
obeys the simplified time-independent ionic current-voltage relation, obtained by
considering an instantaneous change of sodium conductance, i.e. taking m D
m1 .v/, h D h1 .vr / ' 0:98, j D j1 .vr / ' 0:99, where vr is the resting potential.
In this simplified approximation, the ionic current model is given by the non-linear
current-voltage law
ZOOM
200 1
−200
−400 0
Iion ( A/cm2)
(μ A/cm2)
−600
μ
−800
ion
I
−1000 −1
−1200
−1400
−1600 −2
−100 −80 −60 −40 −20 0 20 40 60 −90 −85 −80 −75 −70 −65 −60
v (mV) v (mV)
Fig. 4.1 Cubic-like behavior of the ionic current per unit area of membrane surface Iion (A=cm2 )
as a function of the transmembrane potential v (mV)
with
where IL is the leakage current. Using the GN Na , vNa and m1 .v/ expressions in the
sodium current of the LR1 model [308] and choosing GN L D 0:05 mS=cm2, Vl D
80 mV, the behavior of Iion .v/ is of cubic-like type as displayed in Fig. 4.1, i.e. it
has the following properties:
1. Iion 2 C 1 .R/;
2. There exist only three zeros vr < vth < vp ;
0 0
3. RIion
vp
.vr / > 0 and Iion .vp / < 0;
4. vr Iion .v/d v < 0.
x t
xO D tO D ;
L T
and setting
Iion
gN Na D GN Na g.v/ D ;
gN Na
i;e
and recalling that the conductivity coefficients l;t;n are in the range 0.1–2 mS=cm2 ,
we obtain
i;e
cm l;t;n i;e
D m; D 2 O l;t;n ;
gN Na T gN Na L2
In the following, we will denote by v" D vO , u"i;e D uO i;e , and for the sake of simplicity
in the notations, we will drop all theOsymbols out elsewhere.
The R-D systems with excitable dynamics are studied by mathematical tools
from singular perturbation theory, see e.g. [174, 273]. Given the previous singular
perturbation structure, u"i , u"e diffuse quite slowly, while the reaction takes place
much faster; hence, the development of a moving layer associated with a traveling
wavefront solution is to be expected. Exploiting the singular perturbation approach,
4.2 Eikonal Models 129
We define the activation time " .x/ as the time instant at which the potential v"
at x reaches the value .vr C vp /=2. Then the excitation wavefront S" .t/ is also
represented by the level surface of the activation time at the time instant t, i.e.:
Let
which gives the conductivity measured along the direction of the bulk medium
composed by coupling in series the media (i) and (e). Then we introduce the
following indicatrix function
p
˚.x; / D q.x; /: (4.18)
Under the assumptions 1, 2, 3, 4 let (c,a) be the unique bounded solution of the
eigenvalue problem (see Sect. 3.1.3):
a00 C c a0 C g.a/ D 0
(4.19)
a.1/ D vp ; a.1/ D vr ; a.0/ D .vp C vr /=2:
130 4 Reduced Macroscopic Models: The Monodomain and Eikonal Models
A formal inner asymptotic expansion in powers of " of .u"i ; u"e / solution of (4.12)
and v" D u"i u"e can be performed using two different types of variable stretching.
Let be the Euclidean unit vector normal to the wavefront S" .t/ oriented toward
the resting tissue and for s.t/ 2 S" .t/ we define the vector n˚ .s/ D ˚ .s; /. As
in [39] we use a Lagrangian point of view and we consider the moving reference
.s; y; / defined by
yD ; x D s.t/ C n˚ .s.t//; D t; 8s.t/ 2 S" .t/; (4.20)
"
i.e. stretching the space coordinate along the n˚ direction.
Using the moving frame (4.20) the asymptotic expansion for the Bidomain model
(4.12) shows that (see [39], Appendix B), at least formally, the front associated with
(4.13) moves along the relative normal vector n˚ with velocity .n˚ / given at any
s.t/ 2 S" .t/ by:
c " div n˚
" .n˚ / D C O."2 /; (4.21)
m
˚.s; /
./ D .c " div ˚ .s; // C O."2 /: (4.22)
m
˚.s; /
./ D .c " div ˚ .s; //: (4.23)
m
t .x/
D x; D : (4.24)
"
4.2 Eikonal Models 131
Using the fixed Eulerian frame (4.24) and developing the asymptotic expansion
of the Bidomain model (4.12), we obtain the following expression, equivalent to
formulae (59) and (61) derived in [113]:
1
m C " div ˚.x; r " /˚ .x; r "/ D c C O."2 /: (4.25)
˚.x; r "/
Since
r " 1
D and " ./ D ;
jr "j jr " j
we obtain
" ./ ˚.x; /˚ .x; /
m C " div D c C O."2 /; (4.26)
˚.x; / " ./
or equivalently
" ./ ˚.x; / " ./
m D c " div ˚ .x; / C " r ˚ .x; / C O."2 /:
˚.x; / " ./ ˚.x; /
(4.27)
" ./
Since both Eqs. (4.22)–(4.27) imply D c C O."/, then the two eikonal
˚.x; /
equations (4.22), (4.27) are equivalent up to second order terms. Equations (4.23)
and (4.27), disregarding the O."2 / term, are called respectively eikonal–curvature
and eikonal–diffusion equations, [121, 522].
The rigorous justification of the connection between the evolution of a suitable
level set of v and the surface flowing under geometric evolution law, remains
to our knowledge an open problem. A partial rigorous characterization in the
convergence framework was obtained for the stationary Bidomain model in [4].
We introduce the family of vectorial integral Lyapunov functionals dependent on
the couple u" WD .ui ; ue /
Z
F " .u/ WD " Di rui rui C De rue rue dx
˝
Z (4.28)
1
C G.ui ue / dx;
" ˝
where G 0 D g.
The degenerate reaction-diffusion system associated with (4.12) in the couple of
unknowns u WD .u"i ; u"e / can be obtained by taking the gradient flow of the Lyapunov
functional with respect to the positive but degenerate bilinear form in L2 .˝I R2 /
132 4 Reduced Macroscopic Models: The Monodomain and Eikonal Models
Z
b.u; w/ WD .ui ue / .wi we / dx; u D .ui ; ue /; w D .wi ; we /:
˝
We can easily see that, rescaling as in (4.12) the reaction-diffusion equation related
to the Monodomain model (4.8) and using formal asymptotic expansions as before,
the anisotropic evolution law of the front does not coincide with that derived from
the Bidomain model. In fact, although the eikonal-curvature equation up to terms of
order O."2 / presents the same structure
˚.s; /
./ D .c " div ˚ .s; //; (4.30)
m
p
the non-linear function ˚.x; / D q.x; / for the Monodomain model is defined
by
@v 1
g.v/ " div .Q.x; rv/rv/ D Iapp ; (4.32)
@t "
with Q.x; / defined in (4.17) and written explicitly in terms of the conductivity
tensors as
!2 2
T Di .x/ T De .x/
Q.x; / WD De .x/ C Di .x/; D WD Di C De :
T D.x/ T D.x/
(4.33)
Recalling the definition of the function ˚ in (4.18), the motion of the excitation
wavefront is described by the following kinetic equation
1
./ D ˚.x; / div ˚ .x; / ; (4.34)
cm
which is the dimensional form of Eq. (4.23). In terms of the activation time .x/,
r
since D jr j
and thanks to the homogeneity properties of ˚ and ˚ , we obtain
the
134 4 Reduced Macroscopic Models: The Monodomain and Eikonal Models
1
˚.x; r .x// div ˚ .x; r .x// C ˚.x; r .x// D 1 in ˝H :
cm
(4.35)
An equivalent formulation of (4.34) can be derived using the level set approach,
in terms of the zero level set of a function w.x; t/. Setting the depolarized cardiac
region at time t by
and considering a function w.x; t/ positive inside Hd .t/ and negative outside, the
excitation wavefront S.t/ at time t can be represented as
rw
D :
jrwj
@t w
./ D ;
T rw
Eq. (4.34) becomes
@w 1
D ˚.x; rw/ C div ˚ .x; rw/ : (4.36)
@t cm
1
div ˚.x; r .x//˚ .x; r .x// C ˚.x; r .x// D 1 in ˝H :
cm
(4.37)
thus
1 p i e
./ D D ˚./ D ; with D :
jr j i C e
coupled with the membrane model considered. Then, estimating the steady velocity
of the action potential along the cable and the conductivity coefficient , we derive
D p .
While we have focused so far on the use of reduced models for the excitation
phase, we note that the linear anisotropic Monodomain model and the relaxed
Monodomain model could also be used as reduced models in all phases of the
heartbeat. These two models in dimensional form are
Linear Anisotropic Monodomain model
8
ˆ @v
ˆ
ˆ cm div.Dm rv/ C iion .v; w; c/ D Iapp m
in ˝H .0; T /
ˆ
ˆ @t
< @w @c
R.v; w/ D 0; S.v; w; c/ D 0 in ˝H .0; T /
ˆ T
ˆ @t @t
ˆ
ˆ n Dm rv D 0 in H .0; T /
:̂
v.x; 0/ D v0 .x/; w.x; 0/ D w0 .x/; c.x; 0/ D c0 .x/ in ˝H ;
(4.38)
From (3.45), it follows that there exist c1m ; c2m > 0 such that
This property, coupled with the assumptions of Theorem 3.10, guarantees the well-
posedness of the Linear Anisotropic Monodomain model (4.38).
Assuming that 8x 2 ˝H the map ! ˚ 2 .x; / is uniformly strongly convex,
i.e. there exists c > 0 independent of x such that ˚ 2 .x; / c2 jj2 is convex, then
the mapping ! ˚.x; /˚ .x; / is uniformly strongly monotone, i.e.
This monotonicity property, coupled with the assumptions of Theorem 3.10 on the
non-linear reaction terms of FitzHugh-Nagumo type, ensures the well-posedness of
the Relaxed Monodomain model (4.39).
Considering the Eikonal-Diffusion equation (4.37) with mixed Dirichlet-
Neumann boundary conditions, the same monotonicity property should guarantee
the well-posedness.
Regarding the level set formulation (4.36) of the Eikonal-Curvature model, we
remark that the mapping associated to the right hand side of (4.36) is continuous,
degenerate elliptic and geometric. Moreover, it is easy to verify that the map ˚.x; /
satisfies the properties (6.1) and (6.2) of [37]. Thus, proceeding as in [37], one
can prove that Eq. (4.36) admits a unique continuous viscosity solution by applying
Theorem (4.9) of [204], see also [27, 202, 203].
We recall that differently from to the Relaxed Monodomain and Eikonal models,
the convexity assumption on the map ˚ 2 is not needed for the well-posedness of
the Bidomain model, and the only assumption required on the tensors Di;e is the
uniform ellipticity (3.45).
We refer to [3,40] for works investigating a generalized multidomain system and
eikonal models with non-convex indicatrix function ˚.
We have seen before that the solvability of both the eikonal curvature and diffusion
equations, as well as of the relaxed Monodomain model, is guaranteed if the function
! ˚ 2 .x; / is strictly convex. Such property depends on the conductivity tensors
Di .x/, De .x/. After some relatively easy computations, this strict convexity is
equivalent to the following property:
4.4 Dimensional Form of the Reduced Models 137
where
pT Di .x/p pT De .x/p
!.x; p/ WD T
De .x/ p T Di .x/ p; D WD Di C De ;
p D.x/p p D.x/p
T 2 T 2
p Di .x/p p De .x/p
Q.x; p/ WD De .x/ C Di .x/:
pT D.x/p pT D.x/p
where
Di;e .x/ D diag.li;e .x/; ti;e .x/; ni;e .x//:
Since A.x/ is an orthogonal matrix, then it easy to verify that the previous property
(4.40) is equivalent to
min T D .x/ 4! .x; /T ŒQ .x; /1 ! .x; / > 0 8x 2 ˝H ; (4.41)
jjjj2 D1
and
(4.42)
Therefore the property (4.42) depends on the six parameters li;e .x/; ti;e .x/; ni;e .x/.
Assuming axially symmetric anisotropy, F depends on the anisotropy ratios
defined by
li le ti
i D ; e D ; rt D :
ti te te
138 4 Reduced Macroscopic Models: The Monodomain and Eikonal Models
In order to visualize the convexity property, we introduce the unit sphere of the
indicatrix function ˚.x; / at x
called Frank diagram, and the unit sphere associated with the dual function ˚
usually called the Wulff form of the anisotropy in cristal growth and multiphase
problems.
Neglecting the dependence on x, the boundary of B˚ is given by the set
@B˚ D 2 Rn W D ; jjjj2 D 1 :
˚./
implying that
Thus
Q./
@B˚ D 2 Rn W D ˚ D ; D ; jjjj2 D 1 :
˚./ ˚./ ˚./
In Figs. 4.2–4.7, we display the Frank diagram and Wulff form using a nom-
inal anisotropy and various estimates of the conductivity coefficients taken from
[100, 123, 131, 420, 421] and related to a normal cardiac tissue. From the graphics,
we see that all these estimates guarantee the convexity of ˚ and ˚ . In Fig. 4.8,
we display the Frank and Wulff diagrams for the extreme case called reciprocal
anisotropy (i D 1=e ), where it is evident that convexity is lost. We remark that
at the concave portion of the Frank diagram corresponds a caustic in the Wulff
diagram. We recall that the conductivity coefficients are average values over a unit
volume of tissue and in particular ti is related to the density of the gap-junctions
in the across fiber direction. Particularly in a pathological tissue as in presence
of myocardial ischemia, the decreasing number of across gap-junctions yields a
reduction of ti . Considering the conductivity values in Fig. 4.6, but with ti reduced
in order to have i D 20, the diagrams displayed in Fig. 4.9 show that the two sets
are not convex, as clearly evidenced by the presence of caustics in the Wulff form,
see Fig. 4.10.
4.4 Dimensional Form of the Reduced Models 139
15
10
150
5
100
0
50
0 −5
−50
−10
−100
−150 −15
−250 −200 −150 −100 −50 0 50 100 150 200 250 −10 −5 0 5 10
Fig. 4.2 Frank and Wulff diagrams for the conductivity values li D 0:2, ti D 0:02, le D 0:8,
te D 0:2 (all in mS cm1 )
40
50 20
−20
−50 −40
−80 −40 0 40 80 −20 −10 0 10 20
Fig. 4.3 Frank and Wulff diagrams for the conductivity values li D 1:74, ti D 0:193, le D
6:52, te D 2:36 (all in mS cm1 ) (From [100])
140 4 Reduced Macroscopic Models: The Monodomain and Eikonal Models
40
50 20
−20
−50 −40
−80 −40 0 40 80 −20 −10 0 10 20
Fig. 4.4 Frank and Wulff diagrams for the conductivity values li D 2:78, ti D 0:263, le D
2:94, te D 1:33 (all in mS cm1 ) (From [421])
40
40 20
20
0
−20
−20
−40 −40
−60 −40 −20 0 20 40 60 −30 −15 0 15 30
Fig. 4.5 Frank and Wulff diagrams for the conductivity values li D 3:43, ti D 0:596, le D
1:17, te D 0:802 (all in mS cm1 ) (From [420])
4.4 Dimensional Form of the Reduced Models 141
40
20
40
20 0
−20
−20
−40 −40
−60 −40 −20 0 20 40 60 −20 −10 0 10 20
Fig. 4.6 Frank and Wulff diagrams for the conductivity values li D 3, ti D 0:315, le D 2,
te D 1:35 (all in mS cm1 ) (From [123])
60
40
30
20
15 0
0 −20
−15 −40
−30 −60
−50 −25 0 25 50 −20 0 20
Fig. 4.7 Frank and Wulff diagrams for the conductivity values li D 1:55, ti D 0:243, le D
2:32, te D 1:04 (all in mS cm1 ) (From [131])
142 4 Reduced Macroscopic Models: The Monodomain and Eikonal Models
10
300
200
5
100
0 0
−100
−5
−200
−300 −10
−300 −200 −100 0 100 200 300 −10 −5 0 5 10
Fig. 4.8 Frank and Wulff diagrams for the conductivity values li D 0:2, ti D 0:02, le D 0:02,
te D 0:2 (all in mS cm1 )
40
20
40
0
20
0 −20
−20
−40 −40
−80 −60 −40 −20 0 20 40 60 80 −20 −10 0 10 20
Fig. 4.9 Frank and Wulff diagrams for the conductivity values li D 3, ti D 0:15, le D 2,
te D 1:35 (all in mS cm1 )
4.5 Numerical Comparison 143
32
31.5
31
−9.5 −9 −8.5 −8
All values are given in 1 cm1 . The membrane model considered is the Luo-
Rudy I model [308]. The external stimulus of 200 mA=cm3 lasting 1 ms is
applied in a small volume of about 0:04 0:04 0:02 cm3 at the center of
the epicardial surface. The computed AT is defined at each point in space x
as the unique instant ta during the upstroke phase of the action potential when
v.x; ta / D 50 mV.
The space discretization of all the four models considered is performed by Q1
finite elements in space and semi-implicit finite differences in time; for details see
Chap. 7. The AT considered as a reference solution is computed by solving the
Bidomain model on a mesh of 512 512 128 finite elements, yielding a total
amount of 67 897 602 degree of freedom (dof).
In Table 4.1, we report the l 2 relative errors of the AT computed by solving
the Bidomain, Linear Monodomain and Relaxed Monodomain models on four
increasing Q1 finite element meshes with respect to the reference AT. At the finest
level, the errors of the Linear Monodomain and Relaxed Monodomain models are
about 4:5 and 1:5 %, respectively.
The Eikonal-Diffusion equation is solved on a very coarse mesh of 64 64 16
elements, with a total amount of only 71 825 dof. The relative error in l 2 norm
between the AT computed by the Eikonal-Diffusion equation and the reference AT
is 0.0772. The AT maps reported in Figs. 4.11–4.14 confirm that, even on a such
coarse mesh, the Eikonal-Diffusion equation is able to approximate very accurately
the AT.
Table 4.1 Relative errors in l 2 norm of the activation time computed by solving the Bidomain,
Linear Monodomain and Relaxed Monodomain models on four increasing Q1 finite element
meshes with respect to the reference activation time computed by solving the Bidomain model
on a mesh of 512 512 128 finite elements. The degrees of freedom (dof) are also reported. Note
that in the case of the Bidomain model the dof are doubled because the unknowns are both the
transmembrane potential v and extracellular potential ue . The Bidomain dof for the computation of
the reference solution are 67 897 602
Mesh dof Bidomain Linear monodomain Relaxed monodomain
128 128 32 549 153 0.0301 0.0649 0.0387
192 192 48 1 825 201 0.0198 0.0493 0.0239
256 256 64 4 293 185 0.0129 0.0463 0.0185
384 384 96 14 377 825 0.0047 0.0471 0.0158
4.5 Numerical Comparison 145
a EPI
b EPI
22
20
0.2 0.2 20
18
0 16 0
15
14
−0.2 12 −0.2
10 10
−0.4 −0.4
8
6
−0.6 −0.6
5
4
−0.8 2 −0.8
0
−0.5 0 0.5 −0.5 0 0.5
c EPI
d EPI
22
22
20
0.2 20 0.2
18
18
0 0 16
16
14
14
−0.2 −0.2 12
12
10
10
−0.4 −0.4
8 8
6 6
−0.6 −0.6
4 4
−0.8 2 −0.8 2
Fig. 4.11 Epicardial activation time isochrones computed by solving the Bidomain model (a), the
Linear Monodomain model (c) and the Relaxed Monodomain model (d) with a mesh of 384
384 96 finite elements, and the Eikonal-Diffusion equation (b) with a mesh of 64 64 16
finite elements (b). Below each panel are reported the minimum, maximum and step in mV of the
displayed map and the colorbar denotes the range of values of the displayed equipotential lines
146 4 Reduced Macroscopic Models: The Monodomain and Eikonal Models
a MID
b
MID
18 20
0.2 0.2
16
0 0
14
15
−0.2 −0.2
12
−0.4 10 −0.4
10
8
−0.6 −0.6
−0.8 −0.8
5
−0.3 −0.2 −0.1 0 0.1 0.2 0.3 −0.3 −0.2 −0.1 0 0.1 0.2 0.3
c MID
d MID
20
20
0.2 0.2
0 0
15
15
−0.2 −0.2
−0.4 −0.4
10
10
−0.6 −0.6
−0.8 −0.8
5 5
−0.3 −0.2 −0.1 0 0.1 0.2 0.3 −0.3 −0.2 −0.1 0 0.1 0.2 0.3
Fig. 4.12 Mid-myocardial activation time isochrones computed by solving the Bidomain model
(a), the Linear Monodomain model (c) and the Relaxed Monodomain model (d) with a mesh of
384 384 96 finite elements, and the Eikonal-Diffusion equation (b) with a mesh of 64 64 16
finite elements (b). Below each panel are reported the minimum, maximum and step in mV of the
displayed map and the colorbar denotes the range of values of the displayed equipotential lines
4.5 Numerical Comparison 147
a ENDO
b
ENDO
0.3 0.3
0.2 20 0.2 22
0.1 0.1
20
18
0 0
−0.1 −0.1 18
16
−0.2 −0.2
16
−0.3 14 −0.3
−0.4 −0.4 14
12 −0.5
−0.5
12
−0.6 −0.6
10
−0.7 −0.7 10
c ENDO d ENDO
0.3 0.3
22
0.2 0.2 20
20
0.1 0.1
18
0 0
18
−0.1 −0.1
16
−0.2 16 −0.2
−0.3 −0.3
14
14
−0.4 −0.4
−0.5 −0.5 12
12
−0.6 −0.6
10 10
−0.7 −0.7
Fig. 4.13 Endocardial activation time isochrones computed by solving the Bidomain model (a),
the Linear Monodomain model (c) and the Relaxed Monodomain model (d) with a mesh of 384
384 96 finite elements, and the Eikonal-Diffusion equation (b) with a mesh of 64 64 16
finite elements (b). Below each panel are reported the minimum, maximum and step in mV of the
displayed map and the colorbar denotes the range of values of the displayed equipotential lines
148 4 Reduced Macroscopic Models: The Monodomain and Eikonal Models
a 22
b
20
0.2 0.2 20
18
0 16 0
15
14
−0.2 12 −0.2
10 10
−0.4 −0.4
8
6
−0.6 −0.6
5
4
−0.8 2 −0.8
0
0.2 0.4 0.2 0.4
0.11 22.49 1.00 0.00 23.20 1.00
c d
22
22
20
0.2 20 0.2
18
18
16
0 16 0
14
14
−0.2 −0.2 12
12
10 10
−0.4 −0.4
8 8
6 6
−0.6 −0.6
4 4
−0.8 2 −0.8 2
Fig. 4.14 Transmural activation time isochrones computed by solving the Bidomain model (a),
the Linear Monodomain model (c) and the Relaxed Monodomain model (d) with a mesh of 384
384 96 finite elements, and the Eikonal-Diffusion equation (b) with a mesh of 64 64 16
finite elements (b). Below each panel are reported the minimum, maximum and step in mV of the
displayed map and the colorbar denotes the range of values of the displayed equipotential lines
Chapter 5
Anisotropic Cardiac Sources
The electrical activity of the heart is revealed and usually detected by measuring the
time–dependent extracellular or extracardiac potential u.x; t/.
Definition 5.1. The electrogram (EG) at a point x inside or outside the myocardium
is defined as the time course of the potential u.x; t/, i.e. with x fixed and t variable.
In electrocardiography what is really measured is a difference between the
potential at an observation point and a reference potential. In the usual practice
body surface, epicardial, intramural and intracavitary EGs are recorded against
the potential average on a given set of points or against the potential at a distant
electrode. The classical reference potential is the average of the potentials at three
different points on the torso (Wilson Central Terminal), see Sect. 1.5. More generally
we can consider as reference potential the average potential on a given surface.
In this study we choose the reference potential which yields zero average on the
insulated boundary of the volume conductor where the heart is embedded (see
[510]).
As described in Sect. 3.3, the representation at macroscopic level of the cardiac
tissue is given by the anisotropic Bidomain [225], constituted by two interpene-
trating anisotropic continua, intra- and extracellular, connected by the distributed
cellular membrane. In the Bidomain model the following potentials are consid-
ered:
• ui .x; t/; ue .x; t/ intra- and extracellular potential;
• v.x; t/ D ui .x; t/ ue .x; t/ transmembrane potential;
• u0 .x; t/ the extracardiac potential;
then the potential u.x; t/ is given by ue .x; t/ inside the myocardium and by u0 .x; t/
outside. The bioelectric sources generating the electric potential field in the heart
(extracellular potential ue ) and outside (extracardiac potential u0 ) are related to
the gradient of the transmembrane potential distribution v.x; t/, as detailed in the
following of this chapter.
We recall that the current vector densities associated to the intra, extracellular
and extracardiac potentials are given by Ji D Di rui , Je D De rue and
J0 D D0 ru0 . Assuming that there are no externally applied sources, then the
total field Ji C Je in the cardiac tissue ˝H must be solenoidal, see Sect. 3.3; the
same holds for J0 since there are no sources in ˝0 . Hence we have:
If ˙ D @˝H \ @˝0 denotes the parts of the epi and/or endocardium in contact with
˝0 , then the flux continuity across ˙ implies the transmission conditions:
ue D u0 .Ji C Je / n D J0 n on ˙:
.Ji C Je / n D 0 on H J0 n D 0 on 0 :
and we define the equivalent cardiac source for the extracardiac and extracellular
potentials by
n
n Η
ΓH
Ω0
Γ
0
Σ
where ŒŒ ' ˙ denotes the jump of ' through ˙, i.e. ŒŒ ' ˙ D '˙ C 'j˙ with
'j ˙ the traces taken on the positive and negative side of ˙ with respect to the
˙
oriented normal. Therefore problem (5.4) provides the extracellular or extracardiac
potential u.x; t/ from the knowledge of v.x; t/.
In the remaind of this Chapter we will assume:
• ˝ and ˝H bounded Lipschitz open set of R3 ;
• The conductivity tensors Di .x/; De .x/ 2 C 1 .˝H /;
• 8t 2 Œ0; T the transmembrane potential v.x; t/ 2 W 2;p .˝/ with p > 3.
These regularity properties imply that Jv 2 W 1;p .˝H / and div Jv 2 Lp .˝H /,
hence n Jvj@˝ 2 W 1=p;p .@˝H / and the first Green formula holds, see e.g. [305].
H
The conormal derivative, in the boundary value problem (5.4), exhibits a jump
discontinuity on ˙. Defining the following bilinear form
Z
a.'; / D O
.D.x/r'/ r d x; 8'; 2 W 1;2 .˝/ (5.5)
˝
We remark that 8t 2 Œ0; T problem (5.6) admits a solution u.x; t/ 2 W 1;2 .˝/
unique up to an additive constant. This constant depends on t and it is related to the
reference potential chosen.
It is easy to verify, applying the Green formula, that the solution of (5.6) satisfies
problem (5.4) in the sense of distributions and of the trace theorems (see e.g. [305]).
1
˝ D ı. x/; D when x 2 ˝
j j (5.10)
1
˝ D 0; D ı. x/ when x 2 ;
j j
O
Taking into account that, if x; x0 … ˙, then ŒŒ .Dr / n ˙ D 0 otherwise, if
O
x 2 ˙, ŒŒ .Dr / n ˙ D ı. x/, from (5.11) we have:
Z Z
u.x; t/ C O
div Dru d O
.Dru/ n d
˝0 [˝H
Z
O
ŒŒ .Dru/ n ˙ d uref D 0;
˙
Z
1
where uref D u.x0 ; t/ or uref D u d according as .I x/ satisfies (5.9)
j j
or (5.10). Hence taking into account (5.4) and denoting by w.x; t/ the potential
difference with respect to the chosen reference potential, i.e. w.x; t/ D u.x; t/uref ,
it follows:
Z Z
w.x; t/ D div Jv d C Jv n d
˝H @˝H
and applying the first Green formula we obtain the integral representation:
Z Z
w.x; t/ D Jv r d D .Di rv/ r d : (5.12)
˝H ˝H
The previous derivation can be easily justified, in the framework of Sobolev spaces,
when DO is assumed regular in ˝. When DO exhibits, as in real cases, a jump
discontinuity on the interface ˙ then the Green function is solution of a boundary
154 5 Anisotropic Cardiac Sources
value problem with discontinuous coefficients and cannot be uniquely defined in the
sense of distributions.
Rigorous mathematical derivation of (5.7). To rigorously introduce Green
functions associated to the previous Neumann problems we consider a notion of
weak solution similar to the one proposed by Stampacchia for the Dirichlet problem
(see [343, 495]).
We denote by A D div Dr O and by @A D DrO n with D.˝/ and D. / the
set of measures with bounded variation and support on ˝ and on respectively.
Problem (5.8) is a non homogeneous Neumann problem of the type:
A ! D ˝ in ˝
(5.13)
@A ! D on
with data as measures given by (5.9) and (5.10). Extending the classical approach
of Stampacchia [495], we introduce the following notion of weak formulation
of (5.13):
Definition 5.2. Given Z˝ ; measures
Z of bounded variation on ˝ and
respectively, such that d˝ C d D 0, a function ! 2 L1 .˝/=R is a
˝
weak solution of problem (5.13) if
Z Z Z
!A ' D ' d˝ C ' d (5.14)
˝ ˝
We remark that the set of the test functions is not empty since, due to Corollary 7.8
in [343], the following regularity result holds:
Theorem 5.1. LetR ˝ be a domain with Lipschitz continuous boundary, f0 2
Lp .˝/ such that ˝ f0 dx D 0 and F 2 .Lp .˝//3 with p > 3. If z 2 W 1;2 .˝/=R
is the unique solution of
Z Z
a.z; '/ D f0 ' dx C r' F dx 8' 2 W 1;2 .˝/;
˝ ˝
then
max jzj C jjf0 jjLp C jjFjj.Lp /3 :
˝N
Z
.I x/ A ' d D '.x/ 'ref (5.15)
˝
R
where 'ref D '.x0 / or 'ref D j1 j ' d ; moreover, since A is self-adjoint,
.I x/ D .xI /. R
N with
For g 2 C 0 .˝/ 0 N
˝ g dx D 0, we consider ' 2 C .˝/ \ W
1;2
.˝/ solution
of
A ' D g in ˝
(5.16)
@A ' D 0 on :
From (5.15) we have the following integral representation for the solution of
N
problem (5.16) 8x 2 ˝:
Z
'.x/ 'ref D .I x/ g./ d : (5.17)
˝
Using Fubini’s theorem and formula (5.17) with 'ref D 0, it follows that:
Z Z Z Z Z
' d˝ C ' d D g.x/ .xI / d˝ C .xI / d dx
˝ ˝ ˝
and, comparing with (5.14), we obtain that the solution !.x/ of problem (5.13),
apart from an additive constant !ref , is given by:
Z Z
!.x/ D .xI / d˝ C .xI / d C !ref : (5.18)
˝
Going back to the variational problem (5.6) and defining the potential u.x; t/, it is
easy to verify that u.x; t/ is also a weak solution of problem (5.13) for the data
N
˝ ; defined by the functional on C 0 .˝/
Z Z Z
' 7! ' div Jv C ' Jv n C ' Jv n:
˝H ˙ H
0
From the representation formula (5.18), since 2 W 1;p .˝/ with p 0 < 3=2,
applying the Green formula, u.x; t/, apart from an additive constant, is given by
Z
u.x; t/ uref D Jv r d : (5.19)
˝H
156 5 Anisotropic Cardiac Sources
This formula coincides with (5.12), which represents the solution having zero
reference potential and holds 8x 2 ˝N since, from the regularity Theorem 5.1, we
N In conclusion the following Proposition holds:
have that u.x; t/ 2 C 0 .˝/.
Proposition 5.1. The solution of (5.6) with zero reference potential is given by:
Z
w.x; t/ D Jv r .I x/ d N 8t 2 Œ0; T ;
8x 2 ˝;
˝H
N
where, 8x 2 ˝, .I x/ is solution of problem (5.15).
In the following, we denote by
Finally, we observe that in both the differential and integral formulations of the
Bidomain representation of the multicellular cardiac tissue, the macroscopic cardiac
bioelectric sources generating the extracellular and extracardiac potentials consist of
the volume dipole density Jv .
During the excitation phase of the heartbeat, the transmembrane potential v exhibits
a steep transition layer propagating throughout the myocardium with a thickness
of about 1 mm. At every point of the cardiac domain, this upstroke phase lasts
about 1 ms. Therefore, the accurate simulation of the excitation process requires
discretization methods based on small space and time steps (of the order of 0.1 mm
and 0.01 ms). This fact constrained 3–D simulations, see e.g. [102, 104, 226, 544]
and the recent survey [545], and in order to reduce the computational costs for the
simulations of potential maps and electrocardiograms, approximate models of the
cardiac sources have been developed and used.
The evolving thin layer of cells undergoing the depolarization upstroke and
sweeping the whole ventricle is called the excitation wavefront and the excitation
wavefront surface is defined by means of the so called excitation or activation time
of the cell. During the excitation phase of the heart beat, each cell undergoes a
fast upstroke of the transmembrane potential, starting from the resting state vr ,
increasing monotonically during the depolarization upstroke and reaching finally
a depolarized state vd .
Therefore, at any point x 2 ˝H there exists a unique instant '.x/, called the
activation time, such that
vr C vd
v.x; '.x// D :
2
5.3 Approximate Representation of Cardiac Sources 157
Hd .t/ D fx 2 ˝H W '.x/ tg
Thus, under the assumption of equal anisotropy ratio, the cardiac sources are
represented as a dipole layer distributed on the epi-endocardial heart surface. For
this reason, this model is called heart surface source model, see [197–199,329,570].
The dipole distribution has dipolar moment proportional to the transmembrane
potential v.; t/, with oblique direction parallel to Di n and oriented outward the
cardiac tissue since nT Di n > 0.
Remark 5.1. If the transmembrane potential v is constant on the heart surface,
then (5.21) predicts a zero potential w0 at any point x outside the cardiac tissue.
Indeed, for v D c constant, due to (5.20) we have
158 5 Anisotropic Cardiac Sources
Z Z
v nT Di r .I x/ d D c nT Di r .I x/ d
@H @H
Z
Dc div.Di r .I x// d x D 0: (5.22)
˝H
It is reasonable to assume that the cardiac fibers are tangential to the cardiac surface,
i.e. al ; an ?n on @H . Then we have Di n D ti n, hence
Z
w0 .x; t/ D ti v.; t/nT r .I x/ d :
@H
This model describes the evolution of w0 during the entire heartbeat. Given the time
interval Œ0 T , the average of the extracardiac potential w0 can be written as
Z Z Z
1 T
1 T
w0 .x; t/dt D v.; t/dt ti nT r .I x/ d :
T 0 @H T 0
This formula can be used to obtain the following two interesting expressions for
time averages of the extracardiac potential.
(a) QRS – T average. During the QRS – T interval, the average transmembrane
potential, shifted with respect to the resting value vr ,
Z T
1
v./ D .v.; t/ vr /dt;
T 0
where T is the time of the T-wave ending, is the normalized area below the
action potential time course. Hence we can express the QRS – T average of the
extracardiac potential
Z T
1
w0 .x/ D w0 .x; t/dt
T 0
as a heart surface integral involving the normalized area v./ below the
transmembrane potential v,
Z
w0 .x/ D v./ti nT r .I x/ d :
@H
(b) QRS average. During the QRS interval, the action potential is in the depolar-
ization phase and it can be approximated with an activation profile v.; t/ D
A.t './/, where './ is the activation time at point . Then the QRS average
Z Td
1
w0 .x/ D w0 .x; t/dt;
Td 0
5.3 Approximate Representation of Cardiac Sources 159
where Td is the ending time of the QRS interval, has the representation
Z Z Td
1
w0 .x/ D A.t './/dt ti nT r .I x/ d : (5.23)
Td @H 0
for a non-negative measurable function g and a Lipschitz function ', and setting
r './
n ./ D , we have
jr './j
Z Z
w.x; t/ D A0 .t '.x//n ./T Di r .I x/d d
R ' 1 . /DS
Z Z
D A0 .t / .Di n /T r .I x/d d : (5.24)
R S
By approximating the action potential profile during the depolarization phase with
the jump discontinuity described by
A.t / D H . t/.vd vr / C vr ;
In this formula, the cardiac excitation sources are represented as a dipole distribution
supported on the excitation wavefront St , having oblique directions parallel to Di n
and with dipolar moment per unit surface area given by .vd vr /kDi n k2 .
Assuming an infinite extracardiac medium and the same constant conductivity
for the extracardiac and bulk tissue, i.e. D0 D Di C De D 0 I , then the Green
function becomes
1 1
.I x/ D ;
40 j xj
which coincides with the representation of the cardiac excitation source in terms of
normal dipoles uniformly distributed over the excitation surface St , also known as
uniform double layer, see [535, Ch. 6]. If we define the moment per unit surface
p D i .vd vr /;
Setting
@
d!.x; / D j xj1 d D nT r j xj1 d ;
@n
1
w.x; t/ D p!.x/: (5.28)
40
This formula shows that the magnitude of the potential at a point x is proportional
to the solid angle !.x/ under which the activation wavefront is seen from x. The
solid angle formula (5.27) and the associated potential representation (5.28) yield
immediately some important consequences:
(a) The potential field external to a closed surface St is zero;
(b) The potential field due to an open surface St is completely defined by its
boundary (rim);
(c) The potential is positive at a point x located “in front” of the surface St , i.e.
where nT . x/ < 0, since there the solid angle is negative; viceversa, the
potential is negative at a point x located “behind” the surface St , i.e. where
nT . x/ > 0, since there the solid angle is positive.
These properties can be applied to the simple case of a single depolarizing wave-
front spreading through the myocardium to obtain an approximate explanation of the
ECG inflections. Indeed, by (c) a depolarization front propagating toward a positive
electrode located at point x produces a positive signal, while a depolarization front
propagating away from a positive electrode located at point x produces a negative
signal. Moreover, if an excitation wavefront St1 at time t1 propagates to a larger
wavefront St2 at time t2 > t1 toward a positive electrode, then the potential increases,
162 5 Anisotropic Cardiac Sources
w.x; t1 / < w.x; t2 /, while if the propagation is away from the electrode then the
potential decreases, w.x; t1 / > w.x; t2 /. The previous prediction takes not into
account the role of the anisotropic structure of the myocardium, which is instead
incorporated in the following Sects. 5.4 and 5.5.
We present now a useful split form of the potential field, previously introduced in
[116–118, 121, 510] to investigate some features of the potential u.x; t/. It is used
also as a tool for a more accurate computation of u.x; t/. We focus first on the case,
frequently considered, of a cardiac tissue which is characterized by conductivity
tensors that are axially symmetric with respect to the local fiber direction; the
extension of the split form to tensors with no axial symmetry can be found at the
end of this paragraph.
If a D a.x/ is the unit vector defining locally the fiber direction and li;e , ti;e are
the conductivity coefficients along and across fiber in the intra- and extracellular
medium, then the assumption of axial symmetry implies that the conductivity
tensors Di ; De can be defined as
D D Di C De D t I C .l t /a ˝ a;
where t D ti C te and l D li C le . The conductivity coefficients may depend
on the position x according to the local state, normal or diseased, of the various part
of the myocardial tissue. It is easy to verify the following identity:
Di D ˛ D C ˇ a ˝ a (5.29)
l
with ˛ D ti =t , ˇ D li ti , l D li C le , t D ti C te . In fact, from the
t
definition of the bulk conductivity tensor D, we have that
t I D D .l t /a ˝ a:
5.4 Cardiac Source Splitting 163
ti
Di D .D .l t /a ˝ a/ C .li ti /a ˝ a
t
i i
D t D C li ti t .l t / a ˝ a
t t
ti i l
D D C l t
i
a ˝ a;
t t
Jv D Di rv D Ja C Jc (5.30)
where
We call Ja ; Jc axial and conormal current densities since they define two distribu-
tions of dipolar current sources in the heart with dipole axes parallel respectively to
the fiber direction a and to the vector Drv.
Setting
Z
1
w.x; t/ D u.x; t/ u.x; t/ d
j j
and using the split form (5.30) of Jv , relation (5.12) can be written as:
with
Z Z
wa .x; t/ D Ja r d and wc .x; t/ D Jc r d (5.33)
˝H ˝H
where
Z
uhs .x; t/ D ˛v .Dr / n d heart surface component (5.35)
˙
Z
1
ud .t/ D ˛v d drift component (5.36)
j j H
˛v.x; t/ x 2 ˝H
uj .x; t/ D jump component (5.37)
0 x 2 ˝0
Z
unh .x; t/ D v .Dr˛/ r d non-homogeneous component. (5.38)
˝H
Note that ud .t/ is independent of the observation point x hence it acts as a drift
component; moreover the component unh .x; t/ gives a contribution only for tissue
with non homogeneous anisotropy, i.e. when ˛ is not constant.
The above decomposition of the conormal component holds if the observation
point x does not lie on the non insulated boundary ˙ (e.g. the endocardium in
contact with intracavitary blood). If x 2 ˙ then the splitting of the conormal
component must account for the discontinuity of the two components uhs and uj
on ˙. More precisely applying classical limit formulae for the generalized double-
layer potentials on ˙ (see formula 15.3 of Theorem 15.II in [332]) we obtain the
following trace and jump relationships:
Z
1
u˙
hs D ˛vj ˛v .Dr / n d ; ŒŒ uhs ˙ D ˛ vj˙
2 ˙ ˙
5.4 Cardiac Source Splitting 165
where u˙ hs denotes the trace of the function on the positive and negative side of ˙
with respect to the oriented normal n. We remark that wc .x; t/ D uhs .x; t/ C ud .t/ C
uj .x; t/ is continuous across ˙ and has the following expression:
Z
1
x 2 ˙; wc .x; t/ D ˛vj˙ ˛v .Dr / n d C ud .t/ C uj .x; t/ C unh .x; t/:
2 ˙
Recent findings suggest that the cardiac tissue anisotropy is not symmetric across
fibers, but, due to the presence of myocardial sheets (or laminae) (see [352]), three
material principal axes can be identified, thus it presents an orthotropic structure.
Actually, the split form of the potential w.x; t/ can be extended to the case of no
axial symmetry for the conductivity tensors Di ; De . In fact we can assume that Di;e
have three distinct eigenvalues li;e ; ti;e and ni;e with the same principal axes given
by the set of orthonormal eigenvectors ak ; k D l; t; n with ak aj D ıjk and al D a,
i;e
i.e. tangent locally to the fiber. Setting A D .al ; at ; an /, Ddiag D diag.li;e ; ti;e ; ni;e /
and using the property that AAT D AT A D I , where AT and I denote respectively
the transpose of A and the identity matrix, we obtain:
Di;e D ADi;e T
diag A :
Di D ˛D C ˇal ˝ al C .I al ˝ al /; (5.40)
with ˛ D .ni ti /=.n t /; ˇ D li ˛l ; D ti ˛t . In fact, given ˛ 2 R,
we have
166 5 Anisotropic Cardiac Sources
Di ˛D D .ti ˛t /I
C .li ti ˛.l t //al ˝ al
C .ni ti ˛.n t //an ˝ an :
Choosing ˛ D .ni ti /=.n t /, the last term vanishes and we get the split
form (5.40). We remark that I al ˝ al is the orthogonal projection matrix on a
plane perpendicular to the fiber direction al .
If ni;e D i;e ti;e , then the media (i) and (e) tend to become axially symmetric
for i;e ! 1 with the limit values ni;e D ti;e . Setting i;e D 1 i;e , if i = e ! 1
l
as i;e ! 1 then it is easy to verify that ˛; ˇ; tend to ti =t , li ti ,0
t
respectively, which are the values obtained directly under the assumption of axial
symmetry.
In the orthotropic case, the split of the conductivity tensor Di leads to a
corresponding split of the impressed current Jv yielding:
Jv D Jc C Ja C Jo
From the relation between the field components and their sources we establish some
correlations between pattern features displayed by the full potential and some struc-
ture of the underlying cardiac sources. The potential components deduced from the
potential split depend on v.x; t/, which is a quantity of direct electrophysiological
significance since it characterizes the electric state of the myocardial cells. The
analysis of these components can shed light on the properties of the full potential
field. In the following, we focus only on the axisymmetric components of the
splitting.
Axial component (wa ): the potential wa can be related to the source Ja by
Z Z
wa .x; t/ D JTa r d D ˇ.aTl rv/aTl r d :
˝H ˝H
This expression shows that the potential wa is generated by volume sources repre-
sented by a current dipole density with dipole moment per unit volume proportional
to the vector .aTl rv/al . Therefore the dipole current sources are parallel to the local
fiber direction al and we call wa axial potential. Since the fiber architecture defines
the direction of the current source flow, we expect that structural properties like epi–
endocardial fiber rotation will be the major determinants affecting the axial potential
pattern.
5.5 Interpretation of the Field Components 167
When the excitation wavefront propagates through the heart muscle, it occupies
a thin layer of myocardium, 1 or 2 mm thick, that is passing from the resting to the
excited state. Within this layer, the distribution of axial sources has greater dipole
strength per unit volume in those regions in which the wavefront is spreading mainly
along fibers, i.e. where rv is nearly parallel to al .
Jump component (uj ): we recall that the resting value vr has been set to zero
so that in our model v measures the deviation of the transmembrane potential from
the resting state. We remark that the potential component uj .x; t/ D ˛v.x; t/
is zero in ˝0 , whereas in ˝H it is proportional to the spatial distribution of the
transmembrane potential. Since v, in our model, displays an upstroke in a thin layer
of tissue containing the wavefront and it is practically constant everywhere else, this
field component does not generate an appreciable current flow at distance from the
front. Moreover in the fully activated region the effect of uj on the full potential u
reduces only to a shift proportional to the jump of the transmembrane potential from
the resting to the plateau value. For this reason we call uj jump component.
Heart surface component (uhs ): the potential component uhs is given by
Z
uhs .x; t/ D .˛vDn/T r d : (5.41)
˙
We remark that uhs contributes to the full potential u only when the heart surface is
in contact with an extracardiac conducting medium (i.e. the non-insulated cardiac
interface ˙ is not empty) and depends on the trace on ˙ of the transmembrane
potential v. Since the sources lie on the non-insulated heart surface ˙, we refer to
this field component, as usual in the literature (see e.g. [197, 198, 483]), as heart
surface component.
The differential formulation shows that uhs exhibits a jump through ˙ equal to
˛v and the moment of the dipole layer source on ˙ is proportional to this jump; on
the unexcited parts of ˙, i.e. where v is practically zero, this jump vanishes.
In order to describe in a different and more meaningful way the relationship
between the field component uhs and its heart surface source, we make the following
simplifying assumptions: uniform anisotropy on @H , homogeneity on @˝0 (i.e. li;e ,
ti;e , 0 constant on ˙) and the fibers tangent to the epi- and endocardial surfaces
(i.e. aTl n D 0 on ˙). Hence in particular on ˙ we have Dn D t I , D0 n D 0 I
implying Dn D D0 n with D t =0 ; consequently uhs can be viewed as the
potential generated by a source term Jhs satisfying in the sense of distribution (see
[110, 305]) the equation
very close to vJ on ˙d and ˛ is constant, it follows that the normal dipole layer is
practically uniform on ˙d . Thus, the most remarkable property of uhs is that it can
be well approximated by means of the classical model, i.e. the potential generated
by a normal and uniform dipole layer on ˙d in the conducting medium described
O
by the tensor D.
Non-homogeneous component (unh ): the field component unh is given by
Z
unh .x; t/ D v.Dr˛/T r d :
˝H
Let us denote by Hd .t/ the depolarized region, by Hr .t/ the resting region and
by ˙d and ˙r their intersections with ˙ (one of the two can be void). The two
regions Hd .t/ and Hr .t/, are characterized by an activation time '.x/ whose value
is respectively smaller or greater than the fixed time instant t. Remembering that
5.6 The Limit Case: Oblique Dipole Layer Model 169
where the axial and the heart surface components, in the limit case, are given by the
following surface integrals:
Z Z
ua .x; t/ D ˇ vJ .aTl nSt /aTl r d and uhs .x; t/ D ˛ vJ nTSt Dr d :
St ˙d
(5.45)
In Eq. (5.44), the current sources are concentrated on the wavefront surface St and
are defined by a dipole layer with local dipole direction Di n. This model, also called
oblique dipole layer model, was proposed in [108] and subsequently investigated in
[110, 113].
We remark that the uhs component presents a jump equal to ˛vJ through ˙d
whereas it is continuous across ˙r . This means that different depolarized regions
Hd having the same ˙d on ˙ give rise to the same field component uhs . This is
reminiscent of a property of the potential field generated by a uniform normal dipole
layer. Indeed, the potential generated by these sources, when they are imbedded in
an infinite homogeneous isotropic medium (i.e. D .jxj1 jx0 j1 /=.40 /)
is given by uhs D ˛vJ !.x/, where ! is the solid angle subtended in x 2 Hr by the
surface ˙d . Thus for a given ˙ the field component uhs depends only on the rim of
the wave front intersection on ˙.
In summary, we have shown that when the anisotropic myocardial tissue is in
contact with conducting media, the potential presents a component uhs , having, in
the anisotropic context, the same features as the classical solid angle formula.
170 5 Anisotropic Cardiac Sources
In particular, this component has the peculiarity that it acts only when either
the epicardial or endocardial surface or both are not completely insulated and is
revealed, becoming measurable, only when activation reaches these non-insulated
boundaries; moreover the magnitude of its contribution is practically independent of
the wave front shape but depends only on the activated areas enclosed by the wave
front rim lying on the noninsulated boundaries. The jump component uj , related to
activated region (see Eq. (5.44)), introduces a shift of the potential. The sum of these
two components uj and uhs is the conormal component uc , usually referred to as the
heart surface source model, see [197, 198, 483].
From the split form it follows that for tissue having uniform anisotropy the full
potential is derived by superimposing the axial potential to the previously mentioned
potential components (due to the heart surface model and activating volume).
For tissue having uniform anisotropy, i.e. li;e , ti;e constant in ˝H , as assumed
in the numerical simulations, the full potential is derived by superimposing the
axial potential to the conormal component which in turn is the sum of the heart
surface component uhs and the potential jump uj across the wave front. The ability
of expressing the field as the sum of the axial and conormal components enables
us to underline the difference between the field produced by the classical conormal
model and the actual cardiac field.
It is worth mentioning here that the estimates of the intra- and extracellular
conductivities do not verify the condition of equal anisotropy ratio. Nevertheless
many computational studies in electrocardiology have often used this assumption.
Under this hypothesis ˇ D 0 hence ua drops out and consequently the cardiac fiber
arrangement plays no role in determining the structure of the current source density.
Cardiac anisotropy influences only the current conduction flow in the bulk medium,
described by the tensor D but frequently handled as an isotropic conducting medium
[108, 113, 198, 295, 483]. Therefore under the equal anisotropy ratio condition the
model reduces to u D uc C uhs C uj and coincides with the heart surface source
model used in [197,198,483]. As a consequence this model reduces to the conormal
one.
The more realistic physiological situation of a Bidomain model with unequal
anisotropy ratio, involves the heart surface model with the addition of the axial
potential component; the latter introduces a strong dependence of the cardiac
sources on the fiber anisotropy and architecture.
Figure 5.2 illustrates the connections between the epicardial potential patterns and
the intramural currents produced by the various source components described in the
previous section. Potential distributions are shown on the epicardial surface and
on three intramural sections: vertical, horizontal and oblique. On the intramural
5.7 Experimental and Simulation Results 171
Full Axial
Fig. 5.2 Source splitting of the full extracellular potential into axial, conormal and heart surface
components at 30 ms after endocardial pacing. In each panel, the potentials are shown on epicardial
(a), meridian (b), horizontal (c), oblique (d) sections of a truncated ellipsoid in contact with
intracavitary blood, with i D 9:5; b D 6 mS cm1
of the wavefront in contact with the intracavitary blood, indicates that the heart
surface component presents a jump through this basis. This jump characterizes a
source, lying on this interface, generating a current that spreads through the wall
thickness and reaches the insulated epicardial surface tangentially. If we disregard
the medium anisotropy, the heart surface features are similar to those exhibited by
the potential field generated in an isotropic medium by a classical source model, i.e.
a normal dipole layer lying on the same wavefront basis and having dipole moment
proportional to the jump component.
Figure 5.3 shows other simulation results on an oblique section in a test with
intracavitary blood and extracardiac conducting medium (outside the red contour
delimiting the cardiac tissue). The full (left panel) and conormal (right panel)
components are shown on an oblique section of the domain. The full component
shows that at 30 ms the epicardial breakthrough has not yet occurred, so in the
potential map on the epicardium we observe a typical pattern characterized by
the presence of two positive maxima separated by a negative area surrounding a
central minimum. Most of the epicardial current lines, issuing from the area of
maximum positivity, point toward the minimum site, either directly or after bending.
Consequently this minimum acts as a sink for the current flow, whereas the conormal
component predicts that any portion of the wavefront surface acts as a source for the
current flow.
These simulated results are confirmed by the experimental results of Fig. 5.4,
where the current lines are shown (in red) both inside the cardiac domain and in
the torso, on the indicated diagonal section. As in the simulated results of Fig. 5.3,
28
2.80 25
2.60 23
2.40 20
2.20 17
2.00 14
1.80 11
1.60 8
1.40 6
1.20 3
1.00 0
0.80 -7
0.60 -15
0.40 -22
0.20 -30
0.00 -37
-0.20 -45
-0.40 -52
-0.60 -60
-0.80 -67
-1.00 -75
Fig. 5.3 Source splitting; full (left) and conormal (right) components of extracellular potentials at
30 ms on truncated ellipsoid, with i D 9:5; b D 6 mS cm1
5.7 Experimental and Simulation Results 173
Fig. 5.4 Experimental results of extracellular potentials related to an isolated heart embedded in
a torso tank (Courtesy of prof. B. Taccardi)
the current lines issue from two positive maxima and point toward the sink (marked
in green in the right panel of Fig. 5.4). For more details, analysis and physiological
interpretation of the potential and current pathways in the extracardiac medium, see
[350, 513, 514, 516].
Chapter 6
The Inverse Problem of Electrocardiology
Since the end of the 1970s, automated instruments have been employed to record
potential Body Surface Maps (BSM), see e.g. [144, 504, 505, 511]. The spreading
of these techniques of potential recording is motivated by the fact that BSM
provide diagnostic information which could not be obtained by conventional
electrocardiograms in many heart diseases, such as Wolff-Parkinson-White and
myocardial ischemia. In fact some features of the potential surface field, such as
the number, location and time course of potential maxima and minima on the
chest surface, are correlated to the shape of the depolarization wavefronts and to
the spatial distribution of the repolarization processes in the heart, which are the
phenomena generating bioelectrical currents.
BSM exhibit a highly changeable surface pattern as characterized by equipo-
tential lines; moreover there is a high variability of the surface signal magnitude
over the entire heartbeat. The diffusion of BSM technique has raised the problem
of the interpretation of potential distribution on the thorax. The problem of the
best use of the information content of the large amount of data provided by
BSM may be attempted by solving the so-called Forward and Inverse problems of
Electrocardiology. Two approaches are possible:
• (P) in terms of potential alone;
• (S) in terms of cardiac sources.
In the first approach (P), the Forward problem consists of:
simulating the BSM from the epicardial extracellular potential distribution,
while the Inverse problem consists of:
estimating the epicardial extracellular potential distribution from the BSM.
We remark that this approach is justified by experimental studies on animals both
in vitro and in vivo, which show that epicardial maps contain a great amount of
information more directly readable in terms of the underlying cardiac events than
BSM; moreover in this approach it is possible to compare quantitatively computed
epicardial maps with the measured ones collected in experiments on animals.
The second approach (S) consists of modeling the cardiac sources by means of the
so-called equivalent cardiac generators, hence the Forward problem consists of:
simulating the BSM by means of an adequate Equivalent Generator Model,
while the Inverse problem consists of:
identifying an adequate Equivalent Generator of the heart from BSM.
Many groups have investigated methods for reconstructing the cardiac electric
activity with both the approaches, see e.g. [29, 64, 89, 106, 107, 111, 143, 210, 215,
222, 325, 353, 407, 440–442, 534, 550].
In relation to the feasibility of the inverse problem in terms of potential alone, the
following question arises: to what extent can the potential distribution on the heart
surface be computed from BSM? We discuss here the relevant features concerning
this question.
The human body ˝ can be considered as an isotropic linear resistive conducting
medium excluding the cardiac region ˝H . At any time instant of the heartbeat the
electric field can be considered quasi-static (see [394]) and the volume conductor
˝ is embedded into an insulating medium (the air), i.e. the normal derivative of
the potential is zero on D @˝. We set ˝0 D ˝ n ˝H , i.e. the extracardiac
medium, k.x/ the electrical conductivity and H D @˝H represents a fixed surface
surrounding the heart and lying in proximity of the heart surface; in the following,
H is referred to as epicardial surface. At any time t let V .x/ be the electric potential
distribution in ˝0 . If u.x/ D V .x/jH is known, then V .x/ in ˝0 is characterized by
the following mixed boundary problem:
8
ˆ
ˆ div k.x/rV .x/ D 0 in ˝0
<
V .x/ D u.x/ on H
ˆ @V .x/
:̂ D0 on
@n
and the inverse problem consists of estimating V .x/ on H . If the observed potential
z on ˙ were measured error-free, the Cauchy problem for the elliptic operator (6.1)
would define a unique solution V .x/, but in a highly unstable fashion, since, as it
6.1 Inverse Problem in Terms of Potential Alone 177
is well known, the Cauchy problem for elliptic operators is an ill-posed problem
in usual Sobolev spaces, see [236, 249, 374, 518]. Let v 2 L2 .H / define the state
1
y.xI v/ D y.v/ as the unique solution in H 2 .˝0 / of
8
ˆ
ˆ div k.x/ry.v/ D 0 in ˝0
<
y.v/ D v on H
ˆ @y.v/
:̂ D0 on ;
@n
then we introduce the following operator
Av D y.v/j˙
If z 2 A.L2 .H //, i.e. there exists a unique solution of the Cauchy problem,
then u D V .x/jH 2 L2 .H / satisfies Au D z, hence u is the unique minimizer
of J . But the problem is still unstable, i.e. the operator A admits an unbounded
inverse operator in the spaces H s , 8s 2 R. In order to solve the problem one must
approximate it with a family of stable problems. Many methods are available to
perform this stabilization. We first present Tikhonov regularization techniques, by
imposing smoothing constraints justified by the physical problem. We consider the
following Tikhonov zero-order, first-order and second-order regularizations:
then the inverse problem (6.2) is approximated by the following family of stable
problems dependent on :
We refer to [304] for the existence, the uniqueness of the minimizer u of (6.4) and
for the convergence of u to u.
Some Authors have also recently considered L1-regularizations considering the
total variation of v, i.e.
Z
J .v/ D J.v/ C jrvjd; v 2 U; > 0 “small”;
H
X
n X
n
u.x/ D ui w0i .x/ then y.xI u/ D ui yi .x/;
i D1 i D1
X
n
Ah uh D y h .xI uh /j˙ h D uk ykh .x/j˙ h ; (6.6)
kD1
P
where uh D nkD1 uk w0k .x/, fw0k gkD1;n is a basis of the finite element space used
on 0h and u D .u1 ; : : : ; un /T is the vector of the nodal values of uh on 0h . Setting
6.1 Inverse Problem in Terms of Potential Alone 179
z D f.Ah uh /.x˙ h
j /gj D1;m the vector of nodal values of Ah uh on ˙ then in terms of
z and u the relation (6.6) implies
z D T u;
where
is the so called transfer matrix between 0h and ˙ h . Moreover the matrix T
depends only on the geometric data ˝. The first step of the numerical procedure
consists in the computation of T ; T is obtained solving the n mixed boundary value
problems (6.5), which differ only for the Dirichlet data on 0 , by implementing a
suitable version of the frontal method which performs a triangular factorization only
once for all the n problems. The matrix T is a very ill-conditioned matrix reflecting
the ill-posedness of the inverse problem (6.2). We remark that the transfer matrix can
also be computed by means of methods based on boundary integral representations
of the potential. Since the accuracy of the transfer matrix is not a relevant factor
in the experimental inverse problem, all the following results will be related to the
transfer matrix computed by the three dimensional method outlined above, which is
convenient in order to reduce computational time and storage requirement. It is easy
to verify that the discretized form of the regularized problem (6.4) can be restated
in the following least square form:
Let ˝ be the body volume and ˝H the heart muscle. We suppose that ˝ and ˝H are
open connected sets of class C 2 and ˝H ˝. We recall that at a macroscopic level
the excitable cardiac tissue ˝H is conceived as the superposition of two continuous
media: the intra-(i) and extra-(e) cellular media. We denote by Di and De the intra-
and extracellular conductivity tensors of these media. The anisotropic conduction
properties characterized by the tensors Di;e are related to the fiber structure of the
heart muscle. We make the following assumptions regarding the anatomical fiber
architecture:
(i) The fibers are mathematically described by regular curves;
(ii) Through any point of ˝H passes a unique fiber;
(iii) On those parts of @˝H , which lie on the heart surface (epi- or endocardium),
the fibers are tangent to @˝H .
During the excitation phase of the myocardium, a layer of cardiac cells, changing
in time, undergoes the so-called depolarization process, i.e. a rapid change occurs
in the transmembrane electric potential v.x; t/, with an approximately monotone
variation from a resting value vr to a plateau value va . The following assumption is
usually made:
(iv) The depolarization of a cell is instantaneous and is the same for all cells (vr <
va constants).
For x 2 ˝H , let li;e .x/ be the (i)-(e) conductivity coefficients along the fiber
direction at point x. Assuming axial symmetry around the fiber direction, let ti;e .x/
be the (i)-(e) conductivity coefficients in any direction perpendicular to the fiber
direction at point x. If fal ; at ; an g denotes a local basis of R3 with al parallel to the
fiber direction, then the (i)-(e) conductivity tensors are defined by:
The axis at ; an are defined up to a rotation around al . Moreover Di;e do not depend
on the orientation of al parallel to the longitudinal fiber direction. If n is a unit vector
at point x, then:
H C .t/ D fx 2 ˝H W v.x; t/ D va g;
H .t/ D fx 2 ˝H W v.x; t/ D vr g;
C
with H [ H D ˝H , H C \ H D ; and S D @H C \ @H , the excitation
wavefront surface. We also assume that vr ; va are constant and S is an orientable
regular surface. Denoting by u.x/ the potential in ˝ and by J D 0 ru, Ji D Di rv
the current density in ˝ and in the intracellular medium, the current conservation
implies (see Sect. 3.3):
8
< div.J C Ji / D 0 in ˝H
div J D 0 in ˝ n ˝H (no sources)
: T
n JD0 on D @˝ (insulating boundary):
The boundary condition on D @˝ describes the fact that the thorax is surrounded
by air, which is an insulating medium. In terms of potential, we have:
8
ˆ
ˆ div Di rv in ˝H
< 0 u D
0 in ˝ n ˝H
ˆ @u
:̂ D0 on :
@n
Setting
va vr va vr i va vr i
M D Di D Adiag.mt ; mt ; ml /AT ; mt D t ; ml D l ;
0 0 0
(6.9)
where A D .al ; at ; an /, and using the results of [305] (for the Sobolev space H s see
e.g. [305]), the following result holds:
Theorem 6.1. Assuming that:
• The elements of M can be extended in the whole ˝ as functions of C 1 .˝/;
• S is a surface of class C 2 and if S is open there exists an open subset ˝ 0 of ˝
of class C 2 such that ˝ 0 ˝,
1
then there exists a solution u 2 H 2 .˝/, 8 > 0 of problem (6.8) defined up to an
additive constant.
Using the Green’s formula it is also possible to prove the following theorem, see
[110]:
Theorem 6.2. Under the assumptions of Theorem 6.1, the solution of (6.8) in
1
H 2 .˝/, 8 > 0, satisfies the following boundary value problem:
8
ˆ
ˆ u D 0 in ˝ n S
ˆ
ˆ
ˆ
ˆ @u
< D0 on .in the sense of H 1 . //
@n
ˆ ŒuS D ujS C ujS D n M n
ˆ
T
.in the sense of H .S //
ˆ
ˆ
ˆ @u @u @u
:̂ D D divS ˇ .in the sense of H 1 .S //;
@n S @n jS C @n jS
(6.10)
we derive that, under the assumption (iv), the mathematical model of the excitation
wavefront S consists of:
an oblique dipole (double) layer on S having the same orientation of the vector
M n and moment density on S given by jjM njj.
Hence M can be viewed as a dipolar moment tensor. Introducing the unit vectors
al , at such that al is parallel to the longitudinal fiber direction, at is perpendicular
to al and coplanar with n, al and orienting at , al toward the resting tissue so that
at n 0 and al n 0, it is easy to verify:
Z Z Z
@s @s
nT M rsdS D mt .n at / dS C ml .n al / dS:
S S @at S @al
The source can be viewed as the superposition of a transverse dipole layer with
density mt .n at / and an axial dipole layer with density ml .n al /, where, in this
context, transverse and axial means respectively perpendicular and parallel to the
local fiber direction.
Moreover the following decomposition holds:
Z Z Z
@s @s
n M rsdS D
T
mt dS C .ml mt /.n al / dS;
S S @n S @al
which shows that the oblique dipole layer potential can be thought of as the
superposition of a normal dipole layer with density mt and of an axial dipole layer
with density .ml mt /.n al /. If mt D ml constant, we recover the classical model,
i.e. the source is represented by a normal dipole layer. We remark that mt D ml
implies ti D li , i.e. the intracellular medium is isotropic; however experimental
measurements have shown that ti ¤ li and the experimental data of [30, 134] did
not agree with the prediction of the normal dipole layer model, thus questioning
the general validity of the classical model. The model presented here generalizes
the classical model and also the axial model introduced in [134] for explaining the
discrepancy with the prediction of the classical model in dog heart experiments. Our
source model is an anisotropic oblique dipole model where the anisotropy derives
from the properties of the tensor M .
184 6 The Inverse Problem of Electrocardiology
ˇ nb D 0 on @S: (6.13)
That is the wavefront surface S and the heart surface @˝H are, respectively, tangent
or perpendicular or else intersect at an arbitrary angle in the point of @S in which
al is tangent to @S .
6.4 Numerical Approximation of the Integral Representation of the Potential 185
For (ii) we do not know error estimate results. However in the following we will
present this second strategy, that is the most used in practice due to the high
computational costs required by the approximation of a double surface integral in
the (i) procedure.
The surfaces , S are approximated by means of two polyhedral surfaces h ,
h
S with triangular elements. The potential v is approximated by a piecewise linear
continuous function vh .x/ on h , i.e.
X
N
vh .x/ D vj pj .x/; with pj .xk / D ıjk ; vk D vh .xk /;
j D1
where xk denote the nodes on h . Applying the collocation method at the nodes of
h , the integral equation is approximated by:
Z Z
@s.xk ; /
!.xk /vk C vh ./ dh D nTh M r s.xk ; /dSh for k D 1; : : : ; N;
h @nh; Sh
(6.15)
186 6 The Inverse Problem of Electrocardiology
where
Z
˛.xk / @s.x; /
!.xk / D with ˛.x/ D dh ;
4 h @nh;
i.e. ˛.x/ is the solid angle under which the surface h is seen from the point x, nh
is the normal to the triangular elements belonging to h or S h . Setting
Z
@s.xk ; /
K D akj D !.xk /ıkj C pj ./ dh
h @nh
Z
b D bk D nTh M r s.xk ; /dSh ; v D .v1 ; : : : ; vN /T ;
Sh
Kv D b: (6.16)
With !.x/ defined as above, matrix K admits the zero eigenvalue associated to the
right eigenvector
e D .1; : : : ; 1/T :
We remark that the integral operator associated to the first side of Eq. (6.12) admits
zero as a simple eigenvalue; in the following we assume that zero is a simple
eigenvalue also for the approximate operator K. Hence the solution of the linear
system is defined up to an additive constant; moreover system (6.16) admits a
solution if ` T b D 0, where ` T K D 0T , i.e. ` is the left eigenvector of K, associated
to the zero eigenvalue. While in the continuous
R case the corresponding compatibility
condition is exactly satisfied by the term S nT M rsdS, on the other hand due to
approximation errors, the term b does not satisfy exactly the condition ` T b D 0.
For solving the singular linear system by direct methods, the following deflation
method is used. Let q be a vector such that qT b D 1. It is easy to verify that:
if ` T b D 0 then the solution w of the system Bw D .K C eqT /w D b,
where B is non-singular, satisfies the equations Kw D b; qT w D 0.
For ` T b ' 0, we solve the linear system
.K C eqT /v D b
In this section, we present the inverse problem in terms of wavefront, using the
oblique dipole layer model as a representation of the depolarization wavefront.
The problem can be roughly formulated a s follows: is it possible to determine
the wavefront S from the knowledge of the potential on the thorax ?
We recall that the knowledge of:
• The myocardial fibers direction (al );
• The jump of the intracellular action potential (ua ur );
• The coefficients of the longitudinal and transverse intracellular conductivity
(li ; ti );
• The extracellular and extracardiac conductivity (0 )
characterizes the matrix M given in (6.9).
In the following, we assume that li ; ti are positive constants and li ¤ ti , i.e.
we have homogeneous intracellular anisotropy, or equivalently we assume:
S ! US D uj ;
S ! US D v;
where v is the solution of the integral equation (6.12) of Theorem 6.3. That is, given
S we consider the potential US on the insulating boundary generated by the
188 6 The Inverse Problem of Electrocardiology
oblique dipole layer on S . We remark that in the special case in which al is always
perpendicular or tangent to a given surface S , i.e. al D nS or al nS D 0 then the
oblique dipole layer on S is a normal double layer; if S1 and S2 are two surfaces of
W satisfying the preceding constraint, then it is well known that US1 D US 2 implies
@S1 D @S2 , i.e. the potential characterizes only the boundary @S of S . Assuming
that the wavefront is generally oblique w.r.t the fiber direction, i.e. al is not always
perpendicular or tangent to the wavefront, the following uniqueness result holds (see
[110]):
Theorem 6.4. If S1 ; S2 2 W and on at least one of them, e.g. S1 , al is generally
oblique to S1 , i.e. al nS ¤ 0 or al nS ¤ 1, but there exists at least one point where
al nS D 1 or al nS D 0, then
US1 D US2 ) S 1 D S 2 :
Under the same hypotheses of Theorem 6.4, the uniqueness result holds also for
closed wavefronts, i.e. S1 and S2 are C 2 closed surfaces one inside the other. These
first results and others suggest the conjecture that the uniqueness result should hold
under the same hypotheses of Theorem 6.4 in a more general admissible wavefront
class W and for ml ¤ mt variable in ˝H . Finally we outline the following open
problems:
• Extension of the results to an extracellular anisotropic medium and to a body
volume with piecewise constant conductivity;
• Numerical approaches for solving the inverse problem.
We now describe the inverse problem presented in [142,143,241,534,536] for the
reconstruction of the activation time on the endo- and epicardial surfaces, assuming
homogeneous isotropic properties of the cardiac and extracardiac media. As we have
seen in Chap. 5, in the general anisotropic case, the spatio-temporal transmembrane
potential distribution over the whole heart muscle must be taken into account as
primary cardiac electric source. Indeed, the integral representation of the electric
potential u at a point x of the body surface is
Z
u.x; t/ D Di rv.y; t/ r .y; x/d y;
˝H
where Di is the intracellular conductivity tensor and the Bidomain Green func-
tion, see Eq. (5.7). However, if we assume isotropy for the intra- and extracellular
domains and accept the model error involved, only the transmembrane potential
distribution over the heart’s surface (epi- and endocardium) takes effect as primary
source, because the volume integral is converted to a surface integral by means
of Gauss divergence theorem. In fact, assuming all the media isotropic and the
extracardiac medium unbounded, we obtain
Z
i 1
u.x; t/ D rv.y; t/ r d y;
40 ˝H jy xj
6.5 Inverse Problem in Terms of Wavefront 189
This simplification is necessary because the fiber orientation of the individual patient
is unknown and cannot be assessed by current imaging methods.
The onset of the transmembrane potential v is described by the activation time
. The non-linear relationship between the activation time and the transmembrane
potential v makes the problem non-linear. This non-linear problem can be reduced
to a linear problem in the unknown activation time by approximating the time
evolution of the transmembrane potential by a Heaviside step function H , i.e.
and, integrating over the QRS interval .0; T /, we obtain finally the following linear
problem in
Z Z
T
i .va vr / 1
u.x; t/dt D .y/r n.y/d y;
0 40 @˝H jy xj
which relates the measured body surface potential u with the unknown activation
time on the heart surface. A minimum-norm least-square solution to this linear
problem, however, contains unphysiological characteristics and additional regu-
larization is inevitable. The additional regularization yields a non-linear inverse
problem. For a comparison between the potential- and the activation-based inverse
problems see [89].
An alternative strategy followed e.g. by [335] consists of representing the
transmembrane potential v as an analytical function with a sigmoidal shape,
va 2 t .y/
v.y; t/ D 1 C arctan C vr ;
2 w
The same type of approach can be extended to the entire heartbeat, representing
the transmembrane potential waveform by an analytical function involving some
parameters as vr and va and markers of local depolarization and repolarization time,
see [534, 537].
The inverse method described here attempt to reconstruct the electrical activity
on the epi- and/or endocardial surfaces of the heart. In recent years, there have also
been several attempts to reconstruct cardiac electrical activity inside the myocardial
wall, see e.g. [217, 222, 325, 326, 468, 478, 549–551]; see also [353, 354] for inverse
problems to locate transmural ischemia. This problem is even more difficult than
that of reconstructing surface electrical activity, as it is less constrained and hence
more likely to loose uniqueness of solutions.
Chapter 7
Numerical Methods for the Bidomain
and Reduced Models
In this chapter, we review the main numerical techniques used in the literature for the
space and time discretizations of the Monodomain and Bidomain cardiac models.
Then we present a Bidomain discretization based on the finite element method
in space and finite difference methods in time. The latter can be fully implicit
or semi-implicit, and can employ decoupling techniques and operator splitting
methods between the ordinary and partial differential equation components of the
cardiac models. The chapter is concluded by a review of numerical methods for the
eikonal-diffusion equation.
Many different approaches have been developed for the space discretization of
the Bidomain and Monodomain equations. Finite difference methods have been
studied in [66, 223, 368, 399, 413, 449, 529, 568, 569, 591]. Finite element methods
have been widely used, see e.g. [104, 124, 177, 302, 303, 318, 388, 455, 543, 544].
Finite volume methods, which have the advantage of conserving local flux, have
been developed in [43, 139, 218, 220, 252, 375, 528]. Some researchers have also
investigated spectral element methods [54] and high-order finite elements [13, 14].
Non-conforming finite element discretizations have been studied in [376, 377]. In
order to reduce the computational load of the Bidomain system and of general
parabolic reaction-diffusion systems, also adaptive remeshing techniques have been
developed, see e.g. [35, 36, 90, 91, 124, 147, 290, 336, 524, 556, 577, 578]; these
techniques have been proven successful for problems of moderate size and are
currently under investigation. Numerical methods and simulations for different
eikonal approaches can be found in [268, 271, 272, 365] and in [115, 116, 121, 522].
the usual L2 inner product and elliptic bilinear forms. The variational formulation
of the Monodomain model reads as follows. Given v0 ; w0 2 L2 .˝H /; Iapp 2
L2 .˝H .0; T //, find v 2 W 1;1 .0; T I V / and w 2 W 1;1 .0; T I L2 .˝H /M / such
that 8t 2 .0; T /
8
ˆ @
ˆ
ˆ Cm .v.t/; '/ C a.v.t/; '/ C .Iion .v; w/; '/ D .Iapp ; '/ 8' 2 V
< @t
ˆ
ˆ @
:̂ .w.t/; / D .R.v.t/; w.t//; / 8 2 L2 .˝H /M ;
@t
(7.1)
8
ˆ
ˆ
@
Cm .v.t/; uO i / C ai .ui .t/; uO i / C .Iion .v; w/; uO i / D .Iapp
i
; uO i / 8Oui 2 V
ˆ
ˆ
ˆ
ˆ @t
ˆ
ˆ
ˆ
<
@
Cm .v.t/; uO e / C ae .ue .t/; uO e / .Iion .v; w/; uO e / D .Iappe
; uO e / 8Oue 2 V
ˆ
ˆ @t
ˆ
ˆ
ˆ
ˆ
ˆ
ˆ @
:̂ .w.t/; w/
O D .R.v.t/; w.t//; w/; O v.x; t/ D ui .x; t/ ue .x; t/ 8w O 2 L2 .˝H /M ;
@t
(7.2)
i;e
with the proper initial conditions on v; w and compatibility condition on Iapp .
Finite element discretizations. Let T be a uniform hexahedral triangulation of
h
˝H and V h the associated finite element space, see [414]. Choosing a finite element
basis f i g for V h and appropriate quadrature rule, we denote by
Z Z
M D fmrs D r s dxg; Ai;e D fai;e
rs D .r r/
T
Di;e r s dxg;
˝H ˝H
7.1 Space Discretization of Monodomain and Bidomain Models 193
the symmetric mass matrix and stiffness matrices, respectively, and by Ihion ; Ihapp
the finite element interpolants of Iion and Iapp , respectively. In the following, we
will denote by the same letters finite element functions and the vectors of their
nodal values. In the Monodomain model, the finite element approximation vh of the
transmembrane potential is the solution of
@vh
Cm M C Avh C MIhion .vh ; wh / D MIhapp ; (7.3)
@t
while in the Bidomain model, the finite element approximations ui;h ; ue;h of the intra
ed extracellular potentials are the solutions of the system
8
ˆ @vh
ˆ
< Cm M @t C Ai ui;h C MIion .vh ; wh / D MIapp
h i;h
ˆ
(7.4)
ˆ
ˆ @v
:̂ Cm M h C Ae ue;h MIh .vh ; wh / D MIe;h ;
ion app
@t
where vh D ui;h ue;h . In both cases, these equations are coupled with the
semidiscrete approximations of the gating and concentration system
@wh
D R.vh ; wh /:
@t
The Bidomain system can be written in compact form as
!
@ ui;h ui;h MIhion .vh ; wh / MIi;h
Cm M CA C D app
; (7.5)
@t ue;h ue;h MIhion .vh ; wh / MIe;h
app
where
M M Ai 0
M D ; A D :
M M 0 Ae
The advantage of implicit methods is that they do not require stability constraints
on the choice of the time step, but they can be quite expensive, because at each time
step one has to solve a non-linear system. Fully implicit methods in time have been
considered in [233, 338–340, 342, 574].
7.2 Time Discretization of Monodomain and Bidomain Models 195
Considering for simplicity the Backward Euler (BE) method with time step size
D tnC1 tn , the Bidomain system requires at each time step the solution of the
non-linear system
F .u/ D 0;
defined as
2 ! ! 3
unC1
i MIion .vnC1 ; wnC1 ; cnC1 /
6. M C A / C 7
6 unC1 MIion .vnC1 ; wnC1 ; cnC1 / 7
6 e
! ! 7
6 7
6 uni MIiapp 7
F .u/ D 6 M 7;
6 7
6 une MIeapp 7
6 7
4 5
wnC1 R.vn ; wnC1 / D wn ; cnC1 D cn C S.vn ; wnC1 ; cn /
where
Cm M M Ai 0
D ; M D ; A D :
M M 0 Ae
A first attempt to reduce the computational cost of fully implicit methods is based
on decoupling the gating variables w from the potentials ui ; ue (hence v). This
strategy could be motivated by the nonzero pattern of the Jacobian of a fully-implicit
discretization when this shows a weak coupling between the gating and the potential
variables (for example using the LR1 ionic model the Jacobian is dominated by
its diagonal entries). Another motivation is that these two groups of variables are
associated to two very different submodels (the ionic model and the tissue model,
respectively) that describe different scales and are expressed mathematically by
different systems (an ODE and a PDE system, respectively).
We describe this decoupling technique here by employing for simplicity a
backward Euler scheme for the parabolic reaction-diffusion system.
Decoupled backward Euler (BE) method. Each time step of a decoupled
backward Euler (BE) time discretization of the Bidomain system (7.5) consists
of the following two steps. Analogous decoupled techniques could be applied to
high-order time discretizations, such as, e.g., third-order Rosenbrock methods (see
[338]). For simplicity, we will consider a fixed time step size .
Step 1: Updating gating and ionic concentration variables. Given vn , wn ,
cn , computed at the previous time step tn , first update the gating and ion
concentration variables at time tnC1 by solving
196 7 Numerical Methods for the Bidomain and Reduced Models
F.unC1
i ; unC1
e / D G; (7.7)
where
and
" #
uni MIi;nC1
GD M C app
:
une MIe;nC1
app
where
M M Ai 0 Cm
Abid D M C A D C ; D ; (7.9)
M M 0 Ae
and
!
MŒIion .vn ; wnC1 ; cnC1 / C Ii;nC1
app
F D : (7.10)
MŒ Iion .vn ; wnC1 ; cnC1 / C Ie;nC1
app
As in the continuous model, vn is uniquely determined, while uni and une are
determined only up to the same additive time-dependent constant chosen by
imposing the condition 1T Mune D 0. Hence, at each time step we have to solve
the large linear system (7.8), that, as shown in [104], is very ill-conditioned and
increases considerably the computational costs of the simulations.
Following the techniques from [166], we now investigate the stability of this
semi-implicit scheme. For sake of simplicity, we consider a membrane model with
only one gating variable and without stimulus currents Iiapp ; Ieapp . Since the reaction
terms are taken explicitly, we suppose that Iion and F satisfy the following Lipschitz
condition:
we get
which, thanks to the uniform ellipticity of the conductivity tensors Di;e and to the
Lipschitz condition (7.11), becomes
Denote now by M the index of the final time step (thus T D M ) and sum the
previous inequality for n going from 0 to m 1 with 1 m M , to obtain
X
m1 X
m1 X
m1
jjvm jj20 C jjvnC1 vn jj20 C 2 m .juinC1 j21 C juenC1 j21 / C jjwm jj20 C jjwnC1 wn jj20
nD0 nD0 nD0
X
m1
jjv0 jj20 C jjw0 jj20 C C1 .jjvm jj20 C jjwm jj20 / C C2 .jjvn jj20 C jjwn jj20 /;
nD0
X
m1
.1 C1 /.jjvm jj20 C jjwm jj20 / C 2 m .juinC1 j21 C junC1
e j21 /
nD0
X
m1
jjv0 jj20 C jjw0 jj20 C C2 .jjvn jj20 C jjwn jj20 /:
nD0
Lemma 7.1 (Discrete Gronwall). Let fkn g, fpn g be two sequences of non-negative
real numbers, ˚ m a discrete real-valued function and g0 a non-negative real number
such that ˚ 0 g0 . Suppose also that 8m 1,
X
m1 X
m1
˚ m g0 C pn C kn ˚ n :
nD0 nD0
X
m1
1 T C2
.junC1
i j21 C junC1
e j21 / .jjv0 jj20 C jjw0 jj20 / C max.jjvn jj20 C jjwn jj20 /:
nD0
2m 2m n
We shall now consider the stability of a semi-implicit second order BDF method
(SBDF). Taking unC1i ; uenC1 ; wnC1 as test functions, the associated variational
equations read as follows
Z
3vnC1 4vn C vn1 nC1
ui Cai .unC1
i ; unC1
i /C
˝H 2 Z
.2Iion .vn ; wn / Iion .vn1 ; wn1 //unC1
i D0
Z ˝H
3vnC1 4vn C vn1 nC1
ue Cae .unC1 e ; unC1
e /
˝H 2 Z
.2Iion .vn ; wn / Iion .vn1 ; wn1 //unC1
e D0
Z ˝Z
H
3wnC1 4wn C wn1 nC1
w .2F .vn ; wn / F .vn1 ; wn1 //wnC1 D 0:
˝H 2 ˝H
Adding these three equations, multiplying by 4 and applying to the time discretiza-
tion terms the identity
Thanks to the uniform ellipticity of the conductivity tensors Mi;e and to the Lipschitz
condition (7.11), we get
Denoting now by M the index of the final time step (thus T D M ) and summing
the previous inequality for n going from 0 to m 1 with 1 m M , we will see
the values v1 and w1 appear. They are used during the first time step of the SBDF
scheme and we take them v1 D v0 and w1 D w0 . Then we get
X
m1
jjvm jj20 C jj2vm vm1 jj20 C 4 m .juinC1 j21 C juenC1 j21 / C jjwm jj20 C jj2wm wm1 jj20
nD0
X
m1
2.jjv0 jj20 C jjw0 jj20 / C C1 .jjvm jj20 C jjwm jj20 / C C2 .jjvn jj20 C jjwn jj20 /;
nD0
X
m1
.1 C1 /.jjvm jj20 C jjwm jj20 / C 4 m .juinC1 j21 C junC1
e j21 /
nD0
X
m1
2.jjv0 jj20 C jjw0 jj20 / C C2 .jjvn jj20 C jjwn jj20 /:
nD0
X
m1
1 T C2
.junC1
i j21 C junC1
e j21 / .jjv0 jj20 C jjw0 jj20 / C max.jjvn jj20 C jjwn jj20 /:
nD0
4m 4m n
A further decoupling approach consists of splitting the parabolic PDE from the
elliptic PDE in the parabolic-elliptic formulation of the Bidomain model (7.6). We
will now present some stability estimates originally developed by Fernandez and
Zemzemi [172] for this decoupled method. At each time step, the numerical scheme
consists of the following steps.
7.2 Time Discretization of Monodomain and Bidomain Models 201
Step 2: given vn , update the extracellular potential une by solving the elliptic PDE
Step 3: given une , wnC1 , cnC1 update the transmembrane potential vnC1 by solving
the parabolic PDE
Cm Cm
M C Ai vnC1 D Mvn Ai une Iion .vn ; wnC1 ; cnC1 / C Ii;nC1
app :
we get
By summing and subtracting the term 2 ai .vnC1 ; une / at the left hand side, we get
Let us now focus on the term 2 ai .vn vnC1 ; une / and perform the following
computations
Moving the term 2 ai .vn vnC1 ; vnC1 C une / to the right hand side and using the
Lipschitz condition (7.11), we finally get
Denoting now by M the index of the final time step (thus T D M ), summing
the previous inequality for n going from 0 to m 1 with 1 m M and using
the uniform ellipticity of the conductivity tensors Mi;e , we obtain
X
m1 X
m1
jjvm jj20 C jjvnC1 vn jj20 C .mi jvnC1 C une j21 C 2me june j21 / C
nD0 nD0
X
m1
mi jvm j21 C jjwm jj20 C jjwnC1 wn jj20
nD0
X
m1
jjv0 jj20 C Ci jv0 j21 C jjw0 jj20 C C1 .jjvm jj20 C jjwm jj20 / C C2 .jjvn jj20 C jjwn jj20 /;
nD0
X
m1
.1 C1 /.jjvm jj20 C jjwm jj20 / C .mi jvnC1 C une j21 C 2me june j21 / C mi jvm j21
nD0
X
m1
jjv0 jj20 C Ci jv0 j21 C jjw0 jj20 C C2 .jjvn jj20 C jjwn jj20 /:
nD0
7.2 Time Discretization of Monodomain and Bidomain Models 203
X
m1
.mi jvnC1 C une j21 C 2me june j21 / C mi max jvn j21
nD1;:::;M
nD0
jjv0 jj20 C Ci jv0 j21 C jjw0 jj20 C T C2 max.jjvn jj20 C jjwn jj20 /:
n
In this section, we will present two operator splitting techniques for the Bidomain
model, based on splitting the ODEs of the membrane model from the PDEs. The
first method, Godunov splitting, is first order accurate in time, while the second
method, Strang splitting, is second order accurate in time. Both methods were first
introduced for the Bidomain model by Sundnes et al. [501]. The Godunov splitting
for the Monodomain model was developed by Qu and Garfinkel [410].
Godunov splitting.
Step 1: given the initial conditions vn , wn , cn computed at the previous time step,
find vQ nC1 , wnC1 , cnC1 by solving the ODEs system
8
ˆ dv
ˆ
ˆ Cm D Iion .v; w; c/ tn t tn C
ˆ
< dt
dw
D R.v; w/ tn t tn C
ˆ
ˆ
ˆ ddtc
:̂ D S.v; w; c/ tn t tn C :
dt
Step 2: given the initial condition vQ nC1 computed at Step 1, find unC1
i and uenC1
by solving the PDEs system
8
ˆ
< Cm
dv
C Ai ui D Iiapp tn t tn C
dt
:̂ Cm d v C Ae ue D Ieapp tn t tn C :
dt
Strang splitting.
Step 1: given the initial conditions vn , wn , cn computed at the previous time step,
find vnC1=2 , wnC1=2 , cnC1=2 by solving the ODEs system
204 7 Numerical Methods for the Bidomain and Reduced Models
8
ˆ dv
ˆ
ˆ Cm D Iion .v; w; c/ tn t tn C =2
ˆ
< dt
dw
D R.v; w/ tn t tn C =2
ˆ
ˆ
ˆ ddtc
:̂ D S.v; w; c/ tn t tn C =2:
dt
Step 2: given the initial condition vnC1=2 computed at Step 1, find uQ inC1 and uQ enC1
by solving the PDEs system
8
ˆ
< Cm
dv
C Ai ui D Iiapp tn t tn C
dt
:̂ Cm d v C Ae ue D Ieapp tn t tn C :
dt
Step 3: given the initial conditions vQ nC1 D uQ inC1 uQ enC1 , wnC1=2 , cnC1=2
computed at Step 1 and 2, find vnC1 , wnC1 , cnC1 by solving the ODEs system
8
ˆ dv
ˆ
ˆ Cm D Iion .v; w; c/ tn C =2 t tn C
ˆ
< dt
dw
D R.v; w/ tn C =2 t tn C
ˆ
ˆ
ˆ ddtc
:̂ D S.v; w; c/ tn C =2 t tn C :
dt
In order to achieve the second order convergence for the Strang splitting scheme,
the sub-problems in each step must be solved to at least second-order accuracy.
Numerical methods for finding the evolving surface S.t/ based on a direct dis-
cretization of (4.23) encounters many difficulties, e.g. front–tracking techniques,
when high curvature and topological changes of the wavefronts occur. One way
to overcome the singularities due to collisions, merging and extinction is the level
set approach of (4.23) (see e.g. [365, 471]) which represents S.t/ as the zero level
surface of a function w.x; t/ formally solving
@w
D ˚.x; rw/.c C " div ˚ .x; rw//: (7.12)
@t
As observed before, during the excitation sequence in a fully recovered tissue
the wavefront surface S" .t/ admits a cartesian representation. Therefore " ./ D
1=jr j hence, dropping O."2 /, the eikonal-curvature equation (4.22) reduces to
In both equations, the term of order " is related to the influence of the wavefront
curvature on the propagation in an anisotropic medium.
In a fully recovered tissue the propagation front admits a cartesian representation
and, from (4.17), (4.18), it is easy to verify that
p Q.x; p/p
˚.x; p/ D pT Q.x; p/p and ˚ .x; p/ D ;
˚.x; p/
where Sa is the boundary of the initial activated region and ta .x/ the corresponding
activation instants. The solution .x/ of this non-linear problem can be viewed as
the steady–state solution of the following parabolic problem associated with (7.15)
8 p
ˆ
ˆ
@w
ˆ
ˆ " divQ.x; rw/rw C c rwT Q.x; rw/rw D 1 in ˝H
< @t
nT Q rw D 0 on ; (7.16)
ˆ
ˆ
ˆ
:̂ w.x; t/ D w .x/ on Sa w.x; 0/ D w0 :
a
with wn the approximate solution at the time tn and n D tnC1 tn . The space
discretization was carried out by the usual Galerkin finite element method. We
remark that the solution of the discrete problem can exhibit spurious secondary
fronts originating at the domain boundary, see [102]. This is due to the fact that
206 7 Numerical Methods for the Bidomain and Reduced Models
in Eq. (7.15) the transport term is dominant with respect to the non-linear diffusion
term. To overcome this problem the term ˚.x; rw/ required special treatment. A
way to avoid these numerical artifacts is to use the upwind scheme proposed by
Osher-Sethian for propagating fronts with curvature dependent speed. This upwind
method proved to be quite efficient and allowed us to solve our equation in every
case and using a mesh-size h of the order of 1 mm (see [102, 115] and also [522]).
Interested readers can find many results of numerical simulations with the eikonal
approach in [102, 114, 115, 117, 121] and [268, 271, 272, 369]. See [186] for recent
developments of numerical techniques to solve the eikonal equations.
Finally, we remark that the non-linear parabolic equation shares the same
non-linear diffusion term of the Relaxed Monodomain model. A similar implicit-
explicit time discretization could be applied for the numerical solution of the
Relaxed Monodomain model where the reaction term, modeling the ionic current
membrane, is treated explicitly instead of the Hamiltonian term. Therefore the semi-
discretization in time of the Relaxed Monodomain model reads:
8 nC1
ˆ
ˆ w n R.vn ; wnC1 / D wn
< nC1
c D c n C n S.vn ; wnC1 ; c n /
ˆ v nC1
vn
:̂ cm div.Q.x; rvn /rvnC1 / C iion .vn ; wnC1 ; c nC1 / D Iapp
nC1
:
n
Chapter 8
Parallel Solvers for the Bidomain System
We have seen in Chap. 7 that most numerical approaches to solve the Bidomain
system lead to the solution of linear systems of equations. Linear solvers can be
classified as direct and iterative solvers. Direct solvers compute the solution of the
linear system in a finite number of steps, for instance by Gaussian elimination or
computing a factorization of the matrix, and in exact arithmetic would provide the
exact solution of the system. LU factorization is one of the most common direct
solvers, together with the Cholesky factorization for symmetric positive definite
matrices. On the other hand, in finite precision arithmetic, rounding errors and
the matrix ill-conditioning may spoil the approximate solution computed by direct
methods. Moreover, these direct methods have a high computational cost (cubic in
the number of unknowns of the linear system) and can suffer from fill-in, that is the
sparsity pattern of the original sparse matrix can be filled during the factorization
process, thus requiring large amounts of memory. One advantage of direct methods
in the context of evolution problems is that, if the system matrix does not change
during the time evolution, one needs to compute the factorization only once as a
preprocessing step.
The most popular methods to solve the linear systems arising from the dis-
cretization of Bidomain and Monodomain systems are iterative methods. In these
methods, approximate solutions of the linear system are computed iteratively until a
stopping criterion, typically involving the residual norm, is reached. Most published
approaches for the numerical discretisation of Monodomain and Bidomain models
leads to symmetric positive definite or semi-definite matrices. In this case, the
best choice of iterative method is the Conjugate Gradient method (CG). In some
cases, due to the formulation chosen and numerical scheme implemented, non-
symmetric matrices may arise, see e.g. [195,381,382], requiring alternative methods
as Krylov subspace solvers, in particular the Generalized Minimal Residual Method
(GMRES).
and
Z Z
..u; v//0 WD u i vi C u e ve
˝H ˝H
We denote by jjj jjj W U ! R the norm induced by the ..; // inner product, i.e.
Z Z Z
jjjujjj D 2
jrui j C
2
jrue j C
2
.ui ue /2 :
˝H ˝H ˝H
(b) Ellipticity:
Z Z Z
abid .u; u/ D Di rui rui C De rue rue C .ui ue /2
˝H ˝H ˝H
Z Z Z
˛i .rui / C ˛e
2
.rue / C 2
.ui ue /2
˝H ˝H ˝H
Given the previous notations, the stationary Bidomain system (8.3) can be rewritten
in the compact form:
Given vn , wnC1 , c nC1 , find unC1 2 U such that
where
or for
Z
i
Iapp D 0; and e
Iapp D 0: (8.7)
˝H
212 8 Parallel Solvers for the Bidomain System
Theorem 8.1. Given f 2 L2 .˝H /, problem (8.8) is equivalent to: find u 2 U such
that
but this follows from (8.8) with v D .1; 0/, hence (8.8) and (8.9) are equivalent.
Remark 8.1. We note that, if (H1)–(H2)–(H3) hold, problem (8.8) is H 2 -
regular, i.e.
h
where Iion is the finite elements interpolants of Iion .
8.1 Bidomain Variational Setting 213
Choose a finite element basis f'i g for Vh and let M D .mrs / and Ai;e D .ai;e
rs / be the
symmetric mass and stiffness matrices defined by
XZ XZ
mrs D 'r 's dx; rs D
ai;e .r'r /T Di;e .x/ r's dx:
E E E E
M M Ai 0
Abid D C
M M 0 Ae
Remark 8.2. Clearly, (8.14) arises from the finite element discretization of (8.8).
By defining the linear operator Abid W U h ! U h as
Abid uh D f h ; (8.15)
jjui uhi jjL2 .˝H / C jjue uhe jjL2 .˝H / C h jjju uh jjj: (8.16)
Let us define e h D .eih ; eeh / WD u uh D .ui uhi ; ue uhe /. Taking f D eih and
v D e h in (8.17), one has
jjeih jj2L2 .˝H / .eih ; eeh / D .eih ; eih eeh / D abid .e h ; '.eih //:
jjeih jj2L2 .˝H / .eih ; eeh / D abid .e h ; '.eih / vh / C jjje h jjj jjj'.eih / vh jjj:
8.2 Abstract Convergence Theory for Schwarz Methods 215
jjeih jj2L2 .˝H / .eih ; eeh / C h jjje h jjj jjeih jjL2 .˝H / : (8.19)
Taking now f D eeh and v D e h in (8.18), and following the same arguments, we
have
jjeeh jj2L2 .˝H / C .eih ; eeh / C h jjje h jjj jjeeh jjL2 .˝H / : (8.20)
jjeih jj2L2 .˝H / C jjeeh jj2L2 .˝H / C h jjje h jjj .jjeih jjL2 .˝H / C jjeeh jjL2 .˝H / /
In this section we recall the main results of the abstract Schwarz theory for Domain
Decomposition methods; for a more complete treatment, see the monographs
[489, 523].
Let U be a finite dimensional Hilbert space with inner product .; /, a.; / a
symmetric elliptic bilinear form on U U and f 2 U . Consider the following
problem:
Find u 2 U such that
.Au; v/ D a.u; v/ 8v 2 U:
Au D f: (8.22)
Im W Um ! U:
216 8 Parallel Solvers for the Bidomain System
We will refer to the case when Um U as the nested case. In this case Im are simply
imbedding operators. Assume also that on each subspace Um there is a symmetric
positive definite bilinear form
am .; / W Um Um ! R
Q m u; v/ D a.u; Im v/
am .T 8v 2 Um :
This formalism allows us to consider inexact solvers for the subproblems. If we use
exact solvers, then am .; / D a.; /. Now define the operators
Tm D Im TQ m W U ! U:
where BAS is the Additive Schwarz preconditioner. The Additive Schwarz method
replaces problem (8.22) by:
Find u 2 U such that
P P
Let um D TQ m T1
AS u, then m Im um D. m Tm /T1
AS u and
X X
am .um ; um / D Q m T1
am .T Q 1
AS u; Tm TAS u/
m m
X
D a.T1 1
AS u; Tm TAS u/
m
D a.T1
AS u; u/:
We recall for a linear operator L self-adjoint with respect to a.; / the Rayleigh
quotient characterization of the extreme eigenvalues,
a.Lu; u/ a.Lu; u/
min .L/ D min ; max .L/ D max :
u¤0 a.u; u/ u¤0 a.u; u/
We recall that for the condition number of L (.L/) in the norm induced by a, it
holds
max .L/
.L/ D :
min .L/
Lemma 8.4. The operators Tm , and hence TAS , are self-adjoint with respect to
a.; /; that is
Proof. The bound on the smallest eigenvalue follows immediately from Lemma 8.3
and Assumption 8.1.
We now bound the largest eigenvalue. We first treat the subspaces Um , for m D
1; : : : ; N . The special subspace U0 is treated separately. Observe that
X
N X
N X
N
a. Tm v; Tm v/ D a.Tm v; Tl v/;
mD1 mD1 m;lD1
X
N
Eml a.Tm v; Tm v/1=2 a.Tl v; Tl v/1=2 ;
m;lD1
8.2 Abstract Convergence Theory for Schwarz Methods 219
X
N
.E / a.Tm v; Tm v/;
mD1
X
N
!.E / am .TQ m v; TQ m v/;
mD1
X
N
!.E / a.v; Tm v/;
mD1
X
N
!.E /a.v; Tm v/;
mD1
X
N X
N
!.E /a.v; v/1=2 a. Tm v; Tl v/1=2 :
mD1 lD1
It follows
X
N X
N X
N
a.Œ Tm 2 v; v/ D a. Tm v; Tl v/;
mD1 mD1 lD1
! 2 2 .E /a.v; v/:
P
Thus the largest eigenvalue of . N 2 2 2
mD1 Tm / is bounded by ! .E /. Hence the
PN
largest eigenvalue of mD1 Tm is bounded by !.E / and
X
N
a. Tm v; v/ !.E /a.v; v/: (8.27)
mD1
Thus
which implies
Let V H be the space of trilinear finite elements associated to the coarse triangulation
T0 and define
Z
VQ H WD fv 2 V H W v D 0g; U0 D U H WD V H VQ H :
˝H
TQ m W U h ! Um
by
where BAS is the two-level Additive Schwarz preconditioner. As in the scalar elliptic
case, it is easy to see that the matrix form BAS of BAS is
X
N X
N
BAS D Bm D T
Rm A1
m Rm : (8.29)
mD0 mD0
8.3 Two-Level Additive Schwarz Methods for the Bidomain System 221
TAS u D g;
with C independent of h; H; ı.
Proof. We apply the general abstract Schwarz theory of the previous section based
on verifying the three assumptions stable decomposition, strengthened Cauchy-
Schwarz inequality and local stability, that provide an upper and lower bound on
.TAS /, see also [523, p. 40].
(a) Since we use exact local solvers, the local stability assumption holds true with
a unit constant.
(b) Using a standard coloring argument, see [523, Ch. 2], we obtain that the spectral
radius .E / of the matrix in the strengthened Cauchy-Schwarz inequality is
bounded from above by the number of colors, hence we have a constant upper
bound.
(c) It only remains to prove a stable decomposition for our subspace decomposition.
We will prove that this assumption is satisfied with a constant C02 D C.1 C H
ı /,
with C independent of h; H and ı.
Let
• QH W V h ! V0 be the L2 -projection onto V0 ,
• I h the linear interpolation operator onto V h and P
• m a partition of unity such that m 2 C01 .˝m0 /, 0 m 1, NmD1 m D 1,
C
jrm j2L1 ; (8.30)
ı2
222 8 Parallel Solvers for the Bidomain System
and
2
h
km Nm k2L1 .K/ C ; (8.31)
ı
X
N
uD Im um :
mD0
We can bound each scalar component as in the standard proof for scalar elliptic
problems (see e.g. [489, 523])
X
N XN
H
jum;i j2H 1 .˝H / C 1C jui j2H 1 .˝H / and jum;e j2H 1 .˝H /
mD0
ı mD0
H
C 1C jue j2H 1 .˝H / : (8.32)
ı
Since a finite bounded number (independent of h, H and ı) of um;i and um;e are
nonzero for any element K, summing over m we obtain
X
N
kum;i um;e k2L2 .K/ C kwi we k2L2 .K/ ;
mD1
X
N
kum;i um;e k2L2 .˝H / C kwi we k2L2 .˝H / :
mD1
8.4 Multilevel Additive Schwarz Methods for the Bidomain System 223
Since
kwi we k2L2 .˝H / D kui ue C QH .ue ui /k2L2 .˝H / 4kui ue k2L2 .˝H / ;
ku0;i u0;e k2L2 .˝H / D kQH .ui ue /k2L2 .˝H / kui ue k2L2 .˝H / ;
we obtain
X
N
kum;i um;e k2L2 .˝H / C kui ue k2L2 .˝H / :
mD0
X
N
H
.jum;i j2H 1 Cjum;e j2H 1 Ckum;i um;e k2L2 / C 1 C .jui j2H 1 Cjue j2H 1 Ckui ue k2L2 /;
ı
mD0
hence
X
N
H
jjjumjjj C 1 C
2
jjjujjj2:
mD0
ı
From the continuity and coercivity of abid .; /, it follows the stable decomposition
X
N
H
abid .um ; um / C 1 C abid .u; u/:
mD0
ı
The method described in the previous section uses two levels with a coarse and a
fine mesh. The smaller is H , the smaller are the local problems, but the algebraic
linear system corresponding to the problem in U0 becomes larger. We can then apply
recursively this two-level technique to the coarse problem and obtain an additive
multilevel method for the Bidomain system. We refer to the original works by
Dryja and Widlund [151], Zhang [584] and later Dryja, Sarkis and Widlund [153]
for an overview of these methods for scalar elliptic problems, including multilevel
diagonal scaling variants and different choices of coarse spaces.
Let us consider L 2 rather than just two mesh levels: on each level k D
0; 1; : : : ; L 1, ˝H is discretized with a shape-regular mesh Tk with elements of
characteristic size hk . Tk is a refinement of Tk1 , with level L 1 being the finest,
hence hL1 D h and h0 D H . On each level k D 0; : : : ; L 1, we define a finite
224 8 Parallel Solvers for the Bidomain System
jrkm jL1 C =ık and jjkm Nkm jjL1 .K/ C.hk =ık /; (8.33)
Define the interpolation operators Ikm as in the two-level case. Defining the
Q km W U h ! Ukm by
projections T
and Tkm D Ikm TQ km , we can introduce the Multilevel Additive Schwarz (MAS)
operator
XX
L1 Nk
TMAS D BMAS Abid WD T0 C Tkm : (8.34)
kD1 mD1
As mentioned before, approximate local solvers could be used as well. The matrix
form BMAS of the preconditioner BMAS is given by
XX
L1 Nk
BMAS D R0T A1
0 R0 C
T
Rkm A1
km Rkm ;
kD1 mD1
0
where Akm is the local stiffness matrix of subdomain ˝km at level k and Rkm is the
0
restriction matrix from the finest level to subdomain ˝km , see [489] for more details.
We can then prove our main result for the multilevel case.
8.4 Multilevel Additive Schwarz Methods for the Bidomain System 225
Theorem 8.5. The condition number of the multilevel additive Schwarz operator
TMAS defined in (8.34) is bounded by
hk1
.TMAS / C max 1C ;
kD1;:::;L1 ık
where C is a constant independent of the mesh sizes hk and the number of levels L.
Proof. Upper bound. The proof of the optimal upper bound follows from the steps
of the original proof by Zhang [584], from Lemma 3.2 to Lemma 3.5. These
results can be applied to our case since they hold also for general conforming
finite elements (Remark 3.2 in [584]) and uniform elliptic operators (Remark 3.3
in [584]).
Lower bound. In order to prove the left inequality, we need a stable decompo-
sition of u 2 U h . For each level k D 0; : : : ; L 2, we introduce the projections
Qk W U h ! U hk induced by the abid .; / bilinear form, i.e.
We denote by Qik and Qek the two components of Qk , i.e. Qk u D .Qik u; Qek u/. It
follows from Theorem 8.2 that, given u 2 U h
u D u0 C u1 C : : : C uL1 :
X
Nk
jjjukmjjj2 C jjjuk jjj2 :
mD1
226 8 Parallel Solvers for the Bidomain System
jukm;i j2H 1 .K/ D jIhk .km uk;i /j2H 1 .K/ D jIhk .Nkm uk;i / Ihk ..km Nkm /uk;i /j2H 1 .K/
The last term above is zero for all elements K in the interior of an ˝km , since km D
Nkm D 1 on the non overlapping regions. Therefore, summing over all the elements,
the last term only includes the elements in the overlapping regions ık ;m . Hence
X
Nk X
Nk X
jukm;i j2H 1 .˝H / C.juk;i j2H 1 .˝H / C h2 N
k jjIhk ..km km /uk;i jjL2 .K/ /
2
X
Nk
C.juk;i j2H 1 .˝H / C ık2 jjuk;i jj2L2 .ı / /: (8.37)
k ;m
mD1
We can bound the last terms using the following result by Dryja and Widlund [152]:
Lemma 8.5. Let ˝H be a shape regular region, in R2 or R3 , of diameter H and let
ı be a strip along its boundary of width ı > 0. Then
H 1
jjujj2L2 .ı / C ı 2 Œ.1 C /juj2H 1.˝H / C jjujj2L2 .˝H / :
ı Hı
Applying this lemma to (8.37), we obtain
X
Nk
hk1 1
jukm;i j2H 1 .˝H / C.juk;i j2H 1 .˝H / C .1 C /juk;i j2H 1 .˝H / C jjuk;i jj2L2 .˝H / /;
mD1
ık hk1 ık
X
Nk
hk1 X
Nk
hk1
jukm;i j2H 1 .˝H / C.1C /jjjuk jjj2 ; jukm;e j2H 1 .˝H / C.1C /jjjuk jjj2 :
mD1
ık mD1
ık
X
Nk
jjukm;i ukm;e jj2L2 .˝H / C jjuk;i uk;e jj2L2 .˝H / C jjjuk jjj2 ;
mD1
X
Nk
hk1
jjjukm jjj2 C.1 C /jjjuk jjj2 :
mD1
ık
8.5 Numerical Results for Multilevel Schwarz Preconditioners 227
Using the coercivity and the continuity of the Bidomain bilinear form abid .; / and
summing over the levels, we finally obtain the stable decomposition
XX
L1 Nk
hk1
abid .u0 ; u0 / C abid .ukm ; ukm / C max .1 C /abid .u; u/:
kD1;:::;L1 ık
kD1 mD1
The numerical experiments were performed on the Linux Cluster IBM CLX/1024 of
the Cineca Consortium (www.cineca.it), with 1024 processors (Intel Xeon Pentium
IV 3 GHz, 512 KB cache), grouped into 512 biprocessor nodes. The total RAM
is 1 TB and the peak performance is declared to 6.1 Tflops. We were allowed
to use up to 240 processors in each run. Our FORTRAN code is based on the
parallel library PETSc from the Argonne National Laboratory [24, 25]. We solve
the Bidomain system on a domain ˝H that is either a cartesian slab (see Fig. 8.2,
left) or the image of a cartesian slab using ellipsoidal coordinates, yielding a portion
of a truncated ellipsoid (see Fig. 8.1). This second choice can be used in simulations
with idealized ventricular geometries, see e.g. [104, 125] These two choices allow
us to also test the performance of our multilevel preconditioner in presence of severe
domain deformations.
The values of the conductivity coefficients used in the numerical tests are
reported in Table 8.1 and the capacitance per unit volume is set to cm D 1 mF cm3 .
The values of the coefficients and parameters in the LR1 model are given in the
original paper [308].
For both types of domains, we denote a level mesh by the triplet T D Ti Tj Tk
indicating the number of subdivisions in each coordinate direction. When we scale
up the mesh, for brevity we define cT D cTi cTj cTk . In all tests, the the
finest mesh size is h D 0:1 mm, which has been shown to accurately resolve the
steep excitation fronts in the Bidomain–LR1 model without numerical artifacts, see
[104, 125].
We perform 5 time steps of 0.05 ms each with the semi-implicit Euler scheme
described in Chap. 7 and at each time step we solve iteratively the discrete Bidomain
system (8.12) by the Preconditioned Conjugate Gradient (PCG, see [445]) method
with the L-level MAS preconditioner (8.34). The PCG initial guess is the discrete
solution of the previous time step and the stopping criterion is a 104 residual
reduction. The problem on the coarsest level is solved by PCG with 108 residual
reduction as stopping criterion. In all runs, we report the PCG condition number,
extreme eigenvalues, iteration counts and cpu timings relative to the 5th time step,
in order to avoid special right-hand side configurations associated with the first few
time steps.
228 8 Parallel Solvers for the Bidomain System
32 = 4⋅ 4⋅ 2
8 = 2⋅ 2⋅ 2 16 = 4⋅ 2⋅ 2
240 = 15⋅ 8⋅ 2
64 = 8⋅ 4⋅ 2 128 = 8⋅ 8⋅ 2
Fig. 8.1 Portions of ellipsoidal domain decomposed in 8, 16, 32, 64, 128, 240 subdomains for
scaled speedup test
Test MAS-3D-1. Increasing levels and increasing global size, cartesian slab. We
first consider in Table 8.2 a cartesian slab, see Fig. 8.2, left panel, and increase the
number of levels L in our MAS preconditioner from 3 to 6, while increasing the fine
meshes as Ti C1 D 2Ti , starting with a fixed coarse mesh T0 D 8 8 2. The number
of processors and subdomains is fixed at 128, but the global problem size increases
8.5 Numerical Results for Multilevel Schwarz Preconditioners 229
Table 8.2 Test MAS-3D-1. Increasing number of levels L for cartesian slabs on Fig. 8.2 with
increasing fine meshes TiC1 D 2Ti starting with a fixed coarse mesh T0 D 8 8 2; fixed number
of processors D 128
L Meshes d.o.f. D max = min it. Time
3 T0 ; 2T0 ; 4T0 19 602 7.76 D 4.217/0.543 15 1.05
4 T0 ; 2T0 ; 4T0 ; 8T0 143 650 7.97 D 4.501/0.564 17 2.74
5 T0 ; 2T0 ; 4T0 ; 8T0 ; 16T0 1.1 M 7.28 D 5.241/0.720 19 5.20
6 T0 ; 2T0 ; 4T0 ; 8T0 ; 16T0 ; 32T0 8.6 M 6.82 D 5.783/0.847 18 11.29
20
18
it.
16
14
12
10
κ
8 2
4 λ
max
2 λmin
0
3 4 5 6
LEVELS
Fig. 8.2 Left: Cartesian slab subdivided into 128 D 8 8 2 subdomains. Right: plot of
min ; max ; ; it. Reported in Table 8.2
from 19 602 d.o.f. (for L D 3, fine mesh T2 D 4T0 ) to 8.6 millions d.o.f. (for
L D 6, fine mesh T5 D 32T0 ). The MAS preconditioner has overlap size ık D hk
and ILU(0) local solvers for the subdomains of each level. The main results of this
table are also plotted in Fig. 8.2, right panel. The condition numbers and iteration
counts do not grow with increasing L (actually they even decrease for higher values
of L), in spite of the strong growth of the problem size. Of course, the cpu timings
reported in the last column grow because the global problem size grow, due to the
growth of the subdomain size on the finest mesh from 43 .L D 3/ to 323 .L D 6/.
These results confirm the optimality result of Theorem 8.5, with
constant bounds
hk1
independent of L and depending only on max 1C :
kD1;:::;L1 ık
Test MAS-3D-2. Scaled speedup on ellipsoidal domains. In Table 8.3, we then
consider a scaled speedup test for our MAS operator with 2, 3, 4 levels. Again,
we fix the overlap size ık D hk and ILU(0) local solvers on the subdomains
of each level. The local size of each subdomain on the finest mesh is kept fixed
at the value 483 and each subdomain is assigned to one processor. The number
of subdomains (hence number of processors) is increased from 8 D 2 2 2 to
240 D 15 8 2 (first and second columns), forming increasing domains ˝H that are
portions of a truncated ellipsoidal domain as shown in Fig. 8.1. These subdomains
230 8 Parallel Solvers for the Bidomain System
Table 8.3 Test MAS-3D-2. Condition numbers, extreme eigenvalues, iteration counts, cpu timings
for 2, 3, 4 level MAS, ık D hk , ILU(0) local solvers: scaled speedup with fixed subdomain size
48 48 48 on the ellipsoidal domains of Fig. 8.1
2 levels: T0 ; T1 D 48T0
Procs. T0 T1 D 48T0 d.o.f. (M) D max = min it. Time
8 222 96 96 96 1.8 183.50 D 2.268/0.012 45 53.1
16 422 192 96 96 3.6 172.44 D 2.265/0.013 45 65.0
32 442 192 192 96 7.2 228.63 D 2.455/0.011 50 65.8
64 842 384 192 96 14.4 226.95 D 2.479/0.011 50 74.7
128 882 384 384 96 28.7 150.31 D 2.413/0.016 43 66.2
240 15 8 2 720 384 96 53.8 149.98 D 2.415/0.016 43 68.1
3 levels: T0 ; T1 D 4T0 ; T3 D 48T0
Procs. T0 T2 D 48T0 d.o.f. (M) D max = min it. Time
8 222 96 96 96 1.8 21.70 D 2.903/0.134 23 23.9
16 422 192 96 96 3.6 23.33 D 3.092/0.132 24 30.4
32 442 192 192 96 7.2 25.07 D 3.263/0.130 24 34.2
64 842 384 192 96 14.4 25.56 D 3.269/0.128 25 32.0
128 882 384 384 96 28.7 30.70 D 3.347/0.109 26 40.2
240 15 8 2 720 384 96 53.8 30.99 D 3.381/0.109 26 40.8
4 levels: T0 ; T1 D 4T0 ; T2 D 16T0 ; T3 D 48T0
Procs. T0 T3 D 48T0 d.o.f. (M) D max = min it. Time
8 222 96 96 96 1.8 10.81 D 3.815/0.348 20 27.2
16 422 192 96 96 3.6 11.65 D 4.141/0.355 20 29.6
32 442 192 192 96 7.2 10.18 D 4.137/0.406 19 31.0
64 842 384 192 96 14.4 9.95 D 4.117/0.414 19 33.1
128 882 384 384 96 28.7 9.96 D 4.089/0.411 19 34.8
240 15 8 2 720 384 96 53.8 9.95 D 4.083/0.410 19 36.0
are the elements of the coarse mesh T0 . The finest mesh is then 48T0 , since the
local mesh size is kept fixed at 483 . The intermediate level meshes in case of 3
and 4 levels are T1 D 4T0 and T2 D 16T0 . With these data, the global size of
the discrete Bidomain system increases then from 1.8 million d.o.f. for the smallest
domain with 8 subdomains to 53.8 million d.o.f. for the largest domain with 240
subdomains. In each of these cases, the last columns of Table 8.3 report the condition
number , extreme eigenvalues max ; min , iteration counts it: and cpu timings of the
PCG solver with MAS preconditioner. The same data are plotted in Fig. 8.3 vs.
the processor count. The results clearly show the scalability of the MAS operator,
since all the reported quantities seem to approach constant values as the number
of subdomains (hence processors) increases. As expected, these asymptotic values
improve with the number of levels, since for increasing L the MAS preconditioner
becomes more powerful.
Table 8.4 reports the results of a scaled speedup test run recently on the BlueGene
Q cluster of the CINECA consortium (www.cineca.it). The preconditioner consid-
ered is the MAS(4), the domains are portions of a truncated ellipsoid, the number
8.5 Numerical Results for Multilevel Schwarz Preconditioners 231
103 55
50
45 2 LEVELS
2 LEVELS
MAS CONDITION N.
MAS ITERATIONS
40
102
35
3 LEVELS 30 3 LEVELS
25
20
101
4 LEVELS 15 4 LEVELS
10
5
100 0
0 50 100 150 200 250 0 50 100 150 200 250
PROCESSORS (SUBDOMAINS) PROCESSORS (SUBDOMAINS)
Fig. 8.3 Test MAS-3D-2. Scalability of 2, 3, 4 level MAS: condition numbers and iteration counts
reported in Table 8.3
Table 8.5 Test MAS-3D-3. Increasing number of levels L for fixed coarse mesh T0 D 8 8 2
(486 d.o.f.) and fine mesh TL1 D 48T0 D 384 384 96 (28.7M d.o.f.), intermediate meshes
TiC1 D 2Ti ; ellipsoidal domain of Fig. 8.1, fixed number of processors D 128
L Meshes D max = min it. Time
2 T0 ; 48T0 150.31 D 2.413/0.016 43 66.2
3 T0 ; 2T0 ; 48T0 86.55 D 3.375/0.039 38 158.5
4 T0 ; 2T0 ; 4T0 ; 48T0 43.31 D 4.374/0.101 33 124.9
5 T0 ; 2T0 ; 4T0 ; 8T0 ; 48T0 17.82 D 5.118/0.287 24 139.1
6 T0 ; 2T0 ; 4T0 ; 8T0 ; 16T0 ; 48T0 10.94 D 5.892/0.539 20 32.7
of processors varies from 64 to 16 384, while the number of d.o.f. increases from
4 million to about 1 billion.
Test MAS-3D-3. Increasing levels and fixed global size, ellipsoidal domain.
In Table 8.5, we fix the coarse mesh T0 D 8 8 2 (486 d.o.f.) and fine mesh
TL1 D 48T0 D 384 384 96 (28.7M d.o.f.), hence the number of subdomains
232 8 Parallel Solvers for the Bidomain System
and processors are fixed to 128 (the number of coarse elements). We then vary the
number of levels L from 2 to 6, with intermediate meshes defined by Ti C1 D 2Ti .
The ellipsoidal domain is the one of Fig. 8.1 decomposed into 128 subdomains. Both
the condition numbers and iteration counts improve considerably when the number
of levels increases, because the improvement of the minimum eigenvalues more than
compensate a slight increase of the maximum eigenvalue.
Test MAS-3D-4. Scaled speedup with variable overlap and ILU/LU local
solvers, cartesian slab. We now focus on the effect of two variables that we have
kept fixed so far: the overlap size ık and the use of inexact/exact local solvers.
Therefore, we repeat a scaled speedup test as in Table 8.1 but on slab domains and
with smaller subdomain size 163 in order to be able to run our MAS preconditioner
with exact LU local solvers. The finest mesh is then always TL1 D 16T0 , while
the intermediate meshes for L D 3; 4 are 2T0 and 4T0 . Table 8.6 reports the
Table 8.6 Test MAS-3D-4. Iteration counts and condition numbers of 2, 3, 4 level MAS: scaled
speedup with fixed subdomain size 16 16 16 and varying overlap size ık D 1; 2; 4 hk , with
inexact ILU(0) local solvers (in brackets exact LU local solvers)
2 levels: T0 ; T1 D 16T0
ık D hk ık D 2 hk ık D 4 hk
Procs. T0 it. it. it.
8 2 2 2 38 (23) 104.5 (36.1) 39 (17) 85.3 (14.6) 34 (14) 58.7 (10.2)
18 3 3 2 47 (24) 143.4 (51.5) 49 (18) 129.1 (19.6) 37 (16) 72.0 (12.3)
32 4 4 2 45 (21) 123.3 (33.1) 41 (17) 96.9 (17.2) 39 (17) 80.3 (12.6)
50 5 5 2 45 (21) 148.1 (30.1) 41 (17) 96.8 (16.7) 43 (18) 95.9 (13.9)
72 6 6 2 41 (20) 117.2 (27.5) 42 (18) 97.9 (19.1) 43 (18) 97.3 (13.9)
128 8 8 2 39 (20) 92.0 (24.5) 42 (18) 100.8 (17.5) 44 (17) 95.4 (11.5)
3 levels: T0 ; T1 D 2T0 ; T2 D 16T0
ık D hk ık D 2 hk ık D 4 hk
Procs. T0 it. it. it.
8 2 2 2 18 (14) 11.6 (6.5) 18 (12) 10.5 (4.6) 17 (12) 9.7 (3.7)
18 3 3 2 19 (14) 13.3 (6.6) 18 (13) 12.0 (4.8) 18 (13) 11.6 (5.1)
32 4 4 2 19 (14) 13.6 (6.4) 19 (13) 12.9 (5.0) 19 (13) 12.2 (4.4)
50 5 5 2 19 (14) 13.6 (6.1) 19 (14) 12.9 (4.9) 19 (13) 12.3 (4.9)
72 6 6 2 19 (14) 13.4 (6.1) 19 (14) 12.8 (4.8) 19 (13) 12.1 (5.2)
128 8 8 2 19 (14) 13.5 (6.0) 19 (13) 12.8 (4.7) 19 (14) 12.0 (4.6)
4 levels: T0 ; T1 D 2T0 ; T2 D 4T0 ; T3 D 16T0
ık D hk ık D 2 hk ık D 4 hk
Procs. T0 it. it. it.
8 2 2 2 18 (16) 8.9 (8.1) 18 (16) 8.9 (7.3) 18 (16) 8.3 (6.7)
18 3 3 2 20 (18) 14.6 (12.3) 20 (18) 14.4 (10.9) 19 (17) 13.9 (8.2)
32 4 4 2 17 (17) 7.5 (7.9) 17 (16) 7.2 (7.8) 17 (16) 7.0 (7.8)
50 5 5 2 17 (17) 7.4 (7.5) 17 (16) 7.1 (7.3) 17 (16) 6.8 (7.2)
72 6 6 2 17 (17) 7.3 (7.1) 16 (16) 7.0 (6.9) 16 (16) 6.7 (6.8)
128 8 8 2 16 (16) 7.3 (6.6) 16 (16) 7.0 (6.3) 16 (16) 6.7 (6.1)
8.5 Numerical Results for Multilevel Schwarz Preconditioners 233
150 55
50
MAS CONDITION NUMBER (ILU(0))
2 LEVELS (δ = 1,2,4)
45
Fig. 8.4 Test MAS-3D-4. Scalability of 2, 3, 4 level MAS with ILU(0) local solvers and variable
overlap ık D 1; 2; 4 hk : condition numbers (left) and iteration counts (right) reported in Table 8.6
55 30
50
25
45 o = 2 LEVELS (δ = 1,2,4)
MAS CONDITION NUMBER (LU)
= 2 LEVELS (δ = 1,2,4)
40 * = 3 LEVELS (δ = 1,2,4)
MAS ITERATIONS (LU)
20
35 = 4 LEVELS (δ = 1,2,4)
30
15
25
10 5
5
0 0
0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140
PROCESSORS (SUBDOMAINS) PROCESSORS (SUBDOMAINS)
Fig. 8.5 Test MAS-3D-4. Scalability of 2, 3, 4 level MAS with LU local solvers and variable
overlap ık D 1; 2; 4 hk : condition numbers (left) and iteration counts (right) reported in Table 8.6
iteration counts and condition numbers of 2, 3, 4 level MAS with inexact ILU(0)
local solvers; in brackets are the same quantities for MAS with exact LU local
solvers. We consider three cases of increasing overlap size ık D 1; 2; 4 hk . The
ILU(0) results are then plotted in Fig. 8.4, and the LU results in Fig. 8.5, where the
2, 3, 4 level data are denoted by ı; ?; : : :, respectively. The scalability of the MAS
preconditioner is evident in all cases; as expected, the LU results are better than
the ILU(0) results and increasing the overlap size improves the conditioning of the
preconditioned operator but only by a marginal amount when L > 2 (in fact the
iteration counts are basically independent of ı for L > 2), except in the 2-level
ILU(0) case, where the results are a bit erratic. Increasing the levels improve the
results in the ILU(0) case, but in the LU case the 3-level MAS seems to be slightly
better conditioned than the 4-level MAS for this choice of meshes and parameters.
234 8 Parallel Solvers for the Bidomain System
NX
L1
uL1 T
RL1m A1
L1m RL1m rL1
mD1
uL2 T
RL2m A1
L2m RL2m rL2
mD1
:::
u0 A1
0 r0
u1 u1 C R0T u0
X
N1
u1 u1 C T
R1m A1
1m R1m .r1 A1 u1 /
mD1
:::
uL1 uL1 C RL2
T
uL2
X
NL1
uL1 uL1 C T
RL1m A1
L1m RL1m .r
L1
AL1 uL1 /
mD1
u uL1
In the following, we will denote the MHS preconditioner with L levels MHS(L).
8.5 Numerical Results for Multilevel Schwarz Preconditioners 235
20
−10
−20
10
−30
−40
0
−50
−60 −10
−70
−20
−80
Fig. 8.6 Test MHS-2D-1. Patterns of level lines of the transmembrane and extracellular potentials
during the excitation phase (t D 25 ms after the stimulus). Reported below each panel are the
minimum, maximum and step in mV of the displayed map
li;e 0
Di;e D :
0 ni;e
The global mesh is fixed to be of 256256 elements (132 098 unknowns), 512512
elements (526 338 unknowns) and 1 024 1 024 elements (2 101 250 unknowns).
The cases of 4, 16, 64 subdomains (processors) and 2–7 levels are considered. The
model is run for 200 time steps of 0.05 ms, i.e. for a time interval of 10 ms on
the CLX Linux cluster. In Tables 8.7 and 8.8, we report the average number of
PCG iterations per time step, the average condition number and the average time
needed to solve the linear system. The results show the scalability of the MHS
preconditioner, since both the condition numbers and the iteration counts remain
almost constant when the number of subdomains and levels increase.
Test MHS-2D-2. Increasing levels and increasing global size, square domain.
The aim of this test is to study the behavior of the preconditioner increasing the
number of levels from 2 to 9. The size of the fine mesh varies form 4 4 elements
(50 unknowns) with 2 levels to 512 512 elements (526 338 unknowns) with 9
levels, hence the dimensions of the square domain increase from 0:04 0:04 to
5:12 5:12 cm2 . The number of subdomains (and processors) is kept fixed to 4. As
236 8 Parallel Solvers for the Bidomain System
Table 8.7 Test MHS-2D-1. Condition numbers, iteration counts, cpu timings for 2, 3, 4 level
MHS. Fixed global size, increasing the number of subdomains, square domain
2 levels 3 levels 4 levels
Procs. Mesh it. Time it. Time it. Time
4 2562 1.05 4 2:93 1.05 4 1:17 1.05 4 0.98
16 2562 1.04 4 0:84 1.05 4 0:43 1.05 4 0.35
64 2562 1.04 4 0:68 1.05 4 0:43 1.05 4 0.35
4 5122 1.04 2 15:04 1.05 3 3:60 1.05 3 3.14
16 5122 1.04 2 4:02 1.05 3 1:25 1.05 3 1.15
64 5122 1.04 2 1:44 1.05 3 0:61 1.05 3 0.50
4 1 0242 1.06 2 81:89 1.07 2 21:96 1.07 2 14.20
16 1 0242 1.06 2 20:71 1.07 2 6:08 1.07 2 3.91
64 1 0242 1.06 2 6:91 1.07 2 1:98 1.07 2 1.29
Table 8.8 Test MHS-2D-1. Same format as in Table 8.7 for 5, 6, 7 level MHS
5 levels 6 levels 7 levels
Procs. Mesh it. Time it. Time it. Time
4 2562 1.05 4 0:98 1.06 4 0:98 1.06 4 0.98
16 2562 1.05 4 0:38 1.05 4 0:33 1.05 4 0.36
64 2562 1.05 4 0:37 1.12 5 0:36 – – –
4 5122 1.05 3 3:10 1.05 3 3:08 1.05 3 3.09
16 5122 1.05 3 0:84 1.05 3 0:95 1.05 3 0.95
64 5122 1.05 3 0:47 1.05 3 0:43 1.05 3 0.57
4 1 0242 1.07 3 14:55 1.07 3 13:20 1.07 3 13.26
16 1 0242 1.07 3 3:44 1.07 3 3:23 1.07 3 3.28
64 1 0242 1.07 3 1:19 1.07 3 1:30 1.07 3 1.28
in the previous tests, we run the model for 10 ms after the stimulus. Table 8.9 reports
the average minimum and maximum eigenvalues, the average condition number, the
average number of PCG iterations and the average solving time per time step. The
condition numbers and iteration counts do not grow with increasing L, in spite of
the strong growth of the problem size. Of course, the cpu timings reported in the
last column grow because the global problem size grow, due to the growth of the
subdomain size on the finest mesh.
Test MHS-2D-3. Scaled speedup on square domains. In this test, we vary the
number of subdomains from 4 to 64, keeping fixed the local mesh in each subdomain
to 64 64 elements (8 450 unknowns), hence varying the global number of degrees
of freedom (d.o.f.) from 33 800 in the smallest case with 4 subdomains to 540 800 in
the largest with 64 subdomains. The number of levels varies from 2 to 7. As in test
MHS-2D-1, we simulate the initial depolarization of a myocardial section, running
the model for 200 time steps on the CLX cluster of CINECA. Tables 8.10 and 8.11
report the average number of PCG iterations and the average solving time per time
8.5 Numerical Results for Multilevel Schwarz Preconditioners 237
Table 8.9 Test MHS-2D-2. Increasing number of levels L for square domains with increasing fine
meshes TiC1 D 2Ti starting with a fixed coarse mesh T0 D 2 2; fixed number of processors D 4
L Meshes d.o.f. D max = min it. Time
2 T0 ; 2T0 50 1.16 D 1.05/0.90 4 0.01
3 T0 ; 2T0 ; 4T0 162 1.05 D 1.01/0.97 3 0.01
4 T0 ; 2T0 ; 4T0 ; 8T0 578 1.04 D 1.00/0.97 3 0.02
5 T0 ; 2T0 ; 4T0 ; 8T0 ; 16T0 2 178 1.04 D 1.00/0.97 3 0.03
6 T0 ; 2T0 ; 4T0 ; 8T0 ; 16T0 ; 32T0 8 450 1.05 D 1.01/0.96 4 0.08
7 T0 ; 2T0 ; 4T0 ; : : : ; 32T0 ; 64T0 33 282 1.05 D 1.01/0.96 4 0.25
8 T0 ; 2T0 ; 4T0 ; : : : ; 64T0 ; 128T0 132 098 1.05 D 1.01/0.96 4 0.99
9 T0 ; 2T0 ; 4T0 ; : : : ; 128T0 ; 256T0 526 338 1.01 D 1.00/0.99 3 3.12
Table 8.10 Test MHS-2D-3. Condition numbers, iteration counts, cpu timings for 2, 3, 4 level
MHS. Scaled speedup with fixed subdomain size 64 64 on square domains
2 levels 3 levels 4 levels
Procs. d.o.f. it. Time it. Time it. Time
4 33 800 1.04 4 0.46 1.05 4 0.25 1.05 4 0.25
9 76 050 1.04 4 0.63 1.04 4 0.33 1.05 4 0.32
16 135 200 1.05 4 0.87 1.05 4 0.43 1.05 4 0.45
25 211 250 1.05 4 1.07 1.05 4 0.51 1.05 4 0.45
36 304 200 1.05 4 1.43 1.05 4 0.53 1.05 4 0.55
49 414 050 1.05 4 1.57 1.05 4 0.66 1.05 4 0.66
64 540 800 1.05 4 1.94 1.05 4 0.86 1.05 4 0.75
Table 8.11 Test MHS-2D-3. Same format as in Table 8.10 for 5, 6, 7 level MHS
5 levels 6 levels 7 levels
Procs. d.o.f. it. Time it. Time it. Time
4 33 800 1.05 4 0.23 1.05 4 0.25 1.05 4 0.26
9 76 050 1.04 3 0.29 1.05 4 0.28 1.05 4 0.30
16 135 200 1.05 4 0.37 1.05 4 0.33 1.05 4 0.36
25 211 250 1.05 4 0.40 1.05 4 0.41 1.05 4 0.43
36 304 200 1.05 4 0.44 1.05 4 0.44 1.07 4 0.43
49 414 050 1.05 4 0.53 1.05 4 0.48 1.08 4 0.50
64 540 800 1.05 4 0.66 1.05 4 0.55 1.09 5 0.58
step. We observe that, in this case too, both condition numbers and iteration counts
remain constant when increasing the number of subdomains and levels.
Test MHS-3D-1. Increasing levels and increasing global size, cartesian slab. As
in test MAS-3D-1 of the previous section, the behavior of the MHS preconditioner
238 8 Parallel Solvers for the Bidomain System
Table 8.12 Test MHS-3D-1. Increasing number of levels L for cartesian slabs with increasing fine
meshes TiC1 D 2Ti starting with a fixed coarse mesh T0 D 4 4 2; fixed number of processors
D 32
L Meshes d.o.f. D max = min it. Time
3 T0 ; 2T0 ; 4T0 5 202 1.97D1.18/0.60 7 0.13
4 T0 ; 2T0 ; 4T0 ; 8T0 37 026 1.53D1.03/0.67 5 0.27
5 T0 ; 2T0 ; 4T0 ; 8T0 ; 16T0 278 850 1.46D1.01/0.68 5 0.94
6 T0 ; 2T0 ; 4T0 ; 8T0 ; 16T0 ; 32T0 2.1M 1.34D1.01/0.74 5 5.9
Table 8.13 Test MHS-3D-2. Condition numbers, iteration counts, cpu timings for 3, 4 level MHS.
Scaled speedup with fixed subdomain size 48 48 48 on the ellipsoidal domains of Fig. 8.1
3 levels 4 levels
Procs. d.o.f. (M) it. Time it. Time
8 1:8 1.17 4 19:11 1.18 4 16.67
16 3:6 1.17 4 21:21 1.18 4 19.82
32 7:2 1.18 4 21:86 1.20 5 20.57
64 14:4 1.18 4 23:8 1.20 5 21.32
128 28:7 1.30 5 25:83 1.30 5 22.02
240 53:8 1.30 5 28:27 1.30 5 23.13
Table 8.15 Test MHS-3D-3. Presence of transmural ischemia: comparison of the three precon-
ditioners BJ, MG(5) and MHS(5) in presence of discontinuous conductivity coefficients due to
ischemia
BJ MG(5) MHS(5)
Procs. it. time it. Time it. Time
8 3 060 156 33:61 2.86 9 16:58 2.33 8 16.73
16 3 384 163 17:48 2.96 10 10:64 2.39 8 8.58
32 3 500 167 10:16 2.95 10 5:14 2.41 8 5.47
64 4 268 183 5:68 3.35 11 3:47 2.83 9 3.44
seems to be attained with 5 levels, hence in the following tests we focus on the
MHS(5).
Test MHS-3D-3. Presence of transmural ischemia. The domain considered
in this case is an anisotropic slab of dimensions 1:28 1:28 0:32 cm3 , with a
transmural ischemic region of dimensions 0:320:320:32 cm3 located at the center
of the slab. The portion of tissue is discretized by a cartesian grid of 128 128 32
elements (1 098 306 unknowns). The excitation process is started by applying a
stimulus of 200 A=cm3 for 1 ms on a small area of 2 2 2 elements at a vertex of
the endocardium. The ischemic condition is modeled by increasing the extracellular
concentration of potassium in the LR1 model from 5.4 mV (control) to 20 mV
(ischemia) (see [569]) and reducing the conductivity coefficients in the ischemia
region as indicated in Table 8.1 (see [235]), hence the conductivity coefficients
present discontinuities on the boundaries of the ischemic region and this makes the
resolution more difficult. See also Chap. 3 for more details on modeling ischemic
conditions.
We compare MHS(5) with the standard one-level Block Jacobi preconditioner
(BJ) and the multigrid preconditioner with 5 levels (MG(5)) proposed in [552].
We observe that the hybrid method is also in this case scalable and comparable
to (sometimes better than) the multigrid preconditioner MG(5) (Table 8.15).
MHS-3D-4. Complete cardiac cycle. In this last test, we simulate a complete
heartbeat (400 ms, about 3 000 time steps) in a portion of ventricle having dimension
1:92 1:92 0:48 cm3 , discretized by a cartesian grid of 192 193 49 nodes
(3:6 106 d.o.f.). We run the simulation on 36 processors of the Linux cluster of
Milan (cluster.mat.unimi.it). Table 8.16 reports the average PCG iterations per time
240 8 Parallel Solvers for the Bidomain System
Table 8.16 Test MHS-3D-4. Prec. it. Time (s) Total time
Complete cardiac cycle
BJ 205 46:02 29 h 49 min
MG(5) 8 11:11 7 h 21 min
MHS(5) 6 9:67 6 h 26 min
it average PCG iterations per time step,
time average execution time per time step
in seconds, total time total simulation time
260 20
240
18 MG(5)
MHS(5)
220
16
ITERATIONS
200
14
180
12
160
10
PCG
140
8
120
100 6
80 4
0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400
Fig. 8.7 Test MHS-3D-4. Time evolution of the PCG iterations with BJ preconditioners (left) and
multilevel preconditioners MG(5), MHS(5) (right)
Fig. 8.8 Test MHS-3D-4. Patterns of level lines of the transmembrane and extracellular potentials
during the excitation phase (t = 40 ms). Reported below each panel are the minimum, maximum
and step in mV of the displayed map
step, the average execution time per time step and the total simulation time. The
detailed iteration counts as function of time during the complete heartbeat are shown
in Fig. 8.7 (left panel for BJ and right panel for MG(5) and MHS(5)). Figure 8.8
shows the spatial maps of the transmembrane and extracellular potentials computed
40 ms after the stimulus was given at a vertex of the domain, i.e. during the
excitation phase. Figure 8.9 shows the transmembrane and extracellular potentials
8.6 Block Preconditioners for the Bidomain System 241
−50 −10
−20
−100 −30
0 100 200 300 400 0 100 200 300 400
TIME (ms) TIME (ms)
Fig. 8.9 Test MHS-3D-4. Time evolution at a fixed point of the transmembrane and extracellular
potentials, computed with the three methods
computed in a fixed point of the domain by the three methods (the three plots are
indistinguishable because superimposed). MHS(5) reduces the computational time
of about 78 and 12 %, compared to BJ and MG(5) respectively.
M C Ai M M C Ai Ai
APP D ; APE D ;
M M C Ae Ai A
Cm
where D is proportional to the inverse time step size, see (7.9).
Denoting by
A11 A12
AD
A12 A22
242 8 Parallel Solvers for the Bidomain System
either one of APP , APE , we will consider the following classical block precondition-
ers for A (see Axelsson [20]):
• The block-diagonal preconditioner
B1 0
BD D I (8.38)
0 B2
I 0 B1 A12
BF D : (8.39)
A12 B1
1 I 0 B2
The main abstract results for these preconditioners are given in the following two
propositions, see Axelsson [20] for a proof. For both propositions, we define the
constant
vT A12 A1
11 A12 v
2
D sup T
(8.40)
v2Rn nKer.A22 / v A22 v
q
and the function .x/ D 12 .1 C x/ C 1
4 .1 x2/ C x 2.
1 ı2 1
where r2 D minf ˛ı221 ; ˛2 1 g, and ˛2 > 1 and/or ı2 > 1,
1
1 1 minf˛1 ; ı1 g
max .BF A/ 1 .r1 / ;
1 2
B1 1
1 D scalar BMAS preconditioner for ct M C Ai ;
B1 1
2 D scalar BMAS preconditioner for ct M C Ae ; (8.41)
B1 1
1 D scalar BMAS preconditioner for ct M C Ai ;
B1 1
2 D scalar BMAS preconditioner for Ai C Ae : (8.42)
We could also build our scalar blocks using the hybrid multilevel preconditioner
1 1
BMHS instead of BMAS .
The following convergence rate bound for the block-diagonal preconditioner can
be found in [373].
Lemma 8.6. For both Bidomain PP and PE formulations (A D APP or A D APE ),
the condition number of the block-diagonal preconditioned operator with MAS
scalar blocks (8.41), (8.42) is bounded by
1 hk1 1 C
.BD A/ c max 1C :
1kL1 ık 1
This result shows that in addition to the standard domain decomposition param-
eters, convergence with the block-diagonal preconditioner BD depends on the
parameter , which in turn depends only on the original Bidomain blocks. The
following estimates on 2 can be found in [381]:
(a) For the PP formulation, 2 .1 C min .Ae ; M//1 , where min .Ae ; M/ is
the minimum eigenvalue of Ae M1 (see [381, Lemma 4.3]) and numerical
experiments show that is very close to 1, so the bound of Lemma 8.6 might
be large and convergence slow;
(b) For the PE formulation, 2 .1 C min .Ae ; Ai //1 , where min .Ae ; Ai / is
the minimum eigenvalue of Ae A1 i (see [381, Lemma 4.1]) and numerical
experiments seem to indicate that is close to 1=2, so the bound of Lemma 8.6
is satisfactory.
The following convergence rate bound for the block-factorized preconditioner
can be found in [373].
Lemma 8.7. For both Bidomain PP and PE formulations (A D APP or A D APE ),
the extreme eigenvalues of the block-factorized preconditioned operator with MAS
244 8 Parallel Solvers for the Bidomain System
1
1 1 .Nc C 1/1
max .B A/ 1 :
F
1
These bounds are pessimistic because, due to Prop. 8.2, they require that 2 < ı1
and predict a large condition number when 2 approaches ı1 , while our numerical
results seem to indicate that the same considerations on for the block-diagonal
case also hold for the block-factorized case.
2
al .x/ D ue cos ˛.r/Cue sin ˛.r/; with ˛.r/ D .1r/ ; 0 r 1:
3 4
Conductivity coefficients. The values of the conductivity coefficients used in all
the numerical tests are the following:
ni D 3:1525 105 1 cm1 ne D 6:757 104 1 cm1 :
Mesh hierarchy. We denote the cartesian mesh used to discretize our domains
by T D Ti Tj Tk , indicating the number of elements in each coordinate direction.
This notation applies to both fine and coarse meshes. When we scale up the mesh by
a factor c, for brevity we define cT D cTi cTj cTk , i.e. the number of elements
in cT is c 3 times the number of elements in T . Figure 8.10 shows the domains
used in our parallel tests described below, reporting for each test the number of
processors (procs.), the coarse mesh (T0 ), the fine mesh (TL1 ), and the number of
degrees of freedom (d.o.f). Our tests will range from 1.8 million d.o.f. on 8 procs.
to 57.4 million d.o.f. on 256 procs.
Stimulation site, initial and boundary conditions. The depolarization process
is started by applying a stimulus of Iapp D 200 mA=cm3 lasting 1 ms on the face of
the domain modeling the endocardial surface. The initial conditions are at resting
values for all the potentials and LR1 gating variables, while the boundary conditions
are for insulated tissue. In all simulations, the fine mesh size is h D 0:01 cm. The
time step size is t D 0:05 ms.
246 8 Parallel Solvers for the Bidomain System
1e5 400
diagonal
350
diagonal
1e4
factorized 300
1e3 250
κ2
it.
200
1e2 150 factorized
100
10
50
coupled coupled
1 0
816 32 64 128 256 816 32 64 128 256
procs. procs.
Fig. 8.11 Test 1, weak scaling on ellipsoidal domains of Fig. 8.10 for PP formulation with
coupled, diagonal and factorized MHS(4) preconditioners. Top: table with condition numbers (),
extreme eigenvalues ( max ; min ), iteration counts (it.) for each preconditioner as a function of the
number of processors/subdomains (procs). Bottom: plots of condition numbers (left) and iteration
counts (right) from the table above
18
16 20 diagonal
diagonal
14
factorized
factorized 15
12
10
κ2
it.
8 10
4 5
2 coupled coupled
0 0
816 32 64 128 256 816 32 64 128 256
procs. procs.
Fig. 8.12 Test 2, weak scaling on ellipsoidal domains of Fig. 8.10 for PE formulation with
coupled, diagonal and factorized MHS(4) preconditioners. Top: table with condition numbers (),
extreme eigenvalues ( max ; min ), iteration counts (it.) for each preconditioner as a function of the
number of processors/subdomains (procs). Bottom: plots of condition numbers (left) and iteration
counts (right) from table above
In this strong scaling (standard speedup) test, we compare again the performance
of coupled, diagonal, factorized MHS preconditioners applied to both PP and PE
formulations. The number of levels is kept fixed to 4. The cardiac domain ˝ is the
portion of ellipsoid in Fig. 8.10, discretized by a fine mesh of 128 128 64 finite
elements, for a total amount of 2 163 330 dof. Since this is a strong scaling test,
the fine mesh is fixed, while the number of subdomains (D number of processors)
increases from 8 to 512. In this way, the subdomain size is reduced when the number
of subdomains is increased. The simulation is run for 0.5 ms during the excitation
phase, i.e. for 10 time steps.
Table 8.17 reports the average condition number, PCG iterations count and CPU
time per time step. The results show that the coupled preconditioners for both PP
and PE formulations have comparable performances and they are better than the
block preconditioners, especially for the PP formulation. The speedups, defined
with respect to the base 8 processors run, are good for the coupled preconditioners
up to 128 processors (up to 64 processors they are even superlinear for the PP
formulation), while they degenerate with 256 and 512 processors, because the local
problems become too small and the increasing communication costs deteriorate the
parallel performance. The block preconditioners have a much worse performance in
the PP formulation (as already seen in Test 1), showing reasonable speedups up to 64
processors but degenerating rapidly afterward, with even negative return (decreasing
speedups) for 512 processors. For the PE formulation the block preconditioners
show better timings, initially even better than the coupled preconditioner since
each iteration of the latter is more expensive, but such timings worsen rapidly
after 128 processors and become triple the coupled preconditioner timings for 512
248 8 Parallel Solvers for the Bidomain System
Table 8.17 Test 3, strong scaling test on ellipsoidal domain for PP and PE formulations with
coupled, diagonal and factorized MHS(4) preconditioners
MHS(4)-coupled MHS(4)-diagonal MHS(4)-factorized
Procs. it. Time it. Time it. Time
PP formulation
8 2.56 7 8.4 1.1e C 5 239 88.6 3.6e C 3 90 49.2
16 2.63 7 3.6 1.3e C 5 255 51.0 3.4e C 3 89 26.5
32 2.63 7 1.9 1.3e C 5 254 24.8 3.4e C 3 89 13.0
64 2.63 7 1.0 1.3e C 5 254 16.5 3.4e C 3 89 6.8
128 3.20 8 0.6 2.1e C 5 319 10.9 4.4e C 3 96 4.8
256 3.21 8 0.5 2.1e C 5 317 10.1 4.5e C 3 96 4.7
512 3.21 8 0.5 2.1e C 5 316 12.8 4.5e C 3 96 7.9
PE formulation
8 2.61 7 8.2 27.62 25 5.9 21.82 21 6.1
16 2.68 7 4.0 27.60 25 3.3 21.62 22 3.3
32 2.68 7 2.0 27.60 25 1.7 21.62 22 1.6
64 2.68 7 1.4 27.60 25 1.0 21.62 22 0.9
128 3.27 8 0.8 27.16 25 0.9 19.67 23 0.8
256 3.29 8 0.6 27.16 25 1.1 19.70 23 0.9
512 3.29 8 0.5 27.16 25 1.5 19.70 23 1.4
Condition numbers (), iteration counts (it.) for each preconditioner, CPU times (time) in seconds
as a function of the number of processors/subdomains (procs)
processors. This is also confirmed by the low speedups that degenerate rapidly and
start decreasing already at 256 processors.
In conclusions, these results show that block preconditioners for the Bidomain PP
formulation are not scalable, while they are scalable for the PE formulation, but with
higher iteration counts and computational costs than the coupled preconditioner.
Therefore, in our parallel tests (with idealized ventricular geometries and structured
finite elements) multilevel Schwarz preconditioners seem to be more efficient when
applied as coupled rather than block Bidomain preconditioners.
Chapter 9
Simulation Studies of Cardiac Bioelectrical
Activity
It is well known that the macroscopic electrical properties of the cardiac muscle
are markedly anisotropic. Experimental studies have confirmed that the Bidomain
representation of the cardiac tissue as superimposition of two continuous conductive
media, the intracellular and extracellular spaces, exhibits different anisotropy ratios
on the order of 2 and 10, with the intracellular domain the most anisotropic one; see
e.g. [100,277,420,421]. The development of optical mappings of cardiac transmem-
brane action potential, starting in the mid 1990s, has led to experimental studies by
several groups that have investigated the effects of unipolar extracellular cathodal or
anodal stimulations of cardiac tissue, see e.g. [156, 157, 281, 349, 480, 559]. These
studies have established that the stimulation by a unipolar electrode produces a
characteristic transmembrane pattern called virtual electrodes response.
After an anodal stimulus, the transmembrane potential distribution exhibits
a virtual anode (VA), i.e. negatively polarized (hyperpolarized) volume around
the stimulating electrode having a dog-bone shape, and by two virtual cathodes
(VCs), i.e. positively polarized (depolarized) regions adjacent to the concave part
of the hyperpolarized anodal dog-bone boundary, see e.g. Fig. 9.4. The central
dog-bone VA is aligned across the fiber direction, while the two adjacent VCs
are aligned along the fiber direction. Conversely, after a cathodal stimulus, the
polarity is reversed, i.e. the transmembrane potential pattern exhibits a central dog-
bone shaped VC aligned across fiber and two adjacent VAs aligned along fiber. It
is well known that only Bidomain models with unequal anisotropy ratios of the
intra- and extracellular media are able to generate the observed virtual electrode
polarization regions, see e.g. [469, 557]. Therefore, the experimental detection of
virtual electrode responses by [281, 349, 559] strongly supports the evidence for the
unequal anisotropy ratios assumption in Bidomain models; see the recent survey of
the Bidomain theory of pacing in [256].
The heart can be stimulated electrically in four different ways, see e.g. [146,207,
300, 301]:
• By turning on a negative current (cathode make, CM);
• By turning on a positive current (anode make, AM);
• By turning off of negative current (cathode break, CB);
• By turning off of positive current, (anode break, AB).
See Fig. 9.1 for a schematic illustration of these four cardiac stimulations.
In [427, 435], the Bidomain model with unequal anisotropic ratio was first
proposed and used to establish a theoretical framework able to explain the make
and break mechanisms of excitability in terms of the underlying virtual electrodes
polarization. The effects and features of make and break excitation mechanisms,
generated by unipolar extracellular cathodal or anodal stimulations, have been
subsequently investigated by simulation studies in [164,390,416,428–430,432,436,
450,451,486] and by experimental studies in [156,157,480,559]; see also the recent
surveys [256, 558] and the book [158].
9.1 Cardiac Excitation and Virtual Electrode Phenomena 251
applied current
excitation
excitation
time time
excitation excitation
applied current
applied current
time time
Fig. 9.1 Schematic illustration of the cathode make (CM), cathode break (CB), anode make (AM),
anode break (AB) excitation mechanisms
We briefly recall the main activation features of these mechanisms. With the
make stimulation, activation occurs with the stimulus onset either at the central VC
during cathodal stimulation (cathode make CM) or at the two VCs along the fiber
direction during anodal stimulation (anode make AM). With the break stimulation,
activation occurs after the stimulus termination at the two VAs during cathodal
stimulation (cathode break CB) or at the central VA during anodal stimulation
(anode break AB).
The break modes CB and AB generated by local electrode stimulation have been
confirmed by optical recordings and Bidomain simulations when a short stimulation
is delivered during the relative refractory period (systolic phase), whereas the make
modes CM and AM have been observed experimentally and simulated during the
diastolic phase. Many studies have underlined the important role played by the
anode break excitation in reentry induction mechanisms [429] and defibrillation
[11, 156, 157, 486, 525].
Here we focus on the 3D investigation of the excitation mechanisms associated
with an anodal stimulation, which can be generated only in tissues with unequal
anisotropy ratio. Additional simulations and results can be found in [131, 132].
252 9 Simulation Studies of Cardiac Bioelectrical Activity
with appropriate initial conditions on v.x; 0/; w.x; 0/ and c.x; 0/. Here cm and iion
denote the capacitance and the ionic current of the membrane per unit volume,
iRapp represents the applied current per unit volume with the compatibility condition
˝ iapp D 0, and b is the conductivity coefficient of the extracardiac medium.
e
i;e
where al .x/ is a unit vector parallel to the local fiber direction and l;t are the
effective intra and extracellular conductivity coefficients measured along (l) and
across (t) the fiber direction.
Computational domain. The cardiac domain ˝H considered in this study is
a cartesian slab of dimensions 0:96 0:96 0:32 cm3 , modeling a portion of the
left ventricular wall. The endocardial surface is in contact with a smaller slab of
dimensions 0:96 0:96 0:16 cm3 , modeling the extracardiac bath, where b D
6e 3 1 cm1 similar to the blood conductivity.
Stimulation site. Diastolic cathodal or anodal stimuli are delivered at the center
of the epicardial surface in a small region of dimensions 0:06 0:06 0:03 cm3
of a tissue at rest. In order to ensure the compatibility condition, we inject a total
stimulation current in a strip of dimensions 0:960:960:08 cm3 in the extracardiac
bath with equal strength and opposite polarity of that used at the subepicardial
level. Moreover, in order to reduce the effects of the delivered bath current on
the potential field in the subepicardial layer, we have lowered the excitability of
the subendocardial layer by strongly decreasing the maximal sodium conductance
during the stimulation interval. In order to simulate the break excitation mechanisms
during systole, we perform an S1-S2 stimulation protocol. We first apply an initial
S1 cathodal stimulus to the resting tissue and simulate the resulting action potential
propagation, that sweeps the cardiac domain until a suitable time during the relative
refractory period (RRP). At this instant, a subsequent S2 cathodal or anodal stimulus
is delivered and we simulate the resulting action potential propagation where the
surrounding tissue has recovered its excitability properties.
Activation time isochrones. During the Bidomain simulations, we process the
distribution of the transmembrane potential v.x; t/ in order to define the activation
time at.x/ as the first time instant for which v.x; at.x// D vup ; we choose vup D
50 mV, a value above the threshold value of v capable of initiating the upstroke
of the action potential. Then the excitation wavefronts are displayed by drawing the
isochrone surfaces of the activation time.
Numerical methods. In all computations, a structured grid of 96 96 48
hexahedral isoparametric Q1 finite elements of size h D 0:1 mm is used in space,
while the time discretization is based on the following double operator splitting
procedure, based on splitting both the ODEs from the PDEs and the elliptic PDEs
from the parabolic one, similar to the second decoupled semi-implicit method
254 9 Simulation Studies of Cardiac Bioelectrical Activity
described in Sect. 7.2.3, with the addition of the blood volume. Thus, at the general
time step t nC1 , given the transmembrane potential vn at the previous time step, we
find:
1. The gating variables wnC1 , by solving the related ODEs with the Backward Euler
method;
2. The ionic concentration variables c nC1 , by solving the related ODEs with the
Forward Euler method;
3. The extracellular and extracardiac potentials uenC1 and unC1 b , by solving the
algebraic linear system resulting from the finite element discretization of the
elliptic PDEs in (9.1);
4. The transmembrane potential vnC1 , by solving the algebraic linear system
resulting from the finite element discretization of the parabolic PDE in (9.1).
This operator splitting strategy yields two large linear systems of algebraic
equations that must be solved at each time step. In order to ensure parallelization and
portability of our Fortran code, we use the PETSc parallel library [24, 25], a suite of
data structures and functions for building large-scale parallel scientific applications,
based on the MPI communication library. The parallel strategy employed assigns
each subdomain to one processor and the information associated with the interior
of the subdomain is uniquely owned by that processor. The processor stores all
subvectors and a block of the matrices (mass, stiffness) associated to each over-
lapping subdomain. Our parallel code employs different DA (Distributed Arrays)
PETSc objects for representing v on the domain ˝H (tissue) and for unC1 e ; unC1
b on
the domain ˝ (tissue and bath). The two large linear systems at each time step are
solved by a parallel conjugate gradient method, preconditioned by the Multilevel
Hybrid Schwarz preconditioner, described in Chap. 8, for the ill-conditioned elliptic
system (461 041 d.o.f.) and the Block Jacobi preconditioner for the well conditioned
parabolic system (310 497 d.o.f.). These preconditioners are based on the multilevel
PETSc objects PCMG (MultiGrid) with ILU(0) local solvers.
Figures 9.2 and 9.3 show the epicardial distribution of the intracellular potential
(top), extracellular potential (middle) and transmembrane potentials (bottom) 2 ms
after the anodal stimulus elicited in a slab without and with transmural fiber rotation,
respectively. In each figure, the panels on the right column refer to a tissue with
the unequal anisotropy ratios of Table 9.1, while the panels on the left column
refer to a tissue slab with equal anisotropy ratio. In the left panel of Fig. 9.2, the
intra- and extracellular potential elliptical patterns have the same major to minor
semi-axis ratio, yielding a transmembrane potential pattern (bottom left panel) with
more rounded elliptical lines around the anodal electrode. Conversely, in a slab
with unequal anisotropy ratio, the intra- and extracellular equipotential contour
lines, displayed in top and middle right panels of Fig. 9.2, have elliptical shapes
9.1 Cardiac Excitation and Virtual Electrode Phenomena 255
20 20
0
0
−20
−20
−40
−40
−60
−80 −60
−100
−80
−120
−100
−140
−120
80
80
60
60
40
40
20
0 20
−20
0
−40
−20
−60
−120 −65
−70
−140
−75
−160
−80
−180
−85
−200 −90
−95
−220
−100
−232.52 −85.07 10.00 −276.21 −54.83 5.00
Fig. 9.2 Slab without transmural fiber rotation. Epicardial distribution of the intracellular potential
(top), extracellular potential (middle) and transmembrane potentials (bottom) 2 ms after an anodal
stimulus. Right panels: tissue with the unequal anisotropy ratios of Table 9.1. Left panels: tissue
with equal anisotropy ratio. Below each panel are reported the minimum, maximum and step of
the displayed map and the colorbar denotes the range of values of the displayed equipotential lines
256 9 Simulation Studies of Cardiac Bioelectrical Activity
20 20
0 0
−20 −20
−40
−40
−60
−60
−80
−80
25 25
20 20
15
15
10
10
5
5
0
0
−5
−10 −5
transmembrane potential
transmembrane potential
−86
−70
−88
−75
−90
−80
−92
−85
−94
−96 −90
−98 −95
−100
−232.22 −85.07 2.00 −214.59 −66.81 2.00
Fig. 9.3 Slab with transmural fiber rotation. Epicardial distribution of the intracellular potential
(top), extracellular potential (middle) and transmembrane potentials (bottom) 2 ms after an anodal
stimulus. Right panels: tissue with the unequal anisotropy ratios of Table 9.1. Left panels: tissue
with equal anisotropy ratio. Below each panel are reported the minimum, maximum and step of
the displayed map and the colorbar denotes the range of values of the displayed equipotential lines
9.1 Cardiac Excitation and Virtual Electrode Phenomena 257
with different ratio of the two semiaxis. The difference of these two epicardial
patterns yields a transmembrane potential distribution with the typical virtual
electrode response (Fig. 9.2, bottom right panel). In fact, the tissue within and
around the anodal electrode is negatively polarized (hyperpolarized) and it exhibits
an epicardial dog-bone shape, developing perpendicularly to the epicardial fiber
direction of 45ı . Two regions of positive polarization (depolarization), i.e. two
virtual cathodes, develop along the fiber direction adjacent to the concave parts of
the epicardial virtual anode boundary.
The same anodal stimulation was applied to a tissue slab with transmural fiber
rotation. In case of unequal anisotropy ratio, see right panels of Fig. 9.3, the
epicardial transmembrane potential pattern exhibits a virtual electrode response
with a twisted hyperpolarized dog-bone shaped region, which is not symmetric
with respect to the epicardial fiber direction because of the counterclockwise
fiber rotation from epicardium to endocardium. In fact, the equipotential surfaces
inside the virtual cathode volumes, shown in Fig. 9.4, are shaped as two horns
pointing counterclockwise when proceeding from the upper (epicardial) face to
the lower (endocardial) face. On intramural sections parallel to the epicardial
face, the equipotential lines inside the virtual anode preserve the same dog-bone
shape as on the epicardium but with a counterclockwise rotation, thus reducing
Table 9.1 Conductivity coefficients of the intra and extracellular media in S cm1 and anisotropy
ratios lti;e D li;e =ti;e , tn
i;e
D ti;e =ni;e
Unequal anisotropy
Medium li ti ni lti i
tn
Intra 2.31724e3 2.43504e4 5.69083e5 9.51622 4.27889
Medium le te ne lte e
tn
Extra 1.54483e3 1.04385e3 3.7221e4 1.47993 2.80447
Equal anisotropy
Medium li ti ni lti i
tn
Intra 1.34511e3 3.36278e4 8.40695e5 4 4
Medium le te ne lte e
tn
extra 5.35282e3 1.3382e3 3.34551e4 4 4
Fig. 9.4 Slab with transmural fiber rotation. Anodal stimulation applied to a slab with unequal
anisotropy ratio. Two different views of the isopotential surfaces of the transmembrane potential
distribution of values 88 and 75 mV, 2 ms after the beginning of the stimulation
258 9 Simulation Studies of Cardiac Bioelectrical Activity
their area and yielding a twisted tote bag; see Fig. 9.4. On the contrary, the same
anodal stimulus applied to a slab with equal anisotropy ratios and transmural fiber
rotation does not yield virtual cathodal regions but only an slightly twisted elliptical
hyperpolarized region centered at the stimulation site; see Fig. 9.3, bottom left panel.
This confirms that virtual electrodes can only be generated by bidomain models
with unequal anisotropy ratios of the intra- and extracellular media. The formation
of the two virtual cathodes elicited by the anodal stimulation explains why an anodal
stimulation is able to excite the myocardium, as shown by the excitation sequence
shown in Figs. 9.5 and 9.6 described below.
Anode make mechanism of excitability. We apply an anodal stimulus, lasting
10 ms and with an amplitude of 0:1112 mA, to the center of the epicardial face
of a tissue slab at rest with unequal anisotropy and with the full membrane
model. Figure 9.5 shows the snapshots of the simulated transmembrane potential
distributions on the epicardial surface and on the transmural diagonal sections.
The densely packed equipotential transmembrane lines indicate the location of the
excitation layers, propagating on the epicardial face and transmurally. The snapshots
in Fig. 9.5 reveal two distinct wavefronts propagating on the epicardial face and
transmurally, which subsequently merge at epicardial and intramural levels. In fact,
distinct activation wavefronts arise from two epicardial locations inside the virtual
cathodes and first propagate in all directions on the epicardial face. In particular,
they propagate outward and inward along the diagonal parallel to the fiber direction,
4 ms 6 ms 8 ms
50 50 50
0 0 0
4 ms 6 ms 8 ms
50 50 50
0 0 0
D1
50 50 50
0 0 0
D2
Fig. 9.5 Diastolic anode make excitation. Anodal stimulation during the diastolic interval with
total current Iapp D 0:1177 mA and 10 ms duration, applied to a slab with unequal anisotropic
calibration and funny current. Top panel. Transmembrane potential distributions on the epicardial
section at 3 time instants during the diastolic interval for cathode make stimulation. Bottom panel.
Transmembrane potential distributions on the two diagonal transmural sections of the slab d1 (top)
and d2 (bottom) at 3 time instants during the diastolic interval. Same conditions as in the top panel
(Reproduced with permission from [131])
9.1 Cardiac Excitation and Virtual Electrode Phenomena 259
a b
50 50 30
A1
C
2
A C 0 0 20
2 2
mV
ms
C1
−50 10
−50
C1 A1
A2
−100 0
−100 0 5 10 15 20
−401.13 −50.90 5.00
ms
c d
D1
20 20
18 15
16 10
14 5
10 D2
20
8
15
6
10
4
5
Fig. 9.6 Diastolic anode make excitation. Anodal stimulation during the diastolic interval with
total current Iapp D 0:1177 mA and 10 ms duration, applied to a slab with unequal anisotropic
calibration and funny current. (a) Transmembrane potential distribution on the epicardial surface
0:5 ms after the beginning of the stimulation. (b)-left. Transmembrane potential waveforms in
points C1 (blue continuous), A1 (red continuous), A2 (red dashed) and C2 (blue dashed), indicated
in figure (a). (b)-right. Activation time profiles on the diagonal across (blue) and along (red) fibers.
The black line indicates the instant (10 ms), when the stimulus ends. (c). Epicardial isochrones of
activation time. (d). Isochrones of activation time on the two transmural diagonals. Same format
as in Fig. 9.6 (Reproduced with permission from [131])
but, when the excitation isochrones reach points lying on a boundary of the virtual
anode, a block of the inward propagation takes place during the stimulation interval,
as also shown by the large slope of the activation time of Fig. 9.6b inside the anodal
area. Therefore, from about 2 to 6 ms, two distinct activation wavefronts propagate
outward, moving faster along fibers. At about 12 ms, the central epicardial area of
the virtual anode is suddenly activated (0.5 ms), since it becomes excitable only
after the stimulus turn off. Subsequently, all fronts finally merge in a unique large
wavefront.
260 9 Simulation Studies of Cardiac Bioelectrical Activity
Excitation starts at about 2 ms, as shown by the activation time plot along
the diagonal parallel to the epicardial fiber direction, displayed in Fig. 9.6b, thus
inducing an anode make excitation. The isochrones of activation time on the
epicardial surface and transmural diagonal sections (Fig. 9.6c, d) show clearly
that the first triggered areas are the two epicardial sites located at the center
of the virtual cathodes, generating two distinct excitation wavefronts propagating
intramurally and separated by a central region within the virtual anode volume,
which is inexcitable until the stimulus is turned off. These two wavefronts are open
surfaces, each having a closed curve as rim lying on the epicardial surface. There
are epicardial and intramural excitation pathways that, starting from the virtual
cathodes, proceed turning around the boundary of the virtual anode volume and
point toward its epicardial center. These pairs of pathways, coming from the virtual
cathodes, cause the merging of the two wavefronts forming an activated tissue
volume bounded by two intramural surfaces, one consisting of a wavefront moving
toward the epicardial boundary and intramurally toward the endocardial face, and
the other being the conduction block surface surrounding the inexcitable region,
which is suddenly activated after the stimulus is turned off. From the inspection
of the transmural isochrones pattern displayed in Fig. 9.6d, we observe that the
major transmural effect of the virtual electrode polarization is the propagation of
two wavefronts, separated by the transmural part of the virtual anode region, that
merge 5 ms after the beginning of epicardial excitation at 1 mm depth.
Cathode make mechanism of excitability. We now consider the application of
a cathodal stimulus, lasting 1 ms and with an amplitude of 0:0378 mA, to a slab
with unequal anisotropy and full membrane model. Figure 9.7 shows the snapshots
of the transmembrane potential distribution on the epicardial surface and on the two
2 ms 4 ms 6 ms
50 50 50
0 0 0
2 ms 4 ms 6 ms
50 50 50
0 0 0
D1
50 50 50
0 0 0
D2
Fig. 9.7 Diastolic cathode make excitation. Cathodal stimulation with total current Iapp D
0:0378 mA and 1 ms duration, applied to a slab with unequal anisotropic calibration and funny
current. Same format as in Fig. 9.5 (Reproduced with permission from [131])
9.1 Cardiac Excitation and Virtual Electrode Phenomena 261
diagonal transmural sections (bottom panel) at instants subsequent to the end of the
stimulus pulse. The densely packed equipotential transmembrane lines indicates the
location of the excitation layer, propagating on the epicardial face and transmurally.
This excitation wavefront has an approximate ellipsoidal shape with the major
axis and one of the two minor axes parallel and perpendicular, respectively, to the
epicardial fiber direction of 45ı ; the remaining minor axis develops intramurally
through the wall thickness and transversally to fiber layers.
Figure 9.8a illustrates the characteristic pattern of the epicardial transmembrane
potential at 0:5 ms (at the beginning of the stimulation interval), with a dog-bone
shaped virtual cathode, located around the site of the stimulation electrode, and
two adjacent virtual anode areas. A cathode make excitation mechanism occurs,
because activation arises before the end of stimulation as confirmed by the activation
time plots at points A1 ; A2 ; C1 ; C2 along the two epicardial diagonals, displayed in
Fig. 9.8b.
a b
−80 50 30
−81 A
1
A1 C2 −82 0 C 20
1
mV
ms
−83
A2
−50 10
−84 C2
C1 A2
−85
−100 0
−86 0 5 10
−85.83 −66.32 0.20 ms
c d
D1
18
16 15
14 10
5
12
Fig. 9.8 Diastolic cathode make excitation. Cathodal stimulation during the diastolic interval
with total current Iapp D 0:0378 mA and 1 ms duration, applied to a slab with unequal anisotropic
calibration and funny current. Same format as in Fig. 9.6 (Reproduced with permission from [131])
262 9 Simulation Studies of Cardiac Bioelectrical Activity
In order to describe in detail the origin, development and shape of the excitation
wavefronts, we plot in Fig. 9.8c, d the activation time isochrones on the epicardial
face and on two transmural diagonal sections. Excitation originates at locations
surrounding the physical cathode and propagates outward in all epicardial and
intramural directions, moving faster along fibers and assuming an approximate
semi-ellipsoidal shape. The epicardial velocities of propagation along and across
fibers are about 0:060 and 0:035 cm/ms, respectively, for unequal anisotropy, as also
confirmed by the activation time slope in Fig. 9.8b.
In order to study the anodal and cathodal virtual electrode phenomena during
systole, an S1 S 2 stimulation protocol has been considered. The cathodal S1
stimulus is applied at the center of the epicardial surface and the entire activation-
repolarization process is simulated. At 380 ms after the S1 stimulus, during the
relative refractory period (RRP) of the repolarization phase, a cathodal or anodal
S 2 stimulus is applied at the same site of the S1 stimulus pulse. In the following,
we denote by T 1 anisotropy the anisotropic calibration of Table 9.1, and by T 2
anisotropy a calibration with a 20 % reduction of intracellular anisotropy ratios, see
also [131].
Anode break mechanism of excitability. We consider an S 2 anodal stimulus of
amplitude 0:2916 mA and duration 10 ms, in a slab with T 1 anisotropy with the full
membrane model. Figure 9.9 shows the snapshots of the transmembrane potential
10 ms 18 ms 22 ms
50 50 50
0 0 0
10 ms 18 ms 22 ms
50 50 50
0 0 0
D1
Fig. 9.9 Systolic anode break excitation. Anodal stimulation during the systolic interval (S1-S2
protocol, S2 at 380 ms) with total current Iapp D 0:2916 mA and 10 ms duration applied to a slab
with unequal anisotropy and membrane model with the addition of If ; Ie ; Ia . Same format as in
Fig. 9.5 (Reproduced with permission from [131])
9.1 Cardiac Excitation and Virtual Electrode Phenomena 263
distribution at 10, 18, 22 ms after the end of the S 2 anodal pulse. During the 10 ms
of stimulation, a large dog-bone shaped region of strongly hyperpolarized tissue
(virtual anode) is generated, while two virtual cathodes develop along the fiber
direction near the concave portion of the anodal region, see Fig. 9.10. Although
the virtual cathode volumes are depolarized above threshold (above 20 mV in
Fig. 9.10a), anode make stimulation does not occur, because the surrounding regions
are in the refractory period and still inexcitable, since the sodium inactivation gates
are closed. When the stimulus is turned off, the combined effects of the membrane
funny current If , of the electroporation current Ie and of the outward current Ia ,
depolarize the tissue inside the virtual anode, inducing anode break excitation with
a delay of about 2 ms. The densely packed equipotential lines in the snapshots of
Fig. 9.9 reveal the development of two epicardial excitation wavefronts propagating
mainly in the direction across the epicardial fibers. The intramural diagonal sections
exhibit a strongly different shape of the excitation layers, that originate from
a b 50 30
A1
0 C
1
C
−20 0 A 2 20
A2 C 2
2
mV
ms
−40
C1 A −50 10
1
−60
−80
−100 0
0 10 20 30
−100
ms
−389.42 17.26 5.00
c d D1
30
30 20
10
25
20
11.89 34.75 1.00
15 D2
10 30
20
5
10
Fig. 9.10 Systolic anode break excitation. Anodal stimulation during the systolic interval (S1-
S2 protocol, S2 at 380 ms) with total current Iapp D 0:2916 mA and 10 ms duration applied to a
slab with unequal anisotropy and membrane model with the addition of If ; Ie ; Ia . Same format as
in Fig. 9.6 (Reproduced with permission from [131])
264 9 Simulation Studies of Cardiac Bioelectrical Activity
the center of the virtual anode and then reach and collide on the epicardial and
intramural boundary of the virtual cathode volumes, which are inexcitable. Then
the wavefront rims move around these obstacles until the activated tissue fully
surrounds them.
In order to describe in detail the origin, development and shape of the excitation
wavefronts, the isochrone lines drawn in Fig. 9.10c, d show clearly that the first
activated point is the epicardial central site of the anodal region. Moreover, there
are intramural excitation pathways that, starting from the epicardial anodal areas,
first point toward the endocardium, progressively bend and proceed around the
intramural boundary of the virtual cathodes, and finally point toward the epicardial
face where a wavefront collision happens.
The initial phase of this anode break mechanism is due to the interplay between
the If ; Ie currents, yielding a transmembrane potential above the threshold
excitability inside the virtual anode region, and the sodium channels dynamics.
Therefore, we have a membrane-based trigger excitation mechanism. Apart from
the different triggering process, the excitation wavefront pattern is strongly similar
to the activation sequence generated by a cathodal pulse applied at the center of the
epicardial surface in the presence of two obstacles (i.e. non conducting volumes),
located inside the virtual cathode regions, which are inexcitable.
Figure 9.11e reports the transmembrane potential distributions after the end of an
S 2 anodal stimulus of 0:1512 mA amplitude, lower than in the previous case, and
10 ms duration. After the pulse turns off, the transmembrane potential in the virtual
cathode regions decays toward the resting value, while that in the virtual anode
area, due to the effect of the membrane funny current, depolarizes with a delay of
about 9 ms from the end of the stimulus pulse. The propagating wavefront assumes
an approximate semi-elliptical shape, as in the diastolic anode break excitation,
because the virtual cathodes at the end of the stimulation interval are excitable
again and do not behave like obstacles. The transition between the two systolic
break excitation patterns happens smoothly when decreasing the strength of the
anodal impulse. This is confirmed by the patterns displayed in Fig. 9.11, showing
that decreasing the pulse amplitude the cathodal volumes become re-excitable and
a dimple like inflection appears in the epicardial isochrones associated with slow
propagation along the fiber direction (panel c). Decreasing further the intensity of
the anodal stimulus allows an almost full re-excitability of the virtual cathodes after
the stimulus end, yielding approximate semi-ellipsoidal propagating wavefronts
with flat isochrone lines in the direction along fibers (panel e). Below the threshold
of 0:1512 mA, anode break excitation does not occur.
The systolic anode break mechanism of excitation has been also simulated for a
slab with T 2 anisotropy and a strongly similar excitation sequences, depending on
the amplitude of the stimulus pulse, was observed (not shown). The velocities of
propagation along (across) fibers are 0:042 cm/ms (0:031) for T 1 anisotropy and
0:040 (0:030) for T 2. The degree of anisotropy influences weakly the stimulus
threshold of the anode break excitation, which is slightly higher for T 2 anisotropy
(0:1809 mA) than for T 1 anisotropy (0:1512 mA).
9.1 Cardiac Excitation and Virtual Electrode Phenomena 265
a b
50 30
35
30 A C
1 2
25 0 20
20
mV
ms
A C
2 1
15
−50 10
10
5
−100 0
0.19 35.95 1.00 0 10 20 30
ms
c d
50 30
35
C
2
30 0 A1 20
C
1
mV
ms
25 A2
−50 10
20
−100 0
15 0 10 20 30
14.79 37.72 1.00
ms
e f 50 30
C2
40
0 C 20
35 A 1
1
mV
ms
30
−50 10
A2
25
20 −100 0
0 10 20 30
18.78 43.92 1.00
ms
Fig. 9.11 Systolic anode break excitation. Anodal stimulations during the systolic interval (S1-
S2 protocol, S2 at 380 ms) with total current Iapp D 0:2160 mA (first row), Iapp D 0:1728 mA
(second row), Iapp D 0:1512 mA (third row), and 10 ms duration applied to a slab with
unequal anisotropy and membrane model with the addition of If ; Ie ; Ia . Panels (a-c-e): Epicardial
isochrones of activation time. Panels (b-d-f)-left: Transmembrane potential waveforms in points
C1 (blue continuous), A1 (red continuous), A2 (red dashed) and C2 (blue dashed), indicated in
Fig. 9.8a. Panel (b-d-f)-right: Activation time profiles on the diagonal across (blue) and along
(red) fibers. The black line indicates the instant (10 ms) when the stimulus ends (Reproduced with
permission from [131])
266 9 Simulation Studies of Cardiac Bioelectrical Activity
10 ms 14 ms 18 ms
50 50 50
0 0 0
10 ms 14 ms 18 ms
50 50 50
0 0 0
D1
50 50 50
0 0 0
D2
Fig. 9.12 Systolic cathode break excitation. Cathodal stimulation during the systolic interval
(S1-S2 protocol, S2 at 380 ms) with total current Iapp D 0:2160 mA and 10 ms duration, applied
to a slab with unequal anisotropy and membrane model with the addition of If ; Ie ; Ia . Same format
as in Fig. 9.5 (Reproduced with permission from [131])
9.1 Cardiac Excitation and Virtual Electrode Phenomena 267
b 50 30
A1
A
a 50
2
0 20
C1
mV
ms
A C2 C
2 0 2
−50 10
C A
1 1
−50
−100 0
0 10 20 30
−100 ms
−118.42 185.88 5.00
d D1
c
35 30
20
30
10
25
0.05 35.08 1.00
20
15 D2
30
10
20
5
10
Fig. 9.13 Systolic cathode break excitation. Cathodal stimulation during the systolic interval
(S1-S2 protocol, S2 at 380 ms) with total current Iapp D 0:2160 mA and 10 ms duration, applied
to a slab with unequal anisotropy and membrane model with the addition of If ; Ie ; Ia . Same format
as in Fig. 9.6 (Reproduced with permission from [131])
From the isochrone lines displayed in Fig. 9.13c, we clearly see that the excitation
sequence between 10 and 21 ms consists of two separated epicardial isochrones,
being the rims of two distinct wavefronts, which collide at 12 ms. As a major
transmural effect of the virtual electrode polarization, Fig. 9.13d shows that each
of the two propagating open wavefront surfaces has a rim partly coinciding with
a curve lying on the transmural boundary surrounding the virtual cathode volume.
The transmural wavefronts merge at about 25 ms in the midwall, then, after collision,
we have a unique wavefront propagating around the virtual cathode volume. Apart
from the virtual cathode volume, the activation patterns associated to the diastolic
anode make and to the systolic cathode break excitation mechanisms are similar,
especially after the merging of the separated wavefronts, as shown by comparing
the isochrones displayed in Figs. 9.13 and 9.6. Despite this qualitative similarity,
Fig. 9.13 shows that, for cathode break excitation, the first activated sites are not
268 9 Simulation Studies of Cardiac Bioelectrical Activity
located in the epicardial central sites in the virtual anodes, but at the two concave
portions of the boundary surrounding the virtual cathode volume. The trigger of
the systolic cathode break mechanism is due to the interplay between the outward
electrotonic current from the cathodal region and the membrane current dynamics at
the locations around the virtual anode. The wavefront propagation coincides with an
activation sequence moving around an obstacle (i.e. unexcitable volume) associated
with the depolarized region under the cathode.
Cathode break excitability mechanism is simulated also for a slab with T 2
anisotropy and exhibits a strongly similar excitation sequence (not shown). The
epicardial velocities of propagation along (across) fibers are 0:049 cm/ms (0:030)
for T 1 anisotropy and 0:040 (0:030) for T 2. These values are lower than the corre-
sponding diastolic velocities, because the tissue during the RRP is less excitable than
at rest. The degree of anisotropy influences the stimulus threshold of the cathode
break excitation, which is higher for the weaker T 2 anisotropy (0:1075 mA) than
for the stronger T 1 anisotropy (0:0578 mA). Finally, the cathode break mechanism
occurs, as expected, independently of the membrane funny current incorporation
(not shown).
Cathodal S−I curve, strong electroporation Anodal S−I curve, strong electroporation
0.5 0.5
0.4 0.4
stimulation threshold (mA)
0.2 0.2
0.1 0.1
0 0
180 190 200 210 220 230 240 250 180 190 200 210 220 230 240 250
S2 stimulation onset (ms) S2 stimulation onset (ms)
Cathodal S−I curve, weak electroporation Anodal S−I curve, weak electroporation
0.5 0.5
0.4 0.4
stimulation threshold (mA)
0.3 0.3
0.2 0.2
0.1 0.1
0 0
180 190 200 210 220 230 240 250 180 190 200 210 220 230 240 250
S2 stimulation onset (ms) S2 stimulation onset (ms)
Fig. 9.14 Cathodal (left) and anodal (right) strength-interval curves for strong (top) and weak
(bottom) electroporation. Bold points and empty circles denote make and break stimulation
responses, respectively (Reproduced with permission from [132])
stimulation responses the two type of responses. When computing the anodal S-I
curves described above, we refine the step from 5 to 1.25 ms around the break/make
transition in order to better resolve the S-I curve in this interval.
S-I curves. Both cathodal S-I curves display a monotonically decreasing behav-
ior in the portions corresponding to the cathode break stimulation response and the
subsequent phase corresponding to the cathode make stimulation response, with an
almost constant threshold stimulus strength. The use of strong electroporation has
a weak effect on the make portion of the S-I curve, while on the break portions
it induces a slight decrease of the threshold stimulus strength with a shift of the
effective refractory period which is increased from 180 to about 190 ms.
The anodal S-I curves displayed in Fig. 9.14 exhibit the following main features:
a dip close to the effective refractory period, a subsequent plateau or dome phase
corresponding to the anode break stimulation response and a sudden decrease close
to the end of the relative refractory period, separating the anode break and make
mechanisms. By relative refractory period, we mean the time interval, between the
effective refractory period and the resting phase, where another action potential can
270 9 Simulation Studies of Cardiac Bioelectrical Activity
will identify the upper and lower face planes with the epicardial and endocardial
surfaces, respectively. The minimal velocity D 10 cm s1 corresponds to a
propagation across fibers on intramural planes parallel to the epicardial surface,
while the velocity t D 25 cm s1 corresponds to a propagation across fibers in
directions perpendicular to the epicardial plane. The initial conditions are .u0i ; u0e / D
.84; 0/ mV, so that v0 D 84 mV and we apply a large stimulus of 200 A for
1 ms on a small volume (3 or 5 mesh points in each direction). The slow inward
current Isi in the LR1 model is scaled by a factor 2/3. This yields an action potential
duration (APD) of about 265 ms.
The fibers rotate intramurally linearly with depth for a total amount of 90ı , i.e.
al .x/ D uex cos ˛.r/ C uey sin ˛.r/; ˛.r/ D .1 2r/=4; r 2 Œ0; 1. The fiber
direction on the epicardium is 45ı with respect to the horizontal side of the panels
displayed in Figs. 9.15–9.17, while on the endocardium is 45ı . In the orthotropic
anisotropy case, we choose a D uex sin ˛.r/ uey cos ˛.r/ and at D uez . In order
to strengthen the role played by the rotational anisotropy, the counter-clockwise
(CCW) intramural fiber rotation amounts to 90ı on a slab thickness of 0.5 cm,
with direction varying linearly from epicardium to endocardium. The stimulus was
applied at an appropriate epicardial vertex or at the center of the slab. In all cases,
the finite element mesh size is h D 0:1 mm.
In order to study the influence of fiber rotational anisotropy on the activation and
repolarization sequences and on the APD distribution, we simulate the electrical
activity in a slab of cardiac tissue, from the onset of excitation to the end of
repolarization. We consider the following test problems:
• Bidomain-LR1 model on a 2 2 0:5 cm3 slab, vertex stimulation (orthotropic
(left Panel) and axisymmetric (right Panel) in Fig. 9.15 top);
• Monodomain-LR1 model on a 220:5 cm3 slab, vertex stimulation (orthotropic
(left Panel) and axisymmetric (right Panel) in Fig. 9.15 bottom);
• Monodomain-LR1 model on a 4 4 0:5 cm3 slab (central stimulation,
axisymmetric in Fig. 9.16, orthotropic in Fig. 9.17).
272 9 Simulation Studies of Cardiac Bioelectrical Activity
EPI
40
40 266 300
300 265
264 20
20
280 262 280 260
0.09 80.44 2.50 271.52 341.02 2.50 260.59 271.79 0.30 0.10 72.15 2.50 274.74 331.63 2.50 259.49 274.65 0.30
268 268
266 266
MID
MID
264 264
262 262
9.73 81.49 2.50 279.15 343.15 2.50 261.05 269.41 0.30 10.91 64.27 2.50 280.09 327.25 2.50 260.98 269.17 0.30
268
266
ENDO
ENDO
266
264
264
262
262
20.63 86.94 2.50 285.84 347.83 2.50 260.89 268.87 0.30 21.41 64.62 2.50 285.32 324.81 2.50 260.18 267.38 0.30
270
268 270
266
265
264
262
260
0.09 86.94 2.50 271.52 347.83 2.50 260.59 271.79 0.30 0.10 72.15 2.50 274.74 331.63 2.50 259.49 274.68 0.30
320 266
40 300 265
300 264 20
20 262
280 280 260
0.12 87.99 2.50 271.82 348.36 2.50 260.37 271.97 0.30 0.12 76.07 2.50 275.28 335.49 2.50 259.43 275.17 0.30
268
268
266
MID
266
MID
264
264
262
262
9.89 90.52 2.50 279.27 354.16 2.50 261.15 269.38 0.30 11.15 68.33 2.50 280.50 331.06 2.50 260.83 269.35 0.30
268 266
ENDO
266
ENDO
264
264
262
262
21.67 68.03 2.50 285.77 328.35 2.50 260.32 267.56 0.30
20.81 99.38 2.50 286.06 359.62 2.50 260.25 268.79 0.30
275
270
270
268
266 265
264
262 260
Fig. 9.15 Orthotropic (left Panels) and axisymmetric (right Panels) Bidomain Models (top row)
and Monodomain Models (bottom row) on a slab 2 2 0:5 cm3 with vertex stimulation and
homogeneous cellular membrane properties. Patterns of level lines of the depolarization time (first
column ACTI), repolarization time (second column REPO) and action potential duration (third
column APD) on epi, midwall and endocardial planes and at the bottom 3D view. Reported below
each panel are the maximum, minimum and step in ms of the displayed map (Reproduced with
permission from [123])
268
266
MID
264
262
11.70 68.66 2.50 280.69 331.30 2.50 260.78 268.99 0.30
266
ENDO
264
262
260
22.03 64.13 2.50 285.63 326.92 2.50 259.50 267.42 0.30
275
270
265
260
0.14 76.43 2.50 276.20 335.81 2.50 259.38 276.06 0.30
Fig. 9.16 Axisymmetric monodomain model with intramural fiber rotation on a slab 4 4
0:5 cm3 with central epicardial stimulation. Same format as in Fig. 9.15 (Reproduced with
permission from [123])
The macroscopic sequences of excitation and recovery and the spatial distribu-
tion of the APDs are illustrated in Fig. 9.15 top, that relates to a Bidomain model
implemented in a slab with orthotropic and axisymmetric anisotropy, respectively.
Both anisotropic slabs exhibit similar qualitative features of the patterns of excita-
tion, recovery and APD distribution.
Activation sequence. It is well known that in the presence of rotational
anisotropy, intramural excitation, starting from an epicardial stimulation site, first
proceeds toward the endocardium but subsequently, due to fiber rotation, comes
back pointing toward the epicardial plane (see e.g. [67,102,116,184,506,508,509]).
Due to these intramural return pathways, propagation undergoes an acceleration,
in particular in epicardial areas where the excitation moves mainly across fibers.
In the orthotropic slab, the lower epicardial velocity across fiber (as compared
to axisymmetric model) causes the appearance of more pronounced dimple-like
inflections.
Repolarization sequence. Repolarization wave fronts exhibit a somewhat
smoother shape and faster propagation compared with the excitation sequence, as
shown by the isochronal lines on the epi, midwall and endocardial planes. Indeed,
the epicardial plane is fully activated in about 80.35 ms and repolarizes in 69.50 ms,
274 9 Simulation Studies of Cardiac Bioelectrical Activity
268
266
MID
264
262
11.02 75.68 2.50 280.00 339.82 2.50 261.04 268.98 0.30
268
266
ENDO
264
262
Fig. 9.17 Orthotropic monodomain model with intramural fiber rotation on a slab 4 4 0:5 cm3
with central epicardial stimulation. Same format as in Fig. 9.15 (Reproduced with permission from
[123])
giving rise to a total amount of APD dispersion of 11.2 ms in the orthotropic case. In
the axisymmetric case, the same section is fully activated and repolarized in about
72 and 57 ms, respectively, with an APD dispersion of 15 ms.
APD distribution. The main common features, displayed by the APD on the
epicardial, intramural and endocardial planes, in both the anisotropic slabs are: a
maximum located at the epicardial stimulation site, higher APD values in regions
where propagation is mainly along fibers, an intramural valley configuration of APD
reduction, a relative minimum in areas of front-boundary collision, and a maximum
located on the endocardial plane. A detailed comparison between orthotropic and
axisymmetric cases shows that the former emphasizes some anisotropic features
displayed by the APD distribution pattern. In particular, the presence of a dimple-
like inflection exhibited by the excitation isochronal lines on intramural horizontal
planes ranging from the epicardium to the midwall, and of a corresponding finger-
shaped valley of low APD are more pronounced in the orthotropic slab, as shown
e.g. by comparing the midwall panels of Fig. 9.15 top.
We remark that the axisymmetric and the orthotropic slabs are fully excited
in about 72 and 87 ms, respectively. The extinction area of excitation is located
at a vertex on the endocardium in the orthotropic slab, whereas it appears at the
9.2 Anisotropic Propagation of Excitation and Recovery Fronts 275
epicardium in the axisymmetric slab. This is the result of the reduced velocity
t across fibers on planes parallel to epicardium being lower than the transmural
velocity across the slab thickness, combined with the effects of the intramural
return pathways of the excitation toward the epicardium. The main quantitative
discrepancy between the two slabs lies at the epicardium, where different activation
and repolarization isochrone shapes in the front-boundary collisions are present;
they are associated with the appearance, in the APD pattern of the orthotropic
slab, of an epicardial relative maximum located at a side of the finger-shaped
valley of APD reduction which started from a relative minimum located at the
boundary.
The APD distributions show several common features in our different simu-
lations, namely: APDs are longest near the pacing site and tend to be shorter
near the front-boundary collisions. Moreover, the spatial distribution of APDs
are influenced by the anisotropic motion of the wave front. Thus, in a slab with
homogeneous cellular membrane properties, i.e. where all individual cells have the
same intrinsic transmembrane action potential, the rotational anisotropy produces
an APD variability displaying an anisotropic pattern. The orthotropic anisotropy
further increases the distortion of the propagating wavefront, emphasizing the
corresponding anisotropic features of the APD dispersion of the axisymmetric case.
Due to the high computational cost of the Bidomain model on a slab with
dimensions 2 2 0:5 cm3 , we performed simulations including both excitation
and recovery using the Monodomain model on the same slab with orthotropic and
axisymmetric anisotropy. We focus the comparison on the shape and propagation
of the excitation and recovery isochrones, as well as the spatial distribution of the
APD.
Activation and repolarization sequences. The comparison of Fig. 9.15 top and
bottom for both orthotropic and axisymmetric slabs evidences a strong qualitative
similarity between the simulated isochrones of the Bidomain and Monodomain
models. This fairly good P match is confirmed by the values of the P correlation
coefficient CC.x; y/ D n1 . .xi x/.yi y//=.x y /; where x D n1 xi is the
q P
arithmetic mean of x, x D 1
n
.xi x/ is the standard deviation of x and
2
analogously for y and y . The correlation coefficient CC for the activation times
of the two models in the entire slab amounts to 0.9971 in the orthotropic case
and to 0.9997 in the axisymmetric case. Similarly, the CC for the repolarization
times on the entire slab amounts to 0.99976 and to 0.9998 in the orthotropic
and axisymmetric cases, respectively. The wave front velocity depends on the
propagating direction and the Bidomain and Monodomain models exhibit a different
anisotropic dependence. In spite of this difference, we found not only a strong
276 9 Simulation Studies of Cardiac Bioelectrical Activity
qualitative agreement but also a small quantitative difference in the activation and
repolarization sequences between the Monodomain and the Bidomain model. The
relative error RE D kTB TM k2 =kTB k2 (with k k2 the Euclidean norm) between
the activation times TB ; TM of the Bidomain and Monodomain models is 0.115 in
the orthotropic case and 0.063 in the axisymmetric case. The relative error between
the repolarization times of the two models is 0.018 in the orthotropic case and 0.009
in the axisymmetric case. We remark that the repolarization isochrones simulated
by both models, especially on the epicardial plane, exhibit a smoother dimple-like
inflection than that appearing in the excitation sequence.
APD distribution. Both orthotropic model simulations exhibit a common
narrow valley of low APD values (see Fig. 9.15 top and bottom), corresponding
to the pathway joining the dimple-like inflections appearing in the epicardial and
intramural excitation isochrones. These dimple-like inflections are a reflection of
the intramural pathways of excitation that, starting from the stimulation site, first
proceed toward the endocardium but subsequently, due to fiber rotation, return
toward the epicardial plane. Therefore, the narrow valley configuration in the APD
distribution is an effect of the rotational anisotropy. We found a good qualitative and
quantitative agreement between the APD related to the Bidomain and Monodomain
models, since CC and RE on the entire slab are 0.9534 and 1.882 103 in the
orthotropic case, while they are 0.987 and 8.9 104 in the axisymmetric case.
The previous comparison validates the use of the Monodomain model as a
tool for investigating the qualitative macroscopic features of the activation and
repolarization sequences, as well as the APD distribution. Therefore, the high costs
required, at present, by the Bidomain simulations can be avoided by implementing
the Monodomain model, which qualitatively preserves the anisotropic features of
the APD distribution.
To further analyze the factors that modulate the APD in a model with homogeneous
membrane properties, we simulate the excitation and recovery sequences elicited
by a central epicardial stimulation, in case of axisymmetric and orthotropic Mon-
odomain slabs with intramural fiber rotation (Figs. 9.16 and 9.17).
The comparison between the APD dispersion patterns should allow us to detect the
role played by:
• The different phases of the wave front sequence of excitation and recovery
(i.e. the initial and extinction wave fronts, the occurrence of front-boundary
collisions) and the wave front shape;
• The anisotropic electrical conductivity (i.e. the different electrotonic currents
flowing along and across the fiber direction);
• The intramural fiber rotation, which causes the occurrence of return pathways of
excitation and recovery toward the stimulated plane.
9.2 Anisotropic Propagation of Excitation and Recovery Fronts 277
Since similar observations apply to both cases, in the results description we refer,
preferentially, to the orthotropic case in which some features are more pronounced.
Activation and recovery sequences. Stimulation at the center of the epicardial
face produces approximately elliptical excitation and repolarization isochrone lines,
a clear sign of their anisotropic propagation. Actually, activation and repolarization
spread symmetrically with respect to the center of the epicardial face. The major
axes of the oblong excitation isochrones are nearly parallel to the epicardial fiber
direction (45ı with respect to the horizontal sides of the panels in Figs. 9.16
and 9.17) while those associated to repolarization isochrones are rotated CCW with
respect to epicardial fiber direction. The oblong epicardial isochrones are slightly
rotated and bulging CCW with respect to the fiber direction, see Figs. 9.16 and 9.17
Panels ACTI, REPO. The spread of excitation and the sequence of recovery show
an acceleration in the cross-fiber direction. Also, the excitation isochrones show
an inflection corresponding to a dimple-like inflection of the wave front. These
findings, i.e. the accelerating propagation across fibers, the bulging and the dimple-
like inflection of the isochrones, are attributed to the influence on the epicardium
of the activation and recovery processes through the deeper layers, where the fiber
direction rotates CCW relative to epicardial fibers.
Proceeding from epi- to endo-cardium, on the intramural planes parallel to the
epicardium the spacing between excitation (recovery) isochrones increases, the
wave front shapes become rounder and we observe a transmural twisting of the
isochrones, i.e. the major axis of the oblong isochrones progressively rotates CCW
with increasing depth. However, their rotation lags behind the rotation of the fiber
direction at corresponding depths (see e.g. the Panels MID of Fig. 9.17, where
the fiber direction is horizontal). It is worth mentioning that the total transmural
rotation of the recovery isochrones is slightly less than the rotation of the excitation
isochrones. The spacing between the recovery isochrones, in the sections parallel to
the epicardial face, is greater than the corresponding excitation spacing, denoting a
faster motion of the recovery fronts.
On the endocardial plane, both the excitation and recovery front-boundary collision
first occur at the center of the face. This happens because our current model does not
incorporate the epi- endocardial obliqueness of the fibers (Streeter [498], Taccardi
et al. [507]), that was included in some of our previous simulations (Colli Franzone
et al. [116, 116]). Subsequent excitation and recovery isochrones have a well
rounded, elliptical shape centered at the point of endocardial BKT of the excitation
and repolarization wave front with the endocardial plane. On the endocardium,
the large spacing between successive isochrones indicates a fast excitation and
repolarization progression with a maximum apparent speed at the BKT point where
a sudden change of the wave front curvature occurs and the high curvature values
bring about high speeds of propagation around the BKT site, see e.g. Colli Franzone
et al. [102], Keener [268], Fast et al. [169]. The results for the excitation sequence
confirm earlier findings of Colli Franzone et al. [102, 116] obtained with an eikonal
approach, and are in agreement with previous experimental data (Taccardi et al.
278 9 Simulation Studies of Cardiac Bioelectrical Activity
[507]). The results relating the repolarization sequence are the main novelty of this
work.
APD distribution. Adding the intramural fiber rotation to the axisymmetric or
orthotropic anisotropic tissue, the epicardial and intramural APD distribution exhibit
a pattern displaying elongated level lines, (see Figs. 9.16 and 9.17), surrounding the
central point of the epicardial or intramural plane, corresponding to the stimulation
site and to first point reached by the excitation wave front. At first, the pattern
exhibits a quasi-elliptical shape with the major axis correlated with that of the
excitation isochrones. This indicates that APD decreases more rapidly when moving
away from the center of the face in the cross-fiber direction than along fibers.
Subsequently, the isochrones lose the quasi-elliptical shape but are still stretched
along a preferential direction, which does not coincide with the local fiber direction,
except at the epicardial level, see Panel MID in Figs. 9.16 and 9.17. The APD
decrease shows that the velocity of the repolarization front exceeds the one of the
activation front, both along and across the preferential directions, and moreover the
difference between the repolarization and the activation velocities is greater across
this direction, as confirmed by the more rapid APD decrease.
At some distance from the center of the face, the stretched portion of the level
lines of the APD, due to boundary effects, becomes straight while the presence of
valleys of APD reduction induces hollow, dog-bone profiles, more pronounced in
the orthotropic slab (see Fig. 9.17). The total dispersion in the axisymmetric and in
the orthotropic slabs amount to 16.7 and 14.8 ms, respectively.
On the intramural sections (from subepicardial (not shown) to midwall ones), two
valleys of decreasing APD values occur. Each valley exhibits a minimum located
at the boundary of the slab. These narrow valleys of low APD values are located
in the regions where excitation isochrones exhibit a dimple-like inflections. These
depression valleys are not related to boundary collisions but are strictly correlated
with the anisotropic influence of the intramural fiber rotation.
In the endocardial plane, the APD distribution displays a saddle point at the
endocardial BKT; the APD increases reaching a maximum when moving away from
the BKT point in a direction parallel to the endocardial fibers of 45ı CW, while
in the cross-fiber direction the APD decreases toward two minima located at the
boundary and corresponding to front-boundary collision points; the axis joining the
two maxima is roughly parallel to the endocardial fiber direction of 45ı CW and the
two maxima are separated by a valley of APD reduction.
We remark that the anisotropic features appearing in the APD pattern are more
pronounced in the orthotropic case; because these emphasized effects are likely
to help in the investigation of their origin and due to the recent interest about the
influence of orthotropic myocardial structure on the cardiac electric activity (see
LeGrice et al. [292,293], Costa et al. [137] for anatomical studies and Colli Franzone
et al. [121] for simulation studies), in the following the simulations have been
performed only for an orthotropic Monodomain model with orthotropic anisotropy.
9.2 Anisotropic Propagation of Excitation and Recovery Fronts 279
perpendicular to the epicardium, V-shaped intramural profiles with the epi- and
endo-cardial foot of the profiles overcoming the peak of the Vs. The simulated
anisotropic behavior of the excitation sequence is in agreement with previous
experimental and simulated observations, e.g. Burgess et al. [67], Taccardi et al.
[506–508], Colli Franzone et al. [102, 114, 116, 121], Keener [268] and Henriquez
et al. [226, 344].
With regard to repolarization, the recovery isochrones on the epi, midwall and
endocardial planes elicited by an epicardial pacing exhibit a somewhat smoother
shape and faster propagation compared with the excitation sequence. In particular,
the epicardial repolarization isochrones propagate across fibers faster than the
excitation wave, yielding a progressive APD shortening in the cross-fiber direction
of propagation as shown by Figs. 9.16 and 9.17. This prediction is in agreement with
the data of Gotoh et al. [208] related to cardiac laminae. The recovery isochrones
through the thickness of the wall show the presence of repolarization return
pathways as in the excitation sequence. These pathways accelerate the repolarization
process in epicardial areas where the recovery proceeds mainly across fibers. Unlike
the excitation propagation, experimental studies of intramural repolarization is less
complete and their interpretation is still the subject of intense research.
The finger-shape valleys of APD reduction associated to the dimple-like inflec-
tions appearing in the excitation and recovery fronts are the result of the interaction
between excitation and repolarization return pathways. Since these valleys are not
present in APD distribution in a slab with parallel fibers, they can be attributed to
the rotational anisotropic structure of the ventricular wall. Moreover, these valley are
narrower in the orthotropic case. These results relating the repolarization sequence
and APD dispersion are the main novelty of this work.
Anisotropic spatial variations of the APD along and across fibers were observed
experimentally in 2D cardiac laminae by Osaka et al. [364] and Gotoh et al. [208]
and on the epicardium of dog hearts by Burgess et al. [67] and Taccardi et al. [512].
More precisely, Taccardi et al. [512] observed a typical tripolar epicardial APD
pattern, with one central maximum and two minima along a cross-fiber direction,
on the left ventricle of exposed dog hearts during left ventricular pacing. Results
were recently confirmed by optical mapping. On the other hand, Bertran et al. [48],
working with pig hearts, did not observe anisotropic APD differences only in 2D
laminae.
We remark that simulation studies and experimental data have shown that
excitation return pathways, proceedings toward the pacing level, have been observed
for pacing sites located at any intramural level, from epi- to endo-cardium. On this
ground, we expect that repolarization return pathways, here reproduced only for
epicardial pacing, will be present also for intramural pacing.
Rules for APD modulation. Even if the plane sections reported cannot com-
pletely capture the complex evolution of the excitation and repolarization surfaces
in a three-dimensional domain, the results suggest the following rules for APD
modulation on plane sections:
9.2 Anisotropic Propagation of Excitation and Recovery Fronts 281
(i) Relatively long APDs occur at points where the excitation and recovery fronts
diverge, with maxima attained at the stimulation site and at each breakthrough
point on intramural planes parallel to the epicardium with the exception of the
endocardial plane;
(ii) A relative APD shortening occurs at points where the excitation and recovery
fronts converge, with minima attained at points of front-boundary collision and
front extinction;
(iii) Moving away from the stimulation site, the APD shortening is greater in the
directions across fibers than along fibers;
(iv) Valleys of relative APD shortening are associated with areas containing
dimple-like inflection in the excitation isochrones and the level lines of the
APD display dog-bone shaped profiles.
Anisotropic curvature as a parameter affecting the APD modulation. The
motion of the excitation wave front first depends on the characteristics of the
anisotropic media, on the other hand it is well known that the shape of the wave
front is a factor which contributes to the wave propagation, see e.g. [102, 114, 116,
169, 268, 273].
In the following, we consider an excitation (or repolarization) wave front locally
convex, relative to the propagating direction, according to whether the front bulges
into the resting (excited) volume; concave when it bulges into the excited (resting)
volumes from which it is coming.
With regard to the electrotonic load, we observe that currents flowing from the
excitation layers of a locally concave front are focusing toward points ahead of
the front, whereas for locally convex front they are scattering; the former case,
producing a more rapid membrane depolarization, yields a speed up of the excitation
propagation, compared to the velocity of a local plane front, whereas a slower
spreading results in the latter case. The opposite electrotonic process takes place
when considering points ahead of repolarization fronts i.e. for locally concave
(convex) fronts we have a depletion (convergence) of repolarizing currents. In spite
of this opposite behavior, we have similar effects on the recovery propagation, i.e.
in the concave case a faster recovery while in the convex case a slowing down
repolarization.
The dependence of the conduction velocity on the wave front shape is quanti-
tatively described by means of the curvature of the wave front. It is well known
that in isotropic homogeneous media with conductivity , the velocity, along the
Euclidean normal n of the front at a point x, oriented in the
pdirection of motion, can
be approximated by .x; n/ D p .1 C K=cm /, with K D k, where k D div n
is twice the mean curvature of the front, see e.g [268, 273]. The velocity of a local
plane front p p also depends on the conductivity of the excitable medium, i.e. we
have p D , with the constant only dependent on the membrane model. Then
a locally convex front, such as an expanding circle, has negative curvature, while a
locally concave front, such as a contracting circle, has positive curvature. Hence, a
locally convex (concave) front propagates with velocity smaller (greater) than that
of a plane wave.
282 9 Simulation Studies of Cardiac Bioelectrical Activity
For an anisotropic Monodomain slab the velocity of the excitation wave front is
described by a similar law of motion .x; n/ D p .x; n/.1 C K=cm /. The principal
term p .x; n/ predicts the velocity of an idealplocally plane wave propagating along
the normal direction, given by p .x; n/ D .x; n/, with .x; n/ D nT Dm .x/n
and Dm .x/ D Di .Di C De /1 De . The additional term is related to the anisotropic
curvature (see Bellettini et al. [37]), whose expression for the Monodomain model
by K D div , where differently from the usual Euclidean normal, .x/ D
is given p
D.x/n= nT Dn; see Bellettini et al. [39, 102, 268].
Burgess et al. [67] observed in experiments on pulmonary cone preparations that
the repolarization pattern can be affected by the direction of the excitation wave
front and that anisotropic differences in conduction velocity induce electrotonic
effects on the cell repolarization. Their findings show that the APD is relatively long
at sites where the activation front decelerates and is relatively short at sites where
the activation front accelerates, suggesting that acceleration and deceleration of the
excitation wave front are associated with a relative APD reduction and prolongation,
respectively. They established a significant correlation between the APD distribution
.
390.0 - 395.0
a b d 385.0 - 390.0
380.0 - 385.0
375.0 - 380.0
370.0 - 375.0
365.0 - 370.0
360.0 - 365.0
355.0 - 360.0
350.0 - 355.0
345.0 - 350.0
340.0 - 345.0
335.0 - 340.0
330.0 - 335.0
325.0 - 330.0
320.0 - 325.0
315.0 - 320.0
310.0 - 315.0
305.0 - 310.0
300.0 - 305.0
295.0 - 300.0
290.0 - 295.0
285.0 - 290.0
280.0 - 285.0
275.0 - 280.0
270.0 - 275.0
265.0 - 270.0
260.0 - 265.0
255.0 - 260.0
250.0 - 255.0
245.0 - 250.0
240.0 - 245.0
235.0 - 240.0
230.0 - 235.0
225.0 - 230.0
220.0 - 225.0
215.0 - 220.0
210.0 - 215.0
205.0 - 210.0
200.0 - 205.0
195.0 - 200.0
42.4 ( 5.0 ) 119.9 0.5 ( 5.0 ) 88.0 21.8 ( 5.0 ) 129.7 190.0 - 195.0
185.0 - 190.0
180.0 - 185.0
175.0 - 180.0
f e 170.0 - 175.0
165.0 - 170.0
160.0 - 165.0
155.0 - 160.0
150.0 - 155.0
145.0 - 150.0
140.0 - 145.0
135.0 - 140.0
c
130.0 - 135.0
125.0 - 130.0
120.0 - 125.0
115.0 - 120.0
110.0 - 115.0
105.0 - 110.0
100.0 - 105.0
95.0 - 100.0
90.0 - 95.0
85.0 - 90.0
80.0 - 85.0
75.0 - 80.0
70.0 - 75.0
65.0 - 70.0
60.0 - 65.0
55.0 - 60.0
50.0 - 55.0
45.0 - 50.0
40.0 - 45.0
35.0 - 40.0
30.0 - 35.0
25.0 - 30.0
20.0 - 25.0
15.0 - 20.0
10.0 - 15.0
5.0 - 10.0
1.1 (5.0) 93.3 15.2 ( 5.0 ) 93.0 0.2 ( 5.0 ) 145.7 0.0 - 5.0
below 0.0
Fig. 9.18 Axisymmetric monodomain model with intramural fiber rotation on a truncated ellip-
soidal domain with central endocardial stimulation. Activation time isochrones on endocardial
(e), mid-myocardial (d), epicardial (a) and different transmural sections (b-c-f). All values are
expressed in ms
9.2 Anisotropic Propagation of Excitation and Recovery Fronts 283
.
390.0 - 395.0
385.0 - 390.0
a b d
380.0 - 385.0
375.0 - 380.0
370.0 - 375.0
365.0 - 370.0
360.0 - 365.0
355.0 - 360.0
350.0 - 355.0
345.0 - 350.0
340.0 - 345.0
335.0 - 340.0
330.0 - 335.0
325.0 - 330.0
320.0 - 325.0
315.0 - 320.0
310.0 - 315.0
305.0 - 310.0
300.0 - 305.0
295.0 - 300.0
290.0 - 295.0
285.0 - 290.0
280.0 - 285.0
275.0 - 280.0
270.0 - 275.0
265.0 - 270.0
260.0 - 265.0
255.0 - 260.0
250.0 - 255.0
245.0 - 250.0
240.0 - 245.0
235.0 - 240.0
230.0 - 235.0
225.0 - 230.0
220.0 - 225.0
215.0 - 220.0
210.0 - 215.0
205.0 - 210.0
200.0 - 205.0
195.0 - 200.0
289.1 ( 5.0 ) 366.2 258.6 ( 5.0 ) 333.6 273.5 ( 5.0 ) 377.2 190.0 - 195.0
185.0 - 190.0
180.0 - 185.0
f
175.0 - 180.0
e
170.0 - 175.0
165.0 - 170.0
160.0 - 165.0
155.0 - 160.0
150.0 - 155.0
145.0 - 150.0
140.0 - 145.0
135.0 - 140.0
c
130.0 - 135.0
125.0 - 130.0
120.0 - 125.0
115.0 - 120.0
110.0 - 115.0
105.0 - 110.0
100.0 - 105.0
95.0 - 100.0
90.0 - 95.0
85.0 - 90.0
80.0 - 85.0
75.0 - 80.0
70.0 - 75.0
65.0 - 70.0
60.0 - 65.0
55.0 - 60.0
50.0 - 55.0
45.0 - 50.0
40.0 - 45.0
35.0 - 40.0
30.0 - 35.0
25.0 - 30.0
20.0 - 25.0
15.0 - 20.0
10.0 - 15.0
5.0 - 10.0
258.7 (5.0) 339.0 261.5 ( 5.0 ) 389.8 258.5 ( 5.0 ) 389.8 0.0 - 5.0
below 0.0
Fig. 9.19 Axisymmetric Monodomain model with intramural fiber rotation on a truncated ellip-
soidal domain with central endocardial stimulation. Repolarization time isochrones on endocardial
(e), mid-myocardial (d), epicardial (a) and different transmural sections (b-c-f). All values are
expressed in ms
a b d
above 259.5
244.5 ( 0.5 ) 250.9 244.9 ( 0.5 ) 258.1 246.1 ( 0.5 ) 251.7 259.0 - 259.5
258.5 - 259.0
258.0 - 258.5
257.5 - 258.0
f e
257.0 - 257.5
256.5 - 257.0
256.0 - 256.5
255.5 - 256.0
255.0 - 255.5
254.5 - 255.0
254.0 - 254.5
253.5 - 254.0
c 253.0 - 253.5
252.5 - 253.0
252.0 - 252.5
251.5 - 252.0
251.0 - 251.5
250.5 - 251.0
250.0 - 250.5
249.5 - 250.0
249.0 - 249.5
248.5 - 249.0
248.0 - 248.5
247.5 - 248.0
247.0 - 247.5
246.5 - 247.0
246.0 - 246.5
245.5 - 246.0
245.0 - 245.5
244.5 - 245.0
244.0 - 244.5
243.5 - 244.0
243.0 - 243.5
242.5 - 243.0
242.0 - 242.5
241.5 - 242.0
241.0 - 241.5
240.5 - 241.0
244.6 (0.5) 257.6 244.1 ( 0.5 ) 253.5 244.1 ( 0.5 ) 258.4 240.0 - 240.5
below 240.0
Fig. 9.20 Axisymmetric monodomain model with intramural fiber rotation on a truncated ellip-
soidal domain with central endocardial stimulation. Action potential duration isochrones on
endocardial (e), mid-myocardial (d), epicardial (a) and different transmural sections (b-c-f). All
values are expressed in ms
the parameter values is not an easy task, because the data available in the literature
refer to diverse experimental setups. The macroscopic conductivities of the media
are here based on conservative values and are determined by requiring that the
simulated excitation wavefront motion has physiologically reasonable velocities of
propagations, see [102, 123]. The conductivity coefficients, in mS cm1 , used in the
orthotropic anisotropic simulations are:
le D 2 te D 1:35 li D 3 ti D 0:465 ne D 0:5te ; ni D 0:1ti :
The initial conditions for the Monodomain model with the phase I Luo-Rudy
model (LR1) [308] are chosen at the steady state, so that the rest potential is v0 D
83:84 mV. In order to elicit the excitation front, we apply a stimulus of 250 A
for 1 ms on a small volume (containing 5 mesh points in each x and ydirection
and 3 mesh points in the zdirection) at the center of the epi- or endocardial face.
Other than potentials and gating variables, at each time-step, we compute also the
activation (ACTI) and the repolarization (REPO) times, defined as the times when
the action potential crosses 60 mV during the upstroke and when it reaches 75:46
in the downstroke phase (i.e. 90 % of the resting value), respectively.
Intrinsic transmural heterogeneity. We consider three different types of
transmural distribution of the intrinsic APDs of the cells, the first homogeneous
(denoted by H-slab) and the other two heterogeneous, (denoted by 3-slab and W-
slab). We assume that the transmural intrinsic heterogeneity is the same along any
transmural epi-endocardial straight line, i.e. in any plane parallel to the epicardium
all cells have the same APD.
• H-slab: the cardiac slab has homogeneous intrinsic properties of the cellular
membrane; the IK current of the LR1 model is multiplied by 2.325, yielding
an APD of 250 ms.
• 3-slab: the cardiac slab is subdivided transmurally into three layers of 1=3 cm
thickness. More precisely, proceeding from the endocardial to the epicardial
surfaces, the values of intrinsic cellular APD amount to 235, 272, 225 ms
on the intervals Œ0; 0:33/; Œ0:33; 0:66/; Œ0:66; 1, respectively. The transmural
heterogeneity is simulated by multiplying the current IK of the LR1 model
by (2.62, 1.952, 2.88), corresponding to intrinsic APDs of (235, 272, 225)
ms, respectively. See Fig. 9.21, left panel and the dashed plots in Fig. 9.23.
Figure 9.21 left panel, displays the three different action potential waveforms
related to the three different intrinsic properties of transmural cells, i.e. they are
3−SLAB W−SLAB
50 50
0 0
AP (mV)
Mid
AP (mV)
Mid
Subendo
Endo
Endo
−100 −100
0 100 200 300 0 100 200 300
Fig. 9.21 Intrinsic action potential waveforms assigned to the intramural heterogeneity of type
3-slab (left) and W-slab (right). Each panel shows subendocardial, midwall and subepicardial
intrinsic action potentials obtained with 0D simulations and assigned to each transmural layer
9.3 Heterogeneous Cardiac Tissue 287
o
A (45 )
o
B (90 )
o
D (0 )
C
EPI
C (−45o)
ENDO
Fig. 9.22 Intramural sections A, B, C, and D of the orthotropic Monodomain slab of size 5 5
1 cm3 , with isochrone contours elicited by epicardial central stimulation
We will then show some results on homogeneous and heterogeneous 3-D curved
walls.
In each case, we will examine the distributions of activation times (ACTI(x)),
repolarization times (REPO(x)) and action potential durations (APD(x) D
REPO(x)–ACTI(x)), generated by a stimulus at the center of the epicardial or
endocardial surface of each slab.
We remark that fiber direction at a depth of 0 cm from the epicardium is parallel
to section C; at 0.375, to section D; at 0.75 cm, to section A, while the fiber direction
is never parallel to section B.
40 280
280
260
3−slab
20 260
240
240
0 220
0 0.5 1 0 0.5 1 0 0.5 1
MIN = 0.13 MAX = 35.26 MIN = 233.20 MAX = 279.96 MIN = 231.10 MAX = 260.09
40 280
280
30
260
W−slab
20 260
240
10
240
0 220
0 0.5 1 0 0.5 1 0 0.5 1
EPI MIN = 0.13 MAX = 35.26 MIN = 236.52 MAX = 283.31 MIN = 235.66 MAX = 255.57
20 270
250
260
0 250 245
0 0.5 1 0 0.5 1 0 0.5 1
MIN = 0.13 MAX = 35.48 MIN = 258.42 MAX = 282.32 MIN = 246.84 MAX = 258.29
40 280 280
260
3−slab
20 260
240
0 240 220
0 0.5 1 0 0.5 1 0 0.5 1
MIN = 0.13 MAX = 35.48 MIN = 245.15 MAX = 278.65 MIN = 223.44 MAX = 260.80
40 280 280
30 270
260
W−slab
20 260
240
10 250
0 240 220
0 0.5 1 0 0.5 1 0 0.5 1
ENDO MIN = 0.13 MAX = 35.48 MIN = 252.94 MAX = 267.68 MIN = 226.79 MAX = 258.57
Fig. 9.23 Profiles along the 1-D transmural line (passing through the stimulation site and
orthogonal to both epicardial and endocardial surfaces) of the activation time (ACTI, first column),
of the repolarization time (REPO, second column), of the action potential duration (APD, third
column) for H-slab (first row), 3-slab (second row), W-slab (third row). Central stimulation:
epicardial (top panel), endocardial (lower panel). The piecewise constant dashed lines in the APD
column are the intrinsic (0D) APD of the cells. Vertical axis reports times in ms and the horizontal
axis reports distances in cm from the endocardial site
290 9 Simulation Studies of Cardiac Bioelectrical Activity
The midwall maximum in 3-slab is related to the collision of the primary recovery
front, originated at the endocardial stimulation site, with a secondary phase front,
originating at the epicardial side at a later instant. In W-slab, two collisions of
the recovery fronts are present because an additional recovery wave, starting from
the intramural site corresponding to a minimum of the repolarization time profile,
spreads toward the endocardium and the other toward the epicardium. This dynamic
behavior is confirmed by the repolarization speed profile in which, the intramural
maxima correspond to points of collision.
The APD dispersions, amounting to 37.36 (3-slab) and to 31.78 ms (W-slab), are
about four and three times greater than in the H-slab.
In comparison with the H-slab case, the endocardial stimulus yields a reduction
or increase of the recovery dispersion depending on the type of heterogeneity.
Conversely, the epicardial stimulus produces recovery dispersions, in both 3-slab
and W-slab, that are roughly twice the H-slab dispersion. For both stimulations, the
APD dispersions are higher than the H-slab APD dispersion, with greater values
in case of endocardial stimulation and 3-slab heterogeneity. Thus, the type of
intrinsic transmural heterogeneities could be inferred by comparing repolarization
profiles and APD dispersions along the 1-D transmural line, where epicardial and
endocardial stimulations produce very different values of repolarization and APD
dispersions.
Comparing the results of the two stimulations, we observe that the repolarization
profiles display greater morphological changes than the APD profiles, suggesting a
major sensitivity of the former to the distribution of transmural heterogeneity.
H−slab
60
40
20
0.14 114.08 4.00
100
80
3−step
60
40
20
0.14 114.08 4.00
100
80
W−slab
60
40
20
ACTI 0.14 114.08 4.00
REPO ENDO
340
320
H−slab
300
280
258.44 358.33 4.00 260
350
3−step
300
340
320
W−slab
300
280
252.95 360.27 4.00 260
REPO
APD EPI
255
H−slab
250
250
3−slab
240
230
223.18 260.80 4.00
250
W−slab
240
Fig. 9.24 Isochrone lines of the activation time (ACTI, top), repolarization time (REPO, middle)
and action potential duration (APD, bottom) on the transmural section C for models H-slab, 3-slab,
W-slab, with endocardial stimulation. Reported below each panel are the maximum, minimum and
step in ms of the displayed map (Reproduced with permission from [126])
9.3 Heterogeneous Cardiac Tissue 293
In the 3-slab case, recovery starts at the stimulation site. After about 13 ms,
a secondary recovery wave starts from the opposite point on the epicardial side.
After collision at midwall level, the merged repolarization front spreads laterally
with a smooth V-shaped profile. In the W-slab case, as discussed in Sect. 9.3.1.1,
the recovery process arises from three different sites and at different times: at
253 ms at the endocardial stimulation site, at 262 ms from the opposite point on the
epicardial side. Finally, at 265 ms, a new repolarization front arises at about midwall
point. After the merging of these three different recovery fronts, the transmural
isochrones exhibit V-shape profiles with some initial undulation associated with the
four different transmural layers of intrinsic heterogeneity.
The repolarization dispersion on section C amounts to (99.89, 110.84, 108.32)
ms for H-slab, 3-slab, W-slab, respectively. These values are lower then those for
activation dispersion, which amount to about 114.08 ms. This confirms that recovery
is slightly faster than excitation, especially in the homogeneous case (Table 9.3).
(c) Transmural APD distribution on section C. The APD dispersion on section
C amounts to (14.05, 37.62, 32.02) ms for H-slab, 3-slab, W-slab, respectively,
These values are comparable to the values of APD dispersion observed along
the 1-D transmural line described in Sect. 9.3.1.1. In the H-slab, the APD distri-
bution, displayed in Fig. 9.24 (bottom panel), exhibits the absolute maximum
at the stimulation site and two symmetric subepicardial minima. Shortening
of the APD and associated relative minima are present at points of front-
boundary collision of the excitation wavefront. Two maxima on the epicardial
side appear symmetrically with respect to the minimum located at the excitation
breakthrough point on the epicardium.
In both heterogeneous slabs, the APD distribution displays a striped pattern; the
stripe having the highest APD has the same location as the intramural layer with the
highest intrinsic APD.
We now consider the excitation, repolarization and APD values on the whole 3-D
slabs and their planar intramural sections parallel to the epicardial face, with both
endocardial and epicardial pacing.
Table 9.3 Dispersion (MAX–MIN value) of activation time (ACTI), repolarization time (REPO),
APD 90 on the transmural sections A and C
Epicardial stimulation, section A Endocardial stimulation, section C
ACTI REPO APD 90 ACTI REPO APD 90
H-slab 100.53 88.00 12.39 114.08 99.89 14.05
3-slab 100.53 107.81 38.42 114.08 110.84 37.62
W-slab 100.53 95.21 30.54 114.08 108.32 32.02
294 9 Simulation Studies of Cardiac Bioelectrical Activity
(a) H-slab. In the homogeneous case, displayed in Fig. 9.25, the spread of excita-
tion and the sequence of recovery, displayed on five planar sections parallel to
the epicardial face (located at z = 0, 0.25, 5, 0.75, 1 cm, respectively), undergo
an acceleration in endocardial areas where the fronts proceed mainly across
fibers, due to the presence of underlying intramural returning pathways of
excitation and recovery, as shown by the isochrones displayed on the transmural
section C with endocardial stimulation (Fig. 9.24). Dimple-like inflections
appear in the isochrone profiles, due to the faster propagation of the fronts in
higher layers where the fiber direction rotates CW relatively to the lower planes.
The recovery isochrones on the endo, intramural and epicardial planes exhibit a
somewhat smoother shape and faster propagation compared with the excitation
sequence.
The rotational anisotropy of the media yields APD patterns strongly correlated
with the excitation wavefront motion and the front-boundary collisions. The APD
spatial distribution shows finger-shape valleys of APD shortening associated to the
dimple-like inflections appearing in the excitation and recovery fronts. The APD
level lines display also dog-bone shaped profiles. These features can be attributed
to the rotational anisotropic structure of the ventricular wall and to the interaction
between excitation and repolarization return pathways. Moreover, APD maxima
occur near the stimulation site and at breakthrough points on each plane section
parallel to the epicardium (except on the epicardial plane), see [123, Fig. 7] for more
detailed analysis of the APD pattern.
Similar features occur in case of epicardial stimulation (not shown), if we
exchange in the description the role of endocardial and epicardial planes.
(b) 3-slab and W-slab. We now compare the H-slab patterns with those of
the heterogeneous 3-slab and W-slab, see Figs. 9.26 and 9.27. Unexpectedly,
the repolarization sequence and the APD patterns on the intramural planar
sections parallel to the epicardium exhibit the same main anisotropic spatial
features as in the homogeneous case. Therefore, the transmural heterogeneity
remains mostly confined in the transmural direction and does not diffuse to the
orthogonal planes, in spite of the combined effects of electrotonic modulation,
orthotropic conduction velocity and fiber rotation. We have computed the
correlation coefficients of the repolarization time and APD between H-slab and
3-slab and between H-slab and W-slab on each of the 101 intramural planar
sections parallel to the epicardial face (the first being the endocardium and the
last the epicardium). These coefficients show a remarkably strong correlation
between homogeneous and heterogeneous repolarization times (the coefficients
differ from 1 by less than 2 104 ) and APD (the coefficients differ from 1
by at most 5:2 102 with epicardial stimulation and at most 7:1 102 with
endocardial stimulation), despite the different magnitude of the repolarization
times and APDs.
9.3 Heterogeneous Cardiac Tissue 295
1 CM (EPI)
100 340
80 255
320
60
300
40 250
20 280
245
0.14 114.08 4.00 258.44 358.33 4.00 244.23 258.32 0.50
Fig. 9.26 Orthotropic Monodomain–LR1 model, heterogeneous 3-slab of size 551 cm3 . Same
format as in Fig. 9.25 (Reproduced with permission from [126])
9.3 Heterogeneous Cardiac Tissue 297
100
340
80 250
320
60
300 240
40
280
20 230
260
0.14 114.08 4.00 252.95 360.27 4.00 224.64 258.07 0.50
In the intramural sections parallel to the epicardium, the APD patterns elicited
by an endocardial (epicardial) central stimulation in the homogeneous or heteroge-
neous slabs are characterized by the following features:
(i) The APD distributions on the endocardial (epicardial) and intramural planes
exhibit a maximum located at the endocardial (epicardial) stimulation site
or at the intramural points first reached by the excitation front, respectively;
see Figs. 9.25–9.27. These maxima are surrounded by level lines elongated
along the local fiber direction. This indicates that APD decreases more rapidly
when moving away from the center of the face in the cross-fiber direction
than along fibers. This finding confirms and extends to a three-dimensional
myocardial slab the experimental findings by Gotoh et al. [208], Osaka et al.
[364], Taccardi et al. [512].
(ii) On the intramural sections (from subendocardial (subepicardial) to midwall),
two valleys of decreasing APD values occur. Each valley exhibits a minimum
located at the boundary of the slab. These narrow valleys of low APD values
are located in the regions where excitation isochrones exhibit a dimple-like
inflections.
(iii) In the epicardial (endocardial) plane, the APD distribution displays a saddle
point at the epicardial (endocardial) BKT point. The APD increases reaching
a maximum when moving away from the BKT point in a direction parallel
to the epicardial (endocardial) fibers of 45ı (75ı ), while in the cross-fiber
direction the APD decreases toward two minima located at the boundary and
corresponding to front-boundary collision points. The axis joining the two
maxima is roughly parallel to the epicardial (endocardial) fiber direction of
45ı (75ı ) and the two maxima are separated by the valley of APD reduction.
We remark that the heterogeneity we consider is only transmural, i.e. the
intrinsic membrane properties are constant in each intramural plane parallel to the
epicardium. Our results show that the presence of transmural heterogeneities cannot
be detected from the epicardial pattern of the APD distribution.
After the endocardial or epicardial breakthrough, the repolarization fronts spread
laterally in the slab (Fig. 9.24) and display similar shapes associated with the 1-D
transmural profiles of Fig. 9.23, independently of the pacing location.
Here r 2 Œ0; 1; 2 Œ min ; max ; 2 Œmin ; max ; a.r/ D a1 Cr.a2 a1 /; c.r/ D
c1 C r.c2 c1 / and ai ; ci ; i D 1; 2 are given coefficients determining the main
axes of the ellipsoid. We denote by r-layers the ellipsoidal surfaces described by
varying ; for a fixed value of r; analogously, we denote by -layers the plane
surfaces described by varying r; for a fixed value of and by -layers the surfaces
described by varying r; for a fixed value of (see Fig. 9.28 and Table 9.4).
We consider transmural 3-wall and W-wall heterogeneous distribution of intrinsic
APD. More precisely:
W-wall. The cardiac tissue is subdivided transmurally into four layers: endo-
cardial (r 2 Œ0; 0:12/), sub-endocardial (r 2 Œ0:12; 0:32/), midmyocardial (r 2
Œ0:32; 0:85/) and epicardial (r 2 Œ0:85; 1). The IK factor assumes the following
values
8
ˆ
ˆ 2:715 r 2 Œ0; 0:12/ (endo)
<
1:952 r 2 Œ0:12; 0:32/ (sub-endo)
factW D
IK
ˆ 2:47 r 2 Œ0:32; 0:85/ (mid)
:̂
2:88 r 2 Œ0:85; 1 (epi);
θ − layer
φ−layer
(θ = π/8)
θ−layer
S
S2 1
r−layer ENDO (r=0)
P1
P2 MID (r=0.5)
P3
φ−layer
EPI (r=1) r−layers
(φ = − π/2)
Fig. 9.28 Left: ellipsoidal domain and r, ; -layers, S1 ; S2 D endocardial and epicardial
stimulation sites, P1 ; P2 ; P3 D Purkinje ventricular junctions. Right: fiber direction on r, ; -
layers
Table 9.4 Parameter calibration for the monodomain and bidomain models
D 103 cm1 Cm D 103 mF=cm2
le D 2 103 1 cm1 li D 3 103 1 cm1
te D 1:3514 103 1 cm1 ti D 3:1525 104 1 cm1
ne D te =2 ni D ti =10
min D =2, max D =2 min D 3=8, max D =8
a1 D 1:5, a2 D 2:7 c1 D 4:4, c2 D 5
n D 500, n D 500, nr D 100
300 9 Simulation Studies of Cardiac Bioelectrical Activity
50 280
MID
−100 220
0 100 200 300 ENDO MID EPI
50 280
MID
APD (ms)
0 260
AP (mV)
ENDO
−50 240
EPI
−100 220
0 100 200 300 ENDO MID EPI
50 270
260
APD (ms)
0
APEX
AP (mV)
250
−50 240
BASE
230
−100
0 100 200 300 APEX MID BASE
Fig. 9.29 Top: W-wall transmural intrinsic action potential profiles (left) and transmural intrinsic
APD distribution (right). Middle: 3-wall transmural intrinsic action potential profiles (left) and
transmural intrinsic APD distribution (right). Bottom: apex-to-base intrinsic action potential
profiles (left) and apex-to-base intrinsic APD distribution (right)
9.3 Heterogeneous Cardiac Tissue 301
This type of transmural heterogeneity composed of three cell layers has been
previously used in [546] for 1D simulations.
Figures 9.30–9.33 display the patterns of excitation, repolarization and APD
isochrones for the three ellipsoidal walls with endocardial stimulation. The same
qualitative features of the excitation, repolarization and APD patterns previously
described for the slab domains are also detected in these ellipsoidal walls. The visual
inspection of the repolarization patterns on r-layers of the H-wall, 3-wall and W-wall
shows that they are very similar, as confirmed by the high CC ( 0:991) reported
in Table 9.6. On the other hand, the APD correlation is lost on some r-layers. This
correlation loss of the intramural APD patterns could be attributed to the interaction
between the wall tapering and the associated curvature variation. This conclusion
differs from the one obtained for cartesian slabs, where both repolarization and
APD patters on intramural r-layers were found to be independent of the transmural
variation of the intrinsic cellular APD.
9.3.2.1 Discussion
In the present investigation we studied the combined influence of (a) the transmural
rotational anisotropy and (b) the transmural intrinsic APDs heterogeneity on the
sequence of repolarization and APD distribution. We have used the Anisotropic
Monodomain model coupled with the Luo-Rudy phase I membrane model, which
has a lower computational cost than more advanced models incorporating additional
ionic currents and a more complete intracellular calcium dynamics, see e.g. [243,
356]. We applied a single stimulus, either epicardial or endocardial, and compared
the results obtained with homogeneous membrane properties (H-slab/wall) with the
results obtained with two types of transmural heterogeneity (3-slab/wall and W-
slab/wall). By considering action potentials elicited by a single stimulus (as opposed
to a sequence of stimuli), our simulations are expected to enhance the disparity
between excitation and recovery sequences and the dispersion of APDs, since it is
well known that periodic stimulations, with relatively short cycle lengths, produce
relatively short APDs with a reduced dispersion, see [479, Fig. 3].
In case of homogeneous membrane (H-slab/wall), the repolarization wave
elicited by epi- or endocardial stimulation spreads anisotropically, like the
corresponding activation pattern. Repolarization exhibits a somewhat faster motion,
with intramural anisotropic patterns of APD distribution in sections parallel to
the epicardial face and on transmural sections A, B, C, and D. Although the
tissue is composed of cells with the same intrinsic properties, the electrotonic
effects generate anisotropic spatial variations (epicardial and transmural) of the
APD. The maximum of APD over the entire volume is consistently located at
the stimulation site. Anisotropic spatial variations of the APD along and across
fibers were observed experimentally in 2-D myocardial laminae by Osaka et al.
[364], Gotoh et al. [208] and on the epicardium of dog hearts by Burgess et al.
[67] and Taccardi et al. [512]. On the other hand, Bertran et al. [48], working with
pig hearts, did not observe anisotropic APD differences in 2-D laminae from the
302 9 Simulation Studies of Cardiac Bioelectrical Activity
144 388
258
93 343
253
45 298
248
0 259 244
0.12 145.71 3.00 258.52 389.76 3.00 244.05 258.42 1.00
144
392
256
93 344
244
45 296 233
0 251 223
Fig. 9.31 Orthotropic Monodomain–LR1 model, W-wall with endocardial central stimulation.
Isochrone lines of the depolarization time (first column ACTI), repolarization time (second column
REPO) and action potential duration (third column APD) on epicardial (first row), mid-myocardial
(second row) and endocardial (third row) sections and on the whole domain (fourth row). Reported
below each panel are the maximum, minimum and step in ms of the displayed map
304 9 Simulation Studies of Cardiac Bioelectrical Activity
EPI
MID
ENDO
93 338 247
45 290 233
Fig. 9.32 Orthotropic Monodomain–LR1 model, 3-wall with endocardial central stimulation.
Isochrone lines of the depolarization time (first column ACTI), repolarization time (second column
REPO) and action potential duration (third column APD) on epicardial (first row), mid-myocardial
(second row) and endocardial (third row) sections and on the whole domain (fourth row). Reported
below each panel are the maximum, minimum and step in ms of the displayed map
9.3 Heterogeneous Cardiac Tissue 305
above 263.5
263.0 - 263.5
244.5 (0.5) 250.9 220.5 (0.5) 226.6 223.2 (0.5) 229.9 251.5 - 252.0
251.0 - 251.5
250.5 - 251.0
250.0 - 250.5
249.5 - 250.0
249.0 - 249.5
248.5 - 249.0
M 248.0 - 248.5
247.5 - 248.0
I 247.0 - 247.5
246.5 - 247.0
D 246.0 - 246.5
245.5 - 246.0
245.0 - 245.5
244.5 - 245.0
244.0 - 244.5
243.5 - 244.0
243.0 - 243.5
242.5 - 243.0
242.0 - 242.5
241.5 - 242.0
241.0 - 241.5
246.1 (0.5) 251.7 247.5 (0.5) 263.1 239.9 (0.5) 246.5 240.5 -
240.0 -
241.0
240.5
239.5 - 240.0
239.0 - 239.5
E 238.5 - 239.0
238.0 - 238.5
237.5 - 238.0
N 237.0 - 237.5
236.5 - 237.0
D 236.0 -
235.5 -
236.5
236.0
O 235.0 - 235.5
234.5 - 235.0
234.0 - 234.5
233.5 - 234.0
233.0 - 233.5
232.5 - 233.0
232.0 - 232.5
231.5 - 232.0
231.0 - 231.5
230.5 - 231.0
230.0 - 230.5
229.5 - 230.0
244.1 (0.5) 258.4 231.0 (0.5) 244.8 237.9 (0.5) 251.3 229.0 - 229.5
228.5 - 229.0
228.0 - 228.5
227.5 - 228.0
227.0 - 227.5
226.5 - 227.0
226.0 - 226.5
225.5 - 226.0
225.0 - 225.5
224.5 - 225.0
224.0 - 224.5
223.5 - 224.0
223.0 - 223.5
222.5 - 223.0
222.0 - 222.5
221.5 - 222.0
221.0 - 221.5
220.5 - 221.0
220.0 - 220.5
below 220.0
Table 9.5 Dispersion (MAX–MIN value) of activation time (ACTI), repolarization time (REPO),
APD 90 on the 1-D transmural line passing through the stimulation site and orthogonal to both
epicardial and endocardial surfaces
Epicardial stimulation Endocardial stimulation
ACTI REPO APD 90 ACTI REPO APD 90
H-slab 35.26 24.62 10.53 35.48 23.90 11.45
3-slab 35.26 46.76 28.99 35.48 33.50 37.36
W-slab 35.26 46.79 19.91 35.48 14.74 31.78
transmural dispersion of APD in the left ventricular wall is on the order of 15–20 ms
during ventricular pacing with cycle length of 350 or 400 ms. In these preparations,
the repolarization sequence is often qualitatively similar to the activation sequence.
When the pacing site was epicardial, both the excitation and the repolarization
fronts returned toward the epicardium in a transmural plane perpendicular to the
epicardial fiber direction near the pacing site. However, confirmatory studies are
needed in order to determine whether these findings occur consistently in varying
experimental conditions. Moreover, Taccardi et al. [512] observed a typical tripolar
epicardial APD pattern, with one central maximum and two minima in the cross-
fiber direction, on the left ventricular epicardium of exposed dog hearts during left
ventricular pacing. Poelzing et al. [395, 396] found that in a narrow subepicardial
layer the intercellular conductance is lower than in deeper layers, due to a local
shortage of gap junctions. This factor contributes to the transmural inhomogeneity
of excitation velocity and recovery durations.
As mentioned before, the data on transmural dispersion of APD currently
available in the literature are not consistent, since there is a disparity between the
transmural dispersion of the APD observed in in vivo hearts, see e.g. [9, 258], and
in vitro experiments in wedge preparations, see e.g. [8, 572]. The transmural APD
dispersion reported by the former authors is less than the one reported by the latter.
In this simulation study on the electrotonic modulation in a three dimensional
structure, we assumed that: (i) the heterogeneity of the intrinsic APDs is transmural,
i.e. the APDs values are constant in each section parallel to the epicardial face;
(ii) the fibers have a constant orientation in each section parallel to the epicardial
face; (iii) an idealized sheet-fiber model (iv) a normal cell coupling (i.e. normal
conductivity coefficients). The simulations show that the electrotonic modulation
is not strong enough to level the intrinsic transmural APD differences, but these
strong differences are not revealed at the epicardial level. Hence, our results do not
explain why in intact hearts several investigators, see e.g. [9, 258], have observed
a strongly reduced dispersion of APD across the ventricular wall as compared with
the dispersion observed in wedge preparations.
In conclusion, the complex 3-D paths of electrotonic currents during the repo-
larization phase in our orthotropic medium do not seem to extend the effects of
transmural inhomogeneity to the sections parallel to the epicardium. These results
suggest the presence of two separate and competing mechanisms that modulate the
repolarization sequence and APD patterns, one due to the rotational anisotropy and
the other to the transmural heterogeneity of intrinsic APD. While the first seems to
prevail in the sections parallel to the epicardium, the second is the major determinant
of the transmural shape of repolarization fronts and APD patterns outside a volume
surrounding the transmural line issuing from the stimulation site.
9.3 Heterogeneous Cardiac Tissue 309
We now extend the previous simulations in ellipsoidal cardiac domains to the case
of apex-to-base and mixed transmural APD heterogeneities; for more results and
details we refer to [130]. In particular, we consider the following heterogeneous
ellipsoidal domains:
Apex-base-wall. The intrinsic APDs of cells vary in this case with respect to the
-direction. We define the intrinsic APD along the apex-to-base direction decreasing
monotonically from apex to base, by introducing a scaling factor factIK of the IK
current depending linearly on according to the following rule
2:88 1:952
factIAB ./ D . min / C 1:952;
K
max min
for 2 Œmin ; max . The resulting dependence of APD on is reported in Fig. 9.29,
showing an almost linear variation from 272 ms at the apex to 225 ms at the base,
with an intrinsic APD dispersion of 47 ms.
Mixed-wall. Both transmural 3-wall and apex-to-base-wall heterogeneities are
introduced in the cardiac tissue by setting factIK D fact3IK C factAB
IK ./, i.e.
8
< 2:62 C factAB ./ 2 Œmin ; max ; r 2 Œ0; 0:33/ (endo)
factIK D 1:952 C factAB ./ 2 Œmin ; max ; r 2 Œ0:33; 0:66/ (mid)
:
2:88 C factAB ./ 2 Œmin ; max ; r 2 Œ0:66; 1 (epi):
E 395
P 390 - 395
385 - 390
380 - 385
I 375 - 380
370 - 375
365 - 370
360 - 365
355 - 360
350 - 355
345 - 350
340 - 345
335 - 340
330 - 335
325 - 330
320 - 325
42.6 ( 5.0) 119.9 315 - 320 279.9 (5.0) 371.7 220.5 ( 1.0) 266.4
310 - 315
305 - 310
300 - 305
295 - 300 277
290 - 295 276 - 277
285 - 290 275 - 276
M 280 - 285
275 - 280
274 - 275
273 - 274
I 270 - 275
265 - 270
272 - 273
271 - 272
D 260 - 265
255 - 260
270 - 271
269 - 270
250 - 255 268 - 269
245 - 250 267 - 268
240 - 245 266 - 267
235 - 240 265 - 266
230 - 235 264 - 265
225 - 230 263 - 264
220 - 225 262 - 263
215 - 220 261 - 262
210 - 215 260 - 261
205 - 210 259 - 260
D 155 - 160
150 - 155
249 - 250
248 - 249
O 145 - 150
140 - 145
247 - 248
246 - 247
135 - 140 245 - 246
130 - 135 244 - 245
125 - 130 243 - 244
120 - 125 242 - 243
115 - 120 241 - 242
110 - 115 240 - 241
105 - 110 239 - 240
100 - 105 238 - 239
95 - 100 237 - 238
90 - 95 236 - 237
85 - 90 235 - 236
0.2 ( 5.0) 146.0 80 - 85 249.2 (5.0) 378.2 219.7 ( 1.0) 268.5 234 - 235
75 - 80 233 - 234
70 - 75 232 - 233
65 - 70 231 - 232
60 - 65 230 - 231
55 - 60 229 - 230
50 - 55 228 - 229
45 - 50 227 - 228
40 - 45 226 - 227
35 - 40 225 - 226
30 - 35 224 - 225
25 - 30 223 - 224
20 - 25 222 - 223
15 - 20 221 - 222
10 - 15 220 - 221
5 - 10 219 - 220
0- 5 218 - 219
0 218
Fig. 9.34 Apex-base-wall with central endocardial stimulation. Same format as Fig. 9.36 (Repro-
duced with permission from [130])
9.3 Heterogeneous Cardiac Tissue 311
E 400 279
P 395 - 400
390 - 395
278 - 279
277 - 278
385 - 390 276 - 277
I 380 - 385
375 - 380
275 - 276
274 - 275
370 - 375 273 - 274
365 - 370 272 - 273
360 - 365 271 - 272
355 - 360 270 - 271
350 - 355 269 - 270
345 - 350 268 - 269
340 - 345 267 - 268
335 - 340 266 - 267
330 - 335 265 - 266
325 - 330 264 - 265
42.6 ( 5.0) 119.9 320 - 325 264.5 (5.0) 349.9 205.2 ( 1.0) 244.7 263 - 264
315 - 320 262 - 263
310 - 315 261 - 262
305 - 310 260 - 261
300 - 305 259 - 260
295 - 300 258 - 259
290 - 295 257 - 258
M 285 - 290
280 - 285
256 - 257
255 - 256
I 275 - 280
270 - 275
254 - 255
253 - 254
D 265 - 270
260 - 265
252 - 253
251 - 252
255 - 260 250 - 251
250 - 255 249 - 250
245 - 250 248 - 249
240 - 245 247 - 248
235 - 240 246 - 247
230 - 235 245 - 246
225 - 230 244 - 245
220 - 225 243 - 244
215 - 220 242 - 243
210 - 215 241 - 242
D 160 - 165
155 - 160
231 - 232
230 - 231
O 150 - 155
145 - 150
229 - 230
228 - 229
140 - 145 227 - 228
135 - 140 226 - 227
130 - 135 225 - 226
125 - 130 224 - 225
120 - 125 223 - 224
115 - 120 222 - 223
110 - 115 221 - 222
105 - 110 220 - 221
100 - 105 219 - 220
95 - 100 218 - 219
90 - 95 217 - 218
0.2 ( 5.0) 146.0 85 - 90 244.0 (5.0) 372.7 213.1 ( 1.0) 261.1 216 - 217
80 - 85 215 - 216
75 - 80 214 - 215
70 - 75 213 - 214
65 - 70 212 - 213
60 - 65 211 - 212
55 - 60 210 - 211
50 - 55 209 - 210
45 - 50 208 - 209
40 - 45 207 - 208
35 - 40 206 - 207
30 - 35 205 - 206
25 - 30 204 - 205
20 - 25 203 - 204
15 - 20 202 - 203
10 - 15 201 - 202
5 - 10 200 - 201
5 200
Fig. 9.35 Mixed-wall with central endocardial stimulation. Same format as Fig. 9.36 (Reproduced
with permission from [130])
312 9 Simulation Studies of Cardiac Bioelectrical Activity
E 395 280
P 390 - 395
385 - 390
279 - 280
278 - 279
380 - 385 278 - 278
I 375 - 380
370 - 375
278 - 278
277 - 278
365 - 370 276 - 277
360 - 365 276 - 276
355 - 360 276 - 276
350 - 355 275 - 276
345 - 350 274 - 275
340 - 345 274 - 274
335 - 340 274 - 274
330 - 335 273 - 274
325 - 330 272 - 273
320 - 325 272 - 272
42.4 ( 5.0) 119.9 315 - 320 289.1 (5.0) 366.2 244.5 ( 0.5) 250.9 272 - 272
310 - 315 271 - 272
305 - 310 270 - 271
300 - 305 270 - 270
295 - 300 270 - 270
290 - 295 269 - 270
285 - 290 268 - 269
M 280 - 285
275 - 280
268 - 268
268 - 268
I 270 - 275
265 - 270
267 - 268
266 - 267
D 260 - 265
255 - 260
266 - 266
266 - 266
250 - 255 265 - 266
245 - 250 264 - 265
240 - 245 264 - 264
235 - 240 264 - 264
230 - 235 263 - 264
225 - 230 262 - 263
220 - 225 262 - 262
215 - 220 262 - 262
210 - 215 261 - 262
205 - 210 260 - 261
D 155 - 160
150 - 155
256 - 256
255 - 256
O 145 - 150
140 - 145
254 - 255
254 - 254
135 - 140 254 - 254
130 - 135 253 - 254
125 - 130 252 - 253
120 - 125 252 - 252
115 - 120 252 - 252
110 - 115 251 - 252
105 - 110 250 - 251
100 - 105 250 - 250
95 - 100 250 - 250
90 - 95 249 - 250
85 - 90 248 - 249
0.2 ( 5.0) 145.7 80 - 85 258.5 (5.0) 389.8 244.1 ( 0.5) 258.4 248 - 248
75 - 80 248 - 248
70 - 75 247 - 248
65 - 70 246 - 247
60 - 65 246 - 246
55 - 60 246 - 246
50 - 55 245 - 246
45 - 50 244 - 245
40 - 45 244 - 244
35 - 40 244 - 244
30 - 35 243 - 244
25 - 30 242 - 243
20 - 25 242 - 242
15 - 20 242 - 242
10 - 15 241 - 242
5 - 10 240 - 241
0- 5 240 - 240
0 240
Fig. 9.36 H-wall with central endocardial stimulation. Activation (ACTI, first column), repolar-
ization (REPO, second column), APD (third column) patterns on the epicardium (EPI, first row),
midwall (MID, second row), endocardium (ENDO, third row), diagonal section (fourth row). Below
each panel are reported the minimum, stepsize (in brackets), and maximum value of the displayed
map (Reproduced with permission from [130])
9.3 Heterogeneous Cardiac Tissue 313
Table 9.6 Average correlation coefficients of repolarization times and APD on r, ; -layers
Endo stim. Epi stim.
Average CC on Average CC on
r-layer -layer -layer r-layer -layer -layer
H-wall,3-wall REPO 0.997 0.838 0.841 0.998 0.842 0.565
APD 0.590 0.460 0.442 0.699 0.386 0.366
H-wall,W-wall REPO 0.999 0.913 0.911 0.999 0.914 0.657
APD 0.830 0.273 0.275 0.873 0.373 0.376
3-wall,mixed-wall REPO 0.853 0.890 0.997 0.882 0.887 0.997
APD 0.146 0.673 0.999 0.150 0.668 0.999
Apex-base-wall, REPO 0.998 0.854 0.829 0.999 0.904 0.562
mixed-wall APD 0.993 0.745 0.455 0.992 0.739 0.375
E 395 279
P 390 - 395
385 - 390
278 - 279
277 - 278
380 - 385 276 - 277
I 375 - 380
370 - 375
275 - 276
274 - 275
365 - 370 273 - 274
360 - 365 272 - 273
355 - 360 271 - 272
350 - 355 270 - 271
345 - 350 269 - 270
340 - 345 268 - 269
335 - 340 267 - 268
330 - 335 266 - 267
325 - 330 265 - 266
320 - 325 264 - 265
35.9 ( 5.0) 159.1 315 - 320 294.5 (5.0) 385.8 224.9 ( 1.0) 274.0 263 - 264
310 - 315 262 - 263
305 - 310 261 - 262
300 - 305 260 - 261
295 - 300 259 - 260
290 - 295 258 - 259
285 - 290 257 - 258
M 280 - 285
275 - 280
256 - 257
255 - 256
I 270 - 275
265 - 270
254 - 255
253 - 254
D 260 - 265
255 - 260
252 - 253
251 - 252
250 - 255 250 - 251
245 - 250 249 - 250
240 - 245 248 - 249
235 - 240 247 - 248
230 - 235 246 - 247
225 - 230 245 - 246
220 - 225 244 - 245
215 - 220 243 - 244
210 - 215 242 - 243
205 - 210 241 - 242
D 155 - 160
150 - 155
231 - 232
230 - 231
O 145 - 150
140 - 145
229 - 230
228 - 229
135 - 140 227 - 228
130 - 135 226 - 227
125 - 130 225 - 226
120 - 125 224 - 225
115 - 120 223 - 224
110 - 115 222 - 223
105 - 110 221 - 222
100 - 105 220 - 221
95 - 100 219 - 220
90 - 95 218 - 219
85 - 90 217 - 218
0.1 ( 5.0) 161.9 80 - 85 270.0 (5.0) 388.3 226.4 ( 1.0) 278.3 216 - 217
75 - 80 215 - 216
70 - 75 214 - 215
65 - 70 213 - 214
60 - 65 212 - 213
55 - 60 211 - 212
50 - 55 210 - 211
45 - 50 209 - 210
40 - 45 208 - 209
35 - 40 207 - 208
30 - 35 206 - 207
25 - 30 205 - 206
20 - 25 204 - 205
15 - 20 203 - 204
10 - 15 202 - 203
5 - 10 201 - 202
0- 5 200 - 201
0 200
Fig. 9.37 Apex-base-wall, Purkinje stimulation. Same format as Fig. 9.36 (Reproduced with
permission from [130])
9.4 QRS Complex and T Wave Morphology in Electrograms 315
Our main finding is that for cardiac tissue with normal cell coupling the
anisotropic diffusion current (electrotonic modulation) is not sufficient to mask the
intrinsic repolarization differences (both transmural and apex-to-base) between each
myocyte and its neighboring cells. Thus, in our orthotropic ellipsoidal wall, the
complex 3D electrotonic modulation of APDs does not fully mix the effects of
the transmural and apex-to-base heterogeneity. The intrinsic spatial heterogeneity
of the APDs is unmasked in the modulated APD patterns only in the appropriate
transmural or intramural sections.
The factors determining the shape and polarity of the T wave in electrocardiograms
(ECGs) are still a matter of debate. According to the classical theory based on
the double layer source model (see e.g. [538] and the classical references therein),
an activation sequence starting from the endocardium and proceeding towards the
epicardium generates ECG waveforms at leads facing the epicardium with an R
wave. By applying the heart surface model (see e.g. [483, 538]), it is generally
believed that if the repolarization sequence follows the activation one, then a
negative T wave is expected, due to opposite direction of the transmembrane
potential gradients during the downstroke phase of the action potential. On the
contrary, measured human Einthoven leads II exhibit QRS complex and T wave with
the same positive polarity. In order to explain this discrepancy, a different transmural
repolarization sequence within the ventricular wall has been proposed, where the
epicardium recovers before the endocardium based on the assumption that the action
potential duration (APD) of epicardial cells is considerably shorter than the APD of
endocardial cells. Experimental evidence of transmural APD heterogeneity and of
transmural repolarization gradients has been established in [571] for isolated cells
and in vitro wedge preparations of canine left ventricles. Computational studies
on isotropic [95, 148, 176, 205] or anisotropic cardiac tissue [58], with transmural
APD cell heterogeneity, have been able to reproduce concordant R and T waves.
Nevertheless, a repolarization sequence directed from epicardium to endocardium
has never been observed in in vivo animal models; see [133, 183, 258, 517]. In
alternative to transmural repolarization gradients, apico-basal APD heterogeneities
have been proposed as a determinant of the T wave polarity. The aim of this section
is to provide some insight on how the tissue anisotropy, cellular APD heterogeneities
and shape of excitation wavefronts influence the polarity of the T wave in unipolar
and bipolar ECGs. The study is based on three-dimensional simulations of the
entire depolarization and repolarization phases of the propagating action potential,
using the insulated anisotropic Monodomain model in a slab of cardiac tissue. Then
the primary source is used to compute the extracardiac potential in an unbounded
homogeneous conducting medium.
316 9 Simulation Studies of Cardiac Bioelectrical Activity
al .x/Due1 cos ˛.r/Cue2 sin ˛.r/; at .x/Due3 ; an .x/Due1 sin ˛.r/ ue 2 cos ˛.r/
˛.r/D 23 .1 r/ 4 ; with r 2 Œ0; 1:
Table 9.7 Parameter calibration for modeling the transmural heterogeneities (left) and conductiv-
ity coefficients in mS cm1 (right)
Slab type H-slab 3-slab Conductivity coefficients
Number of layers 1 3 Intra Extra
Layer thickness (cm) Endo Mid Epi l 3 2
1 0.33 0.34 0.33 t 0:3152 1:3514
fact_IK 1 2.62 1.95 2.88 n 0:0315 0:6757
APD (ms) 266 235 272 225 0 D 6
9.4 QRS Complex and T Wave Morphology in Electrograms 317
PLANE L1
50
TRANSMEMBRANE POTENTIAL (mV)
2 cm
0
EPI
M
1 cm
−50
5 cm ENDO EPI
2 cm
ENDO
−100
PLANE L 0 50 100 150 200 250 300
2
TIME (ms)
Fig. 9.38 Left: cardiac slab and location of the 3 3 transmural ECG electrode array 2 cm above
the epicardial surface (plane L1 ) and 2 cm below the endocardial surface (plane L2 ). Right:
Epicardial, Midmyocardial and Endocardial action potentials in the transmural heterogeneous
3-slab calibration
318 9 Simulation Studies of Cardiac Bioelectrical Activity
DHS D ti I and DAS D .li ti /al .x/aTl .x/ C .ni ti /an .x/aTn .x/:
This splitting of Di induces an analogous splitting of the cardiac sources into the
so called double layer or heart surface (HS) source model (see [483, 538]) and into
a residual anisotropic source (AS). Consequently, the source term is decomposed
as Di rv D DHS rv C DAS rv and since u depends linearly on Di rv in Eq. (9.4),
the full unipolar ECG is decomposed as the sum of an isotropic (heart surface)
and anisotropic components u D uHS C uAS . Given two points xepi 2 cm above
the epicardium (plane L1 ) and xendo 2 cm below the endocardium (plane L2 ), we
introduce also the transmural bipolar ECG given by .xepi ; xendo ; t/ D u.xepi ; t/
u.xendo ; t/: Using the same splitting as before, we also decompose the full bipolar
ECG into the HS component HS .xepi ; xendo ; t/ and the anisotropic component
AS .xepi ; xendo ; t/.
Numerical methods. The cardiac domain ˝H considered is a cartesian slab of
dimensions 5 5 1 cm3 modeling a portion of the left ventricle; see Fig. 9.38-left.
In all computations, a structured grid of 500 500 100 hexahedral isoparametric Q1
elements of size h D 0:1 mm is used in space. The time discretization is based on
an Euler Imex method.
S1 S2
a TRANSMURAL SECTION b TRANSMURAL SECTION
80 60
ACTI
ACTI
60
40
40
20 20
H−slab H−slab
REPO90
REPO90
300 300
250 250
258.50 330.79 4.00 258.24 315.85 4.00
3−slab 3−slab
REPO90
REPO90
300 300
250 250
244.66 326.50 4.00 244.49 312.16 4.00
Fig. 9.39 (a): Single site stimulation (S1 ), isochrones of activation time ACTI (first row) and
repolarization time REPO90 (second and third row) on a transmural diagonal section of H-slab,
3-slab. (b): Multiple sites stimulation (S2 ), same format as in (a)
0 0
mV
mV
−0.5 −0.5
−1 −1
−1.5 −1.5
0 100 200 300 400 0 100 200 300 400
ms ms
Fig. 9.40 H-slab: Unipolar ECGs at site 5 on plane L2 (site 5 is the central one in the 3 3 array,
see Fig. 9.41). (a): Single site stimulation (S1 ). (b): Multiple sites stimulation (S2 ). Continuous
line: full unipolar ECG. Dashed line: HS component. Dashed-Dotted line: AS component
This fact indicates that the full unipolar ECGs are dominated by their anisotropic
AS components. The HS components have little influence on the full ECGs, since
their positive T waves reinforce those of the full ECGs but their R waves at almost
all sites on plane L1 are absent in the QRS complex of the full ECGs. Analogous
morphological considerations hold for the 3-slab (Fig. 9.42), although with slight
differences in magnitude.
S2 : unipolar ECGs. The isochrones of ACTI and REPO90 on a transmural
section of the tissue are reported in Fig. 9.39b. The excitation wavefronts are larger
than those elicited by the ectopic stimulation. On plane L2 , the morphology of the
unipolar ECGs is the same as in the ectopic case described before; see Fig. 9.40b. On
plane L1 , all the AS components have Q waves and positive T waves, while the HS
320 9 Simulation Studies of Cardiac Bioelectrical Activity
a S1 b S2
7 8 9 7 8 9
0.5 0.5
mV
mV
−1 −0.5
4 5 6 4 5 6
0.5 0.5
mV
mV
−1 −0.5
1 2 3 1 2 3
0.5 0.5
mV
mV
−1 −0.5
0 400 0 400 0 400 0 400 0 400 0 400
ms ms ms ms ms ms
0 0
mV
mV
−0.5 −0.5
−1 −1
−1.5 −1.5
0 100 200 300 400 0 100 200 300 400
ms ms
Fig. 9.41 H-slab: Unipolar ECGs on plane L1 . Continuous line: full unipolar ECG. Dashed line:
HS component. Dashed-Dotted line: AS component (only in zoom). (a): Single site stimulation
(S1 ). (b): Multiple sites stimulation (S2 ). (c, d): zoom at site 5
components exhibit R waves and negative (H-slab, Fig. 9.41b, d) or positive (3-slab,
Fig. 9.42b, d) T waves. Nevertheless, the larger magnitude of the AS components
prevail in the full unipolar ECGs and they still exhibit a positive T wave and QRS
complex entirely negative or with RS configuration (small R and large S wave) and
positive T wave. Therefore, the full unipolar ECGs are still dominated by their AS
components.
S1 : bipolar ECGs. Unlike unipolar waveforms, the bipolar ECGs (Figs. 9.43
and 9.44a, c) appear to be dominated by their HS components, all with R and S
waves and biphasic (for H-slab) or positive (for 3-slab) T waves. Since the AS
components of the unipolar ECGs on planes L1 and L2 are almost identical, the
bipolar AS component has a very small magnitude in comparison with the bipolar
HS component, hence they have a negligible influence on the full bipolar ECGs.
The negative minimum of the biphasic T wave appearing in the bipolar ECGs in the
H-slab case is due to the T peak of the unipolar HS components at sites on plane
L2 preceding the T peak of the symmetric sites on plane L1 .
9.4 QRS Complex and T Wave Morphology in Electrograms 321
S1 S2
a 7 8 9
b 7 8 9
0.5 0.5
mV
mV
−1 −0.5
4 5 6 4 5 6
0.5 0.5
mV
mV
−1 −0.5
1 2 3 1 2 3
0.5 0.5
mV
mV
−1 −0.5
0 400 0 400 0 400 0 400 0 400 0 400
ms ms ms ms ms ms
0 0
mV
mV
−0.5 −0.5
−1 −1
−1.5 −1.5
0 100 200 300 400 0 100 200 300 400
ms ms
Fig. 9.42 3-slab: Unipolar ECGs on plane L1 . Same format as Fig. 9.41. (a) S1 . (b) S2 . (c) S1 ,
zoom at site 5. (d) S2 , zoom at site 5
S2 : bipolar ECGs. Bipolar ECGs now display a QRS complex with a large
R wave and negative T wave (Figs. 9.43 and 9.44b, d), similarly to their HS
components. The bipolar AS components have little influence because of their much
smaller magnitude and therefore, as in the ectopic case, the full bipolar ECGs are
dominated by their HS components.
Discussion of the results. By means of three-dimensional simulations based
on the anisotropic Monodomain model, we have studied the influence of tissue
anisotropy, cellular APD heterogeneities and the shape of the excitation wavefront
on the T wave polarity. The simulation results have shown that: (i) unipolar ECGs
exhibit a positive T wave mainly determined by the anisotropy of the cardiac
tissue, irrespective of cellular APD heterogeneities and the shape of the excitation
wavefront; (ii) on the other hand, bipolar ECGs are mainly determined by their
isotropic component and their T wave turns out to be positive only for a single
stimulation in presence of transmural APD heterogeneity, while it becomes always
negative in case of multiple stimulation generating large activation wavefronts,
regardless of the considered cellular APD heterogeneities. We remark that in our
ECG computation, we have used an idealized cardiac geometry taking into account
322 9 Simulation Studies of Cardiac Bioelectrical Activity
S1 S2
a 7 8 9 b 7 8 9
0.5 1.5
mV
mV
−0.5 −0.5
4 5 6 4 5 6
0.5 1.5
mV
mV
−0.5 −0.5
1 2 3 1 2 3
0.5 1.5
mV
mV
−0.5 −0.5
0 400 0 400 0 400 0 400 0 400 0 400
ms ms ms ms ms ms
1 1
mV
mV
0.5 0.5
0 0
−0.5 −0.5
0 100 200 300 400 0 100 200 300 400
ms ms
Fig. 9.43 H-slab: Bipolar ECGs on plane L1 . Continuous line: full unipolar ECG. Dashed line:
HS component. Dashed-Dotted line: AS component (only in zoom). (a): Single site stimulation
(S1 ). (b): Multiple sites stimulation (S2 ). (c, d): zoom at site 5
anisotropic cardiac sources but considering the widely used assumptions of the same
isotropic properties for bulk tissue and for the surrounding infinite extracardiac
medium. Previous studies [117, 119] have shown that including anisotropic bulk
tissue affects the ECGs magnitude far away from the cardiac surface but it does not
alter their QRS morphology. The extension of our conclusions on T wave polarity
to real 12-lead ECGs will require considering realistic geometries for both cardiac
tissue and torso, as well as the anisotropy of the bulk tissue. Nevertheless, the real
12-lead ECGs are far-field signals and we expect that the full ECG and the isotropic/
anisotropic ECG components will be smoothed out due to the greater distance from
the active tissue.
9.5 Extracellular Markers of Excitation and Repolarization Times 323
a S1 b S2
7 8 9 7 8 9
0.5 1.5
mV
mV
−0.5 −0.5
4 5 6 4 5 6
0.5 1.5
mV
mV
−0.5 −0.5
1 2 3 1 2 3
0.5 1.5
mV
mV
−0.5 −0.5
0 400 0 400 0 400 0 400 0 400 0 400
ms ms ms ms ms ms
1 1
mV
mV
0.5 0.5
0 0
−0.5 −0.5
0 100 200 300 400 0 100 200 300 400
ms ms
The time evolution of the cardiac bioelectric activity is described by the time
and spatial distribution of the intracellular and extracellular potential fields
ui .x; t/; ue .x; t/, both dependent on a chosen reference potential. By taking the
difference between the intra- and extracellular potentials, we obtain the distribution
of the transmembrane potential v.x; t/ D ui .x; t/ ue .x; t/; that is independent of
the reference potential. During a heart beat, the transmembrane potential of each
myocardial cell undergoes a time variation called transmembrane action potential
(TAP), characterized by a fast upstroke of about 1 ms and a subsequent slower
downstroke of about 50 ms, associated with the activation and recovery phases,
respectively, also called depolarization and repolarization phases.
Maps of the activation and recovery sequences provide important information for
identifying normal heart activity and cardiac arrhythmias that are often associated
with abnormal recovery times and action potential durations. Methods for determin-
ing activation and recovery times on the epicardial and endocardial surfaces, as well
as transmurally (i.e. in the thickness of the ventricular wall), are essential tools for
understanding the recovery process in normal and pathological conditions at both
experimental and clinical levels.
324 9 Simulation Studies of Cardiac Bioelectrical Activity
While methods for determining activation sequences from direct leads (recorded
directly from the heart) are well established, the assessment of local recovery times
is more difficult, due to the larger time and space scales involved in the recovery
process. The activation time is generally defined as the time instant when the TAP
exhibits the fastest upstroke and the unipolar electrogram (EG) shows the fastest
downstroke (i.e. the intrinsic deflection) during the QRS complex; see e.g. [408]
for a recent update. The assessment of local recovery times is based on indexes
associated with the TAP downstroke phase and with the upstroke of the T wave in
the EG, see Fig. 9.45.
Widely used TAP recovery markers are the time of fastest repolarization RTtap ,
defined as the instant of TAP fastest downstroke during the recovery phase, and the
late recovery time RT90tap , defined as the instant when the TAP reaches 90 % of its
resting value during the downstroke phase.
The potential drop across the membrane can be measured only from a few sites
in a given preparation by microelectrodes (micropipette). Therefore, transmembrane
potentials cannot be simultaneously recorded from hundreds of sites in vivo. Optical
50 50
∂ v
t
0 0
v (mV)
−50 −50
APD APD90
AT RT ATtap RT90
tap
tap tap
−100 −100
0 100 200 300 0 100 200 300
40 40
∂ u ∂ttue
t e
ue (mV)
20 20
ARI
eg ARI90eg
0 0
AT RT AT RT90
eg eg eg eg
−20 −20
0 100 200 300 0 100 200 300
15 15
∂thmap
5 5
hmap (mV)
−5 −5
ARI ARI90
−15 hmap −15 hmap
AT RT AThmap RT90
hmap
hmap hmap
−25 −25
0 100 200 300 0 100 200 300
time (ms) time (ms)
Fig. 9.45 Transmembrane action potential (v), unipolar electrogram (ue ) and hybrid monophasic
action potential (hmap) waveforms and the associated activation and recovery time markers
(Reproduced with permission from [461])
9.5 Extracellular Markers of Excitation and Repolarization Times 325
is related to the moment of fastest repolarization, while the latter indicates the
ending phase of repolarization. The TAP activation and repolarization time markers
(sketched graphically in Fig. 9.45) are defined as follows:
• ATtap .x/ D argmax f@t v.x; t/; t 2 upstrokeg,
i.e. the instant of maximum time derivative of v.x; t/ during the upstroke;
• RTtap .x/ D arg min f@t v.x; t/; t 2 downstrokeg,
i.e. the instant ˚of minimum time derivative of v.x; t/ during the downstroke;
• RT90tap .x/ D t W v.x; t/ D 0:9vr during downstroke ,
i.e. the instant when v.x; t/ reaches 90 % of its resting value vr during down-
stroke.
The TAP activation and repolarization markers ATtap , RTtap and RT90tap are the gold
standards for determining cardiac activation and repolarization times.
Unipolar electrograms (EG) waveforms and markers. A typical morphology
of an electrogram (EG), given by the time course of ue .x; t/ at a given point x 2 ˝H ,
is displayed in the mid panel of Fig. 9.45. The excitation phase is related to the
QRS complex, while the faster terminal repolarization occurs during the T wave.
This wave has positive and negative polarity for sites that repolarize early and late,
respectively, while it has a biphasic waveform for intermediate repolarization states.
A firmly established marker for the activation time from the EGs is the instant ATeg
of minimum derivative during the QRS complex; see e.g. [408, 493]. The most
widely accepted EG repolarization marker is the time RTeg of maximum upslope
during the T wave, see e.g. [221, 566]. Recently, we have proposed an EG marker
for late repolarization based on the time RT90eg of minimum second derivative of
the EG waveform during the T wave, see [460]. Note that the second time derivative
of electrograms has been used in [187, 188, 311] in order to define the end of the T
wave. The EG activation and repolarization time markers (sketched graphically in
Fig. 9.45) are defined as follows:
• ATeg .x/ D arg min f@t ue .x; t/; t 2 QRS complexg,
i.e. the instant of minimum time derivative of ue .x; t/ during the QRS complex;
• RTeg .x/ D argmax f@t ue .x; t/; t 2 T waveg,
i.e. the instant of maximum time derivative of ue .x; t/ during the T wave;
• RT90eg .x/ D arg min f@t t ue .x; t/; t 2 T waveg,
i.e. the instant of minimum second time derivative of ue .x; t/ during the T wave.
Hybrid monophasic action potential (HMAP) waveforms and markers. The
extracellular waveforms, recorded by Weissenburger et al. [553], are the potential
difference between a fixed extracellular recording within a permanently depolarized
region and a generic exploring site. These signals can be obtained directly from
the heart or computed as the difference between the unipolar EG from a fixed
permanently depolarized site and the unipolar EG from a generic exploring site.
These computed signals are obviously independent of the (identical) reference
potential used for the two unipolar EGs. We call these extracellular signals, hybrid
monophasic action potentials (HMAPs) to distinguish them from the classical close
bipolar monophasic action potentials (MAP) of Franz [181], where the exploring
328 9 Simulation Studies of Cardiac Bioelectrical Activity
and permanently depolarized sites are always very close to each other. HMAPs
are contaminated by far-field effects, while MAPs, being close bipolar signals,
are relatively insensitive to far-field effects. A typical morphology of an HMAP
waveform is displayed in the bottom panel of Fig. 9.45. The HMAP signal exhibits
multiple upstroke phases followed by the appearance of a monophasic component.
The fastest upstroke is associated with the activation of the exploring site and the
others with the depolarization of sites around the boundary of the PD volume. The
HMAP activation and recovery time markers are defined in the same way as the
TAP markers:
• AThmap .x/ D argmax f@t hmap.x; t/; t 2 upstrokeg,
i.e. the instant of maximum time derivative of hmap.x; t/ during the upstroke;
• RThmap .x/ D arg min f@t hmap.x; t/; t 2 downstrokeg,
i.e. the instant ˚of minimum time derivative of hmap.x; t/ during the downstroke;
• RT90hmap .x D t W hmap.x; t/ D 0:9vr during downstroke ,
i.e. the instant when hmap.x; t/ reaches 90 % of its resting value during the
downstroke.
We recall that the information content of the HMAP waveform is controversial,
see e.g. [182, 265, 285, 347, 542, 553]. In particular, the reliability of recovery time
markers derived from HMAP signals has been questioned, see [128, 135, 560].
As a byproduct of the AT and RT markers defined above, the following action
potential duration (APD) and Activation–Recovery Interval (ARI) markers are
defined.
APD markers from TAP waveforms.
APD.x/ D RTtap .x/ ATtap .x/; APD90.x/ D RT90tap .x/ ATtap .x/:
ARIeg .x/ D RTeg .x/ ATeg .x/; ARIhmap .x/ D RThmap .x/ AThmap .x/I
ARIeg has been widely used in estimating the APD in experiments on animals
and humans, see e.g. [135, 221, 328] and [84, 194, 517], respectively.
We recall that the excitation process is a self-sustained propagating phenomenon
and the isochrones of the activation time define a propagating excitation front.
Conversely, the repolarization process is not a matter of conduction but it is
a synchronization phenomenon, see e.g. [133]. Nevertheless, the isochrones of
successive repolarization times define a sequence of repolarization fronts and even
if these fronts are not propagating in a classical sense, for convenience of notation
we will use in the following the terms acceleration, deceleration, collision regarding
the sequence of repolarization fronts sweeping the myocardial volume.
9.5 Extracellular Markers of Excitation and Repolarization Times 329
Fig. 9.46 Left: cardiac slab ˝H , PD sites, ischemic region, transmural needles. Right: needle
locations on the epicardial plane and their row and column indexes (Reproduced with permission
from [461])
330 9 Simulation Studies of Cardiac Bioelectrical Activity
Table 9.8 Parameter calibration for modeling the transmural heterogeneities in the three cardiac
slabs H-slab, 3-slab, W-slab
H-slab 3-slab W-slab
# of layers 1 3 4
Endo Mid Epi Endo Sub-endo Mid Epi
Thickness (cm) 0.48 0.16 0.16 0.16 0.058 0.096 0.254 0.072
fact_IK 1 2.62 1.95 2.88 2.71 1.95 2.47 2.88
APD (ms) 266 235 272 225 232 272 242 225
20 200 210 20
40
30 200 210 200
20
150 200 219.95 280.83 2.00 150
10.75 64.61 2.00 233.08 296.08 2.00 221.15 232.79 0.50
60 220 APD
250 215
ENDO
220
40
210
200
20 205 210
Fig. 9.47 3-slab, PD3. Activation (ACTI), repolarization (REPO) and action potential duration
(APD) isochrone lines, computed with space resolution h D 0:1 mm, related to the transmembrane
markers ATtap .x/, RT90tap .x/, APD90.x/. Left panel: intramural sections (epicardium, midwall,
endocardium). Right panel: transmural diagonal section A perpendicular to the epicardium (see
Fig. 9.46). Maximum, minimum and step of the displayed map are reported below each panel
(Reproduced with permission from [461])
transmural pattern with a maximum in the midmyocardial regions, where cells with
a longer intrinsic APD are located.
Postprocessing. We saved the extracellular and the intracellular potential wave-
forms ue .x; t/ and ui .x; t/ at the 12 12 13 locations of the multi-electrode array
described above. These waveforms for ue and for the byproduct v D ui ue are then
post-processed by computing the additional activation and repolarization markers
defined previously. The TAP markers are assumed to be the reference markers.
The EG and HMAP markers are compared with the reference TAP markers, i.e.:
ATeg and AThmap vs ATtap ; RTeg and RThmap vs RTtap ; RT90eg and RT90hmap vs
RT90tap ; ARIeg and ARIhmap vs APD; ARI90eg and ARI90hmap vs APD90.
In order to compare these large data sets, we then compute, for vectors X; Y of
length n D 1 872, the following quantities, reported in the Tables of the Results:
P
• mean.X / D n1 n1 X pi D the average of X ;
• std.X / D kX0 k= n D the standard deviation of X , where X0 D X
mean.X /1;
• corr.X; Y / D .X0 =kX0 k; Y0 =kY0 k/ D the correlation coefficient between X
and Y ;
• mrd D mean.jX Y j/=mean.jY j/ D relative mean error of X with respect to Y ,
where k k; .; / denote the Euclidean norm and scalar product.
332 9 Simulation Studies of Cardiac Bioelectrical Activity
Effects of the PD location on the extracellular potential field. Table 9.9 reports
the comparison between the EG and HMAP repolarization markers and the refer-
ence TAP markers (RTeg vs RTtap , RThmap vs RTtap , RT90eg vs RT90tap , RT90hmap
vs RT90tap ). The sites with discrepancies larger than 30 ms are excluded but, in
each case, the points excluded are less than 3. All the extracellular estimates
show a high reliability as confirmed by a correlation coefficient always greater
than 0.95. The RT90eg marker exhibits the best performance, followed by RTeg
and then by RThmap , with average absolute discrepancies ranging between 1:5 and
3:2 ms. Despite the high correlation coefficient, the RT90hmap estimate exhibits
a larger average discrepancy ranging between 3:1 and 11:3 ms depending on the
location of the PD volume. The relevant increase of both average discrepancies and
standard deviations for all markers with respect to the case involving only isotropic
waveforms (see [461]) indicates that the anisotropic component of the extracellular
potential ue is the major determinant of the discrepancies between EG and TAP
repolarization markers.
The EG markers are weakly influenced by the location of the PD volume.
The global performance of the RThmap marker exhibits a weak dependence on the
position of the PD volume. Conversely, the performance of the RT90hmap marker,
in terms of average discrepancy not of correlation, is strongly dependent on the
PD volume location. This strong performance dependence on the location of the
PD volumes may be related to the different morphology of the EGs in the PD
Table 9.9 RT markers discrepancies, 3-slab, PD volume in PD1, PD2 and PD3. EG and HMAP
markers computed from extracellular waveforms (ue and hmap) versus TAP markers. meanjX
Y j D average of jX Y j; meanjY j D average of jY j; mrd D meanjX Y j/meanjY j D relative
mean discrepancy; std D standard deviation of jX Y j; corr.X; Y / D correlation coefficient
between X and Y
Mean Mean Corr
X Y mrd jX Y j jY j Std .X; Y /
PD1 RTeg RTtap 0.76e2 1.86 244.41 1.72 0.99
RThmap RTtap 1.30e2 3.17 244.43 3.50 0.96
RT90eg RT90tap 0.63e2 1.59 252.94 1.79 0.99
RT90hmap RT90tap 4.46e2 11.29 252.92 6.28 0.97
PD2 RTeg RTtap 0.77e2 1.91 246.87 1.63 0.98
RThmap RTtap 1.23e2 3.04 246.78 3.54 0.95
RT90eg RT90tap 0.64e2 1.64 255.58 1.95 0.99
RT90hmap RT90tap 1.21e2 3.08 255.26 2.51 0.96
PD3 RTeg RTtap 0.78e2 1.92 245.40 1.65 0.99
RThmap RTtap 0.97e2 2.39 245.42 2.23 0.98
RT90eg RT90tap 0.64e2 1.63 254.01 1.75 0.99
RT90hmap RT90tap 2.07e2 5.26 254.02 2.45 0.98
9.5 Extracellular Markers of Excitation and Repolarization Times 333
40
35
30
25
20 PD3
mV
PD2
15
10
PD1
5
−5
0 50 100 150 200 250 300 350
ms
Fig. 9.48 Shifted extracellular waveforms ue .xPDR; t / in the PD volume for xPD 2
PD1; PD2; PD3 (see Fig. 9.46) and CR.t / D j˝1H j ˝H v.x; t / dx (dashed line) for the 3-slab
(Reproduced with permission from [461])
334 9 Simulation Studies of Cardiac Bioelectrical Activity
Table 9.10 Recovery time (RT) markers discrepancies for H-slab, 3-slab, W-slab, MI-slab and
SI-slab with PD volume in PD3. Comparison between the RT markers (RTeg vs RTtap , RThmap vs
RTtap , RT90eg vs RT90tap , RT90hmap vs RT90tap ): meanjX Y j D average of jX Y j; meanjY j D
average of jY j; mrd D meanjX Y j/meanjY j D relative mean discrepancy; std D standard
deviation of jX Y j; corr.X; Y / D correlation coefficient between X and Y
Mean Mean Corr
X Y mrd jX Y j jY j Std .X; Y /
H-slab RTeg RTtap 0.84e2 2.14 253.22 1.98 0.99
RThmap RTtap 0.72e2 1.83 253.15 1.60 0.99
RT90eg RT90tap 0.36e2 0.95 260.32 0.86 0.99
RT90hmap RT90tap 1.17e2 3.04 260.38 2.15 0.98
3-slab RTeg RTtap 0.78e2 1.92 245.40 1.65 0.99
RThmap RTtap 0.97e2 2.39 245.42 2.23 0.98
RT90eg RT90tap 0.64e2 1.63 254.01 1.75 0.99
RT90hmap RT90tap 2.07e2 5.26 254.02 2.45 0.98
W-slab RTeg RTtap 0.56e2 1.39 248.25 1.19 0.99
RThmap RTtap 0.84e2 2.09 247.52 2.18 0.98
RT90eg RT90tap 0.41e2 1.04 255.66 1.09 0.99
RT90hmap RT90tap 2.00e2 5.10 255.79 1.99 0.99
MI-slab RTeg RTtap 0.86e2 2.16 250.76 2.23 0.99
RThmap RTtap 0.77e2 1.94 250.63 2.22 0.99
RT90eg RT90tap 0.51e2 1.31 258.92 1.91 0.99
RT90hmap RT90tap 1.19e2 3.07 258.95 2.24 0.98
SI-slab RTeg RTtap 0.79e2 2.01 252.75 1.67 0.99
RThmap RTtap 0.75e2 1.90 252.24 1.80 0.99
RT90eg RT90tap 0.52e2 1.34 258.86 2.19 0.99
RT90hmap RT90tap 1.14e2 2.95 258.94 2.13 0.98
Fig. 9.49, reporting the regression plots of RTeg vs RTtap , RThmap vs RTtap , RT90eg vs
RT90tap and RT90hmap vs RT90tap on all five simulations. In terms of accuracy, the
RT90eg marker gives the best performance, with a relative average discrepancy of at
most 0.64e2 and an average absolute discrepancy of at most 1:63 ms. The RTeg and
RThmap , as estimates of RTtap , show comparable global performance, because both
the absolute and relative averages of jRTeg –RTtap j and jRThmap –RTtap j are about 2 ms
and 0.8e2 with a range 0.56e2 0.97e2. As previously observed only for the
3-slab, despite a global correlation coefficient of 0:98, the RT90hmap marker exhibits
larger absolute (relative) average discrepancies, ranging between 3:04 (1.17e2)
and 5:3 ms (2.1e2), than the other EG markers.
These global performance findings are also confirmed by the distributions of
the discrepancies RTeg –RTtap , RThmap –RT90tap , RT90eg –RT90tap and RT90hmap –
RT90tap , reported in Fig. 9.50. These distributions clearly show that almost all
discrepancies range between 10 and 10 ms and most of them are confined within
5 and 5. Figure 9.51 reports the discrepancies with respect to the transmural
9.5 Extracellular Markers of Excitation and Repolarization Times 335
RT90eg
RTeg
160 170
160 300 170 310
RTtap RT90tap
RT90
RT
160 170
160 300 170 310
RT RT90
tap tap
Fig. 9.49 Regression lines of RTeg vs RTtap (first column), RT90eg vs RT90tap (second column),
RThmap vs RTtap (third column) and RT90hmap vs RT90tap (fourth column) for all the five slab types
H-slab, 3-slab, W-slab, MI-slab and SI-slab (Reproduced with permission from [461])
position of the electrodes for the 3-slab, i.e. the discrepancy variation over an
intramural plane.
For 3-slab, we found that the largest discrepancies are located in the midmyocar-
dial regions, where the two repolarization fronts, that started from the endocardium
and the epicardium, merge. An analogous consideration apply to the subendocardial
regions in W-slab. We refer to our recent paper [460] for a discussion of the locations
and origin of the RTeg and RT90eg markers largest discrepancies, as well as for a
study of the associated artifacts appearing in the repolarization sequences. Signs
of the presence of these discrepancies are the correlation coefficients of the APD
markers, reported in Table 9.11, which are always lower than those of the associated
RT markers, indicating a loss of precision of the extracellular RT markers. We
remark that in all simulations the global performance indexes (mrd, mean, std)
336 9 Simulation Studies of Cardiac Bioelectrical Activity
3000 3000
NUMBER OF SITES
2000 2000
1000 1000
0 0
−20 −10 0 10 20 −20 −10 0 10 20
RTeg − RTtap RT90eg − RT90tap
3000 3000
2000 2000
1000 1000
0 0
−20 −10 0 10 20 −20 −10 0 10 20
RThmap − RTtap RT90 − RT90
hmap tap
Fig. 9.50 Histograms of discrepancies RTeg RTtap (first column), RT90eg RT90tap (second
column), RThmap RTtap (third column) and RT90hmap RT90tap (fourth column) with 1 ms bins,
for all the five slab types H-slab, 3-slab, W-slab, MI-slab and SI-slab (Reproduced with permission
from [461])
reported in Table 9.10, except the correlation coefficients (corr), are the same for
the pairs ARIeg (ARIhmap ) vs APD and RTeg (RThmap ) vs RTtap , as well as for the
pairs ARI90eg (ARI90hmap ) vs APD90 and RT90eg (RT90hmap ) vs RT90tap , because
the relative and absolute discrepancies of jATeg ATtap j and jAThmap ATtap j are of
the order of 1.e4 and of 0.05 ms, respectively.
When considering exploring sites located a few mm from the PD volume, the
bipolar signal HMAP simulates the condition of the classical monophasic action
potential (MAP), see Franz [181], because the exploring and the PD reference
electrode are very close. In order to evaluate the markers performance at these
particular sites, we now consider in each simulation the six electrodes close to the
PD volume (within 2 mm distance), yielding a total amount of 30 electrodes, and
we compare the RT markers at these sites. The average indexes (mean, std, corr) are
(1.06, 0.60, 0.99) for jRThmap –RTtap j and (1.99, 0.56, 0.96) for jRT90hmap –RT90tap j,
to be compared with (3.84, 1.24, 0.99) for jRTeg –RTtap j and with (2.04, 1.70, 0.95)
for jRT90eg –RT90tap j, respectively.
9.5 Extracellular Markers of Excitation and Repolarization Times 337
20 20
10 10
RT90eg − RT90tap
− RTtap
0 0
eg
RT
−10 −10
−20 −20
ENDO MID EPI ENDO MID EPI
20 20
− RT90tap
10 10
tap
RThmap − RT
0 0
hmap
RT90
−10 −10
−20 −20
ENDO MID EPI ENDO MID EPI
Fig. 9.51 3-slab, PD3. Discrepancies RTeg RTtap (first column), RT90eg RT90tap (second col-
umn), RThmap RTtap (third column) and RT90hmap RT90tap (fourth column) (Reproduced with
permission from [461])
These results show that near the PD volume RThmap (RT90hmap ) is a better
estimate of RTtap (RT90tap ) than RTeg (RT90eg) and consequently confirms that the
downstroke phase of the HMAP yields a more accurate approximation of the TAP
downstroke phase. This fact is also confirmed by the comparison of a suitable scaled
and shifted TAP waveform with the HMAP waveform displayed in Fig. 9.52-top for
a site close to the PD volume. Figure 9.52-bottom shows the plots of TAP, EG and
HMAP waveforms at three endocardial sites, for the 3-slab in the PD3 case. The EGs
exhibit T waves with different polarity, i.e. positive, biphasic and negative. We have
338 9 Simulation Studies of Cardiac Bioelectrical Activity
50 50
0 0
−50 −50
−100 −100
0 100 200 300 0 100 200 300
40 40
20 20
0 0
−20 −20
−40 −40
0 100 200 300 0 100 200 300
RTeg − RT tap = 3.77 RT90eg − RT90 tap = −2.74
40 40
20 20
0 0
−20 −20
−40 −40
0 100 200 300 0 100 200 300
50 50
0 0
−50 −50
−100 −100
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
40 40
20 20
0 0
−20 −20
−40 −40
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
RT − RT = 0.17 (A) 1.03 (B) 2.23 (C) RT90 − RT90 = 2.23 (A) 0.51 (B) 1.03 (C)
eg tap eg tap
40 40
20 20
0 0
−20 −20
−40 −40
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
RThmap − RTtap= 1.71 (A) 1.37 (B) 0.86 (C) RT90 − RT90 = −5.48 (A) −3.77 (B) −3.60 (C)
hmap tap
Fig. 9.52 Top panel. 3-slab, PD3: TAP (first row), EG (second row) and HMAP (third row)
waveforms in the epicardial site located in the needle (2,12), see Fig. 9.46. The TAP waveform,
plotted in dashed line in the third row superimposed to the HMAP, is shifted in order to have
the same resting value of the HMAP waveform and it is scaled by a factor ˛. The vertical lines
indicate RTtap (first row, left), RTeg (second row, left) and RThmap (third row), RT90tap (first row,
right), RT90eg (second row, right) and RT90hmap (third row, right). Bottom panel. 3-slab, PD3: same
format as in the top panel for the waveforms on the three endocardial sites located in the needles
(2,2) (A), (6,6) (B) and (11,11) (C), see Fig. 9.46 (Reproduced with permission from [461])
We briefly recall the well-established mechanism at a cellular level (see [564] Ch. II
Fig. 2.13) of the potential changes during ST interval, evidenced first by Samson
and Sher[453], Prinzmetal et al.[403] and later confirmed by Kleber et al.[279]; see
340 9 Simulation Studies of Cardiac Bioelectrical Activity
also the updated surveys[257, 276] and the references therein. Due to the reduction
of APD, the transmembrane potential in the ischemic cells during the ST interval is
lower than in non-ischemic cells. This transmembrane potential difference produces
a local intracellular current of injury flowing from non-ischemic to ischemic cells,
associated with a local extracellular current flowing in opposite direction. This latter
current is the cause of an elevation of the extracellular potential (ST elevation)
recorded from the ischemic zone and a depression detected in the healthy part.
Thus, during the ST interval, at the tissue level we have a distribution of local
injury currents sources around the ischemic boundary. According to Holland and
Brooks[231], one way to extend the cellular mechanism to the ventricular wall
is based on assuming that the elementary injury sources can be model as: (i) a
dipole layer on the ischemic boundary, with dipole direction normal to this boundary
and with moment proportional to the transmembrane potential jump; (ii) the dipole
layer is embedded into an isotropic homogeneous and infinite conducting medium.
This modeling of the injury sources, known as solid angle theory, predicts at
the epicardial side an ST depression associated with the subendocardial ischemic
volume. This extension of the cellular mechanisms to a tissue level is questionable,
since the cardiac tissue exhibits a strong anisotropic structure and the injury currents
sources are embedded in rotational anisotropic intra and extracellular media.
We apply now the source split of the extracellular potential field described in
Chap. 5 to the particular case of the ST segment in the presence of an ischemic
subendocardial region. Denoting by ˝i the ischemic tissue, define i D @˝i ,
˙i D i \ ˙, and l D i n ˙i , where ˙ is the non-insulated boundary of
˝H , see Fig. 5.1. In the ST segment, the distribution of the transmembrane results
can be practically approximated by a piecewise constant distribution. Denoting
by vIr , vN
r the values of the transmembrane potential in the ischemic and normal
tissue, respectively, then v.x; t/ D vN r C .vr vr /˝i .x/, where ˝i denotes the
I N
vector to i pointing inside the ischemic domain ˝i and ıi .x/ is a Dirac measure
on i . Therefore, the injury current can be represented as an oblique dipolar layer
with moment given by
Jv D .vIr vN
r / Di .x/n.x/ ıi .x/:
Applying the split of the tensor Di given in (5.40), the oblique dipole layer Jv .x/ can
be viewed as the superposition of a conormal dipole with direction Dn, of a axial
dipole layer having the same direction of the fiber al , and of a cross-axial dipole
layer, since the matrix I al aTl is the projection matrix on the plane orthogonal to
al , i.e.
Jv D Jc C Ja C Jo :
9.6 Subendocardial Ischemia, ST Depression and Elevation 341
Here
Jc D ˛Dn
O ıi Ja D ˇO aTl n al ıi ; Jo D O I al aTl n ıi ;
with ˛O D .vIr vN O
r /.t n /=.t n /, ˇ D .vr vr /.l ˛l / and O D
i i I N i
.vr vr /.t ˛t / D .vr vr /.n ˛n /. We remark that since vIr > vN
I N i I N i
r ,
ti > ni and te > ne , it follows that ˛O > 0, ˇO > 0 and O < 0.
The conormal, axial and orthogonal current densities Jc ; Ja ; Jo define three
distributions of dipolar current sources lying on the ischemic boundary with dipole
axes parallel to the conormal direction Dn, to the fiber direction al and orthogonal
to the fiber direction, respectively. Since only rv appears in the current density, we
shift the distribution v by introducing the discrepancy with respect to the normal
resting value vN O D v vN r D .vr vr /˝i .x/.
I N
r , i.e. v
The transmembrane component uvO is given by
j˝i j
uvO .x/ D ˛.v
O Ir vN / ˝i .x/ :
r
j˝H j
In the case of the endocardial surface in contact with the cavitary blood, we have
where uhs is the heart surface component, see (5.41) and (5.42).
The full potential distribution u is the superposition of its field components
For x 2 ˝H [ ˝b , i.e. within the cardiac wall or in the extracardiac medium, the
full potential can be expressed by means of the following integral over the cardiac
domain
Z Z
w.x; t/ D JTv r .; x/ d D .rv.; t//T Di ./r .; x/ d ;
˝H ˝H
(9.5)
where .; x/ is the solution of the following lead field problem
8
ˆ
ˆ O D 1 ./ C ı. x/ 2 ˝H [ ˝b
ˆ
ˆ div Dr
ˆ
< j˝H j ˝H
ˆ O ˙ D 0
ŒŒ ˙ D 0; ŒŒ nT Dr 2˙
ˆ
ˆ
ˆ
:̂ nT Dr
O D 0 on ;
with ˝H ./ the characteristic function of ˝H , and ı. x/ denotes the Dirac delta
function at point x. Note that the boundary condition imposed on the “lead field”
actually reflects the property that w.x; t/, defined by (9.5), has zero average in the
heart domain.
9.6 Subendocardial Ischemia, ST Depression and Elevation 343
Using again as a reference the average potential on ˝H , the split field com-
ponents wa , wo , wc , associate with the sources Jc , Ja and Jo , respectively, can be
represented by means of the following surface integral on the ischemic boundary
Z
wc .x; t/ D ˛O D./n./ r .; x/ d ;
i
Z
wa .x; t/ D ˇO .aTl ./n.//al ./ r .; x/ d ;
i
Z
wo .x; t/ D O I al ./aTl ./ n./ r .; x/ d :
i
w D wIS C wAS
as the decomposition of the full extracellular waveform into its isotropic and
anisotropic components.
Here r 2 Œ0; 1; 2 Œ min ; max ; 2 Œmin ; max ; a.r/ D a1 Cr.a2 a1 /; c.r/ D
c1 Cr.c2 c1 / and ai ; ci ; i D 1; 2 are given coefficients determining the main axes of
the ellipsoid. The intracavitary blood and the extracardiac bath are also modeled as
a set of packed ellipsoidal surfaces in contact with the endocardial and the epicardial
surface, respectively.
We have assumed homogeneous cellular membrane properties, i.e. all individual
cells have the same intrinsic transmembrane action potential, except in the ischemic
region. The three main pathophysiological conditions of myocardial ischemia are
344 9 Simulation Studies of Cardiac Bioelectrical Activity
al .x/ D ue 1 cos ˛.r/ C ue 2 sin ˛.r/; at .x/ D ue 3 ; an .x/ D ue 1 sin ˛.r/ ue 2 cos ˛.r/
2
˛.r/ D .1 r/ ; with r D x3 =c:
3 4
(9.6)
LR1
50
CONTROL
MODERATE
SEVERE
0
mV
−50
−100
0 50 100 150 200 250 300 350 400
ms
Fig. 9.53 Control, moderate and severe ischemic action potentials generated by 0-dimensional
LR1 model
9.6 Subendocardial Ischemia, ST Depression and Elevation 345
We will also consider an isotropic conducting media modeling the cavitary blood
with conductivity coefficient b D 6 103 1 cm1 .
The macroscopic features of the excitation and subsequent repolarization process
are described by extracting from the spatio-temporal transmembrane potential
the sequence of the propagating excitation and repolarization wave fronts. In
particular, we define the excitation time te .x/ at a given point x as the unique time
when v.x; te .x// D 40 mV during the excitation phase. Analogously, during the
repolarization phase we define the recovery time as the unique time instant tr .x/
when v.x; tr .x// D 50 mV. In Figs. 9.54 and 9.55, we report the isochrone lines
300
60
40 230
20 250
36.76 105.22 264.54 332.39 225.88 231.10
300
60
40 230
20
250
19.09 111.58 250.93 339.78 227.15 232.29
235
300
60
40 230
20
250
0.50 124.54 238.34 351.09 225.73 240.92
Fig. 9.54 Contour plots of depolarization time (first column ACTI), repolarization time (second
column REPO) and action potential duration (third column APD) on epi, midwall and endocardial
surfaces in case of healthy tissue with multiple endocardial stimulation. Reported below each panel
are the maximum, minimum and step in ms of the displayed map
346 9 Simulation Studies of Cardiac Bioelectrical Activity
80 300
200
60
200
EPI
150
40
100
20 100
50
38.08 86.85 267.92 314.93 228.06 230.78
80 300
200
60
MID
200 150
40
100
20 100
50
17.35 75.98 187.37 305.17 166.83 235.00
80 300
200
60
ENDO
200 150
40
100
20 100
50
0.27 68.28 105.06 297.00 102.39 242.32
Fig. 9.55 Contour plots of depolarization time (first column ACTI), repolarization time (second
column REPO) and action potential duration (third column APD) on epi, midwall and endocardial
surfaces in case of an acute subendocardial ischemic region. Reported below each panel are the
maximum, minimum and step in ms of the displayed map
of the activation time te .x/ (first column, ACTI), repolarization time tr .x/ (second
column, REPO) and APD = tr .x/ te .x/ (third column) on the epicardium (first
row), midwall (second row) and endocardium (third row), for a healthy tissue with
multiple endocardial stimulation and for a tissue with a subendordial ischemic
region and central endocardial stimulation, respectively.
Extracellular potential distributions and electrograms: healthy tissue. We
report in Fig. 9.56 the decomposition of four epicardial electrograms and in Fig. 9.57
the decomposition of 16 electrocardiograms recorded 2 cm far away from the
epicardium, in case of normal tissue. We observe that all the isotropic components
(in blue) present a mainly positive QRS complex, while the anisotropic components
(in red) present a negative QRS complex. Due to the larger magnitude of the
anisotropic components, the resulting full extracellular/extracardiac waveforms (in
black) have a mainly negative biphasic QRS complex.
9.6 Subendocardial Ischemia, ST Depression and Elevation 347
5
mV
−5
5
mV
−5
0 400 0 400
ms ms
Fig. 9.56 Decomposition of four epicardial electrograms in case of normal tissue. Full extracellu-
lar waveform (black), isotropic component (blue) and anisotropic component (red)
3
mV
−3
3
mV
−3
3
mV
−3
3
mV
−3
0 400 0 400 0 400 0 400
ms ms ms ms
Fig. 9.57 Decomposition of 16 electrocardiograms recorded 2 cm far away from the epicardium
in case of normal tissue. Full extracardiac waveform (black), isotropic component (blue) and
anisotropic component (red)
Atrial or ventricular reentry is a type of arrhythmia where the electric signal is not
completing the normal circuit, but it rather follows an alternative circuit looping
back upon itself and developing a self-perpetuating rapid and abnormal activation,
also called circus movement. Necessary conditions for the onset of reentry include
a combination of unidirectional block and slowed conduction. Reentry is divided
into two major types: anatomical and functional reentry. The circus movement can
occur around an anatomical (e.g. an infarct region) or functional core. Either of two
types may occur alone, or together. In the following, we will focus on functional
reentry, and in particular on the causes of its induction.
A widely accepted theory for the induction of functional reentry in cardiac tissue
is based on a hypothesis proposed by Winfree in [561], known as the critical
point hypothesis, see also the review [431]. This hypothesis can be illustrated by
considering the so-called pinwheel experiment. Suppose that an action potential
wavefront, elicited by a brief electrical stimulus (S1), propagates from left to right
in a two-dimensional sheet of cardiac tissue, as shown in Fig. 9.61, Panels a–c.
9.7 Reentry Phenomena 349
FULL IS AS
Fig. 9.58 Contour plots of the epicardial extracellular potential decomposition during the ST
interval (140 ms) for different percentages (10, 25, 50 %) of the subendocardial ischemic region
thickness with respect to the total transmural wall thickness. Reported below each panel are the
maximum, minimum and step in mV of the displayed map
During the plateau and repolarization phases of the action potential, the tissue is
refractory. The critical phase of the action potential is the time when the application
of a stimulus of appropriate strength induces reentry and it occurs during the so-
called vulnerable window, near the end of the refractory period.
Suppose now that a second brief point stimulus (S2) is applied at the center of
the sheet. If the S2 stimulus is applied early (see Fig. 9.61, Panel d), the tissue
under the stimulation electrode is still refractory and the stimulus does not elicit
a wavefront (see Fig. 9.61, Panel g). On the other hand, if the S2 stimulus is applied
late (see Fig. 9.61, Panel f), the tissue under the electrode has already recovered from
refractoriness and a closed wavefront propagating outward is initiated (see Fig. 9.61,
Panel i). However, if the S2 stimulus is applied at an intermediate time (see Fig. 9.61,
Panel e), when the critical phase is occurring in the tissue under the electrode, the
portion of tissue on the left of the electrode is excited, while that on the right is still
too refractory to be excited. As a result, a wavefront is propagating outward, but it is
not closed (see Fig. 9.61, Panel h). It ends at two points where the contours of critical
350 9 Simulation Studies of Cardiac Bioelectrical Activity
10 10 10 10
0 0 0 0
10 10 10 10
0 0 0 0
10 10 10 10
0 0 0 0
10 10 10 10
0 0 0 0
Fig. 9.59 Epicardial electrograms calculated from a grid of 4 4 electrodes in the case of an acute
subendocardial ischemic region 50 % thick
0 0 0
−5 −5 −5
0 200 400 0 200 400 0 200 400
ms ms ms
Fig. 9.60 Decomposition of the epicardial electrogram above the subendocardial ischemic region
for different percentages (10, 25, 50 %) of the ischemic region thickness with respect to the total
transmural wall thickness. Full extracellular waveform (black), isotropic component (blue) and
anisotropic component (red)
9.7 Reentry Phenomena 351
a t = 20 ms
50
b t = 40 ms
50
c t = 60 ms
50
0 0 0
d t = 265 ms
50
e t = 285 ms
50
f t = 305 ms
50
0 0 0
0 0 0
Fig. 9.61 The pinwheel experiment on a two-dimensional sheet of cardiac tissue. Panels a-b-c:
transmembrane potential patterns at three instants during the propagation of the planar excitation
wavefront elicited by the S1 stimulation. Panels d-e-f: transmembrane potential patterns during an
early (d), intermediate (e) and late (f) S2 central point stimulation. Panels g-h-i: transmembrane
potential patterns some instants after the onset of the early (g), intermediate (h) and late (i) S2
stimulations. Below each panel are reported the minimum, maximum and step in mV of the
displayed map and the colorbar denotes the range of values of the displayed equipotential lines
phase and critical depolarization intersect. Frazier et al. in [185] refer to these points
as critical points. As time proceeds further, the two critical points become the pivots
of two spiral excitation waves. In the cardiac electrophysiology literature, spiral
waves are also called rotors and the particular type of reentry described above is
called figure of eight reentry.
The critical point hypothesis has been verified experimentally by Shibata et al.
[476], who performed the pinwheel experiment in a dog heart. The critical point
hypothesis also inspired an experimental method for the induction of a single rotor in
the heart, known as cross-field stimulation. In this method, two electric stimulations,
S1 and S2, are delivered with an appropriate delay through two lines of electrodes
352 9 Simulation Studies of Cardiac Bioelectrical Activity
aligned perpendicular to each other, see e.g. [185], where cross-field stimulation
was used to initiate rotors in a dog heart.
The pinwheel experiments [476] and the cross-field stimulation experiments
[185] validated strongly the critical point hypothesis, which now, thanks to this
experimental evidence, is usually referred to as critical point theory.
In the pinwheel experiment, a pair of rotors is induced by applying two
successive stimuli, S1 and S2, at different locations. Suppose now that both S1 and
S2 are applied at the same site, so that the S1 critical phase contour and the S2
stimulus contour are both circles centered at the same point. Because two concentric
circles never intersect, the critical point theory predicts that reentry cannot occur.
However, Matta et al. in [322] performed this experiment and initiated reentry.
A possible explanation of this paradox is based on the anisotropic Bidomain
model, and in particular on the virtual electrode polarization pattern due to the
unequal anisotropy rations of the intra- and extracellular media. In [429], Roth
showed by Bidomain numerical simulations that an S1-S2 stimulation protocol
through the same unipolar electrode might induce reentry.
Three-dimensional reentry simulations. We now turn to the simulation of
reentry phenomena in three dimensions, where the possible configurations of
reentrant fronts are much more complex and less understood than in two dimensions
(three-dimensional rotors are often called scroll waves). There is a vast literature on
cardiac reentry simulations and experiments; among the several simulation works,
we refer the interested reader to e.g. [49, 98, 170, 171, 191, 287, 345, 368, 411, 412,
431, 450, 488, 519, 562, 569], Part VII of [587, 588], and the references therein. Due
to the high computational complexity of large scale simulations, virtually all works
in the vast existing literature on cardiac reentry simulations employ some model
simplifications in order to obtain a tractable discrete problem see e.g. [278], [589,
Part V and VII], [369]. The most used simplifications are in the cardiac tissue model
(Monodomain instead of Bidomain, fiber structure without intramural rotation or
full anisotropy), in the ionic model (FHN or intermediate variations instead of
LR models), in the domain (two instead of three dimensions), in the numerical
methods (operator splitting instead of the fully coupled Bidomain system, coarse
time and/or space mesh sizes compared to the length scale of the problem), etc. On
the contrary, we retain here the full complexity of the coupled Bidomain–LR1 model
with intramural fiber rotation and orthotropic anisotropy in three dimensions and we
compare the resulting scroll waves with the one obtained with the Monodomain–
LR1 model. With the latter, we then simulate scroll wave breakups on larger
domains and counterrotating scroll waves on half ventricle.
We start with the simulation of a stable scroll wave using both the anisotropic
Bidomain–LR1 and the Monodomain–LR1 models. The conductivity tensors are
assumed axisymmetric with values
9.7 Reentry Phenomena 353
li D 3 103 ; ti D 3:1525 104; le D 2 103 ; te D 1:3514 103 .1 cm1 /
The breakup of spiral and scroll waves has been intensely studied in recent
years because it could underlie ventricular fibrillation according to some authors
354 9 Simulation Studies of Cardiac Bioelectrical Activity
50
1
m (w1)
0
0.5
V
−50
0
0 200 400 600 800 0 200 400 600 800
1 1
h (w2)
j (w3)
0.5 0.5
0 0
0 200 400 600 800 0 200 400 600 800
0.1 1
d (w4)
f (w5)
0.05 0.5
0 0
0 200 400 600 800 0 200 400 600 800
−5
x 10
0.2
10
Cai (w7)
X (w )
6
0.1 8
6
0
4
0 200 400 600 800 0 200 400 600 800
Fig. 9.62 Time course of the transmembrane potential v, of the LR1 gating variables
m; h; j; d; f; X and of the Ca2C concentration during reentry determined by a stable scroll wave
(800 ms)
and it still remains a complex phenomenon with multiple causes not yet fully
understood, see e.g. Fenton et al. [171], Garfinkel et al. [191]. Among the possible
factors contributing to the breakup of the excitation front and to the instability of
reentry circuits, many researchers indicated steep APD restitution curve, intramural
rotational anisotropy, tissue thickness, filament twist, etc.
We simulated a spiral and scroll wave breakup using the axisymmetric
anisotropic Monodomain with the conductivity coefficients
and the LR1 model with GNa D 16; GK D 0:432; as before, but setting
the conductance of the slow-inward current Gsi D 0:056. It has been shown
that increasing this parameter increases the meandering of the scroll filament and
eventually leads to a breakup regime if the cardiac tissue is thick enough; see
Garfinkel et al. [191].
9.7 Reentry Phenomena 355
Fig. 9.63 Stable scroll wave with Bidomain–LR1 model. Domain: 2 2 0:5 cm3 ; 4x D
0:01; Gsi D 0, t D 90, 100, 110, 120, 130, 140, 150, 160, 170, 180, 190, 200 ms. The colormap
ranges from blue (resting values around 84 mV) to red (excitation front around 10 mV)
Fig. 9.64 Stable scroll wave with Monodomain–LR1 model. Domain: 2 2 0:5 cm3 ; 4x D
0:01; Gsi D 0, t D 90, 100, 110, 120, 130, 140, 150, 160, 170, 180, 190, 200 ms. The colormap
ranges from blue (resting values around 84 mV) to red (excitation front around 10 mV)
of the front evolution at differently spaced time intervals: (a) just after the first
breakup (first rows, t D 500; 505; 510 ms, second row, t D 520; 530; 540 ms) due
to a head-to-tail collision and generating two new spiral tips, one drifting outside
the domain and the other, near the center of the slab, generating a new scroll wave;
(b) a subsequent breakup (third row, t D 705; 715; 725 ms) generating additional
scroll waves and spiral tips; (c) a later time with many broken fronts (last row,
t D 1 250; 1 275; 1 300ms) displaying a spatio-temporal chaotic configuration of
multiple wave reentry. Figure 9.67 shows the approximated scroll filaments at an
early time after reentry initiation (t D 265 ms, left panel) when only one filament
9.7 Reentry Phenomena 357
a b
10 0
0 −10
−10 −20
−20
−30
−30
−40
−40
−50
−50
−60
−60
−70 −70
−80 −80
Fig. 9.65 Transmembrane potential at a given point as a function of time for the two tests of
Fig. 9.63 ((a), Bidomain model) and Fig. 9.64 ((b), Monodomain model).
is present and almost at the end of our simulation (t D 1 207 ms) when many
filaments have appeared (for clarity, the slab height has been magnified). Figure 9.68
compares the time course of the transmembrane potential v of the Monodomain–
LR1 model for the stable scroll wave of Fig. 9.64 (left) and for the unstable scroll
wave with breakups of Fig. 9.66 (right).
Fig. 9.66 Scroll wave breakup with Monodomain–LR1 model, Gsi D 0:056, domain: 6 6
0:6 cm3 . First row: t D 500, 505, 510; second row: t D 520, 530, 540; third row: t D 705, 715,
725; fourth row: t D 1 250, 1 275, 1 300 ms. The colormap ranges from blue (resting values around
84 mV) to red (excitation front around 10 mV)
a second counterrotating scroll wave. The evolution of these two scroll waves is
displayed in Fig. 9.69, where after an initial adjustment, the two scroll waves seem
to reach a stable counterrotating configuration (last panel, t D 500 ms), also known
as figure-8 or double loop reentry, see e.g. Wit and Janse [564].
9.7 Reentry Phenomena 359
a b
6 6
0.6 0.6
6 6
0 0
Fig. 9.67 Scroll wave filaments for the breakup test of Fig. 9.66 near reentry initiation ((a), t D
265 ms) and during breakup regime ((b), t D 1 297 ms). Monodomain–LR1 model with Gsi D
0:056. Domain: 6 6 0:6 cm3 (height of the slab has been magnified)
10
0
0
−10
−10
−20
−20
−30
−30
−40
−40
−50
−50
−60 −60
−70 −70
−80 −80
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
3 3
x 10 x 10
T = 25
T = 0.25
13.55 22.366
−20.0931 −14.3799
−53.7362 −51.1257
−84.015 −84.197
T = 50 T = 100
23.403 21.216
−13.6991 −15.1299
−50.8011 −51.4757
−84.193 −84.187
T = 300 T = 500
19.886 27.703
−15.9061 −10.6991
−51.6981 −49.1011
−83.911 −83.663
Fig. 9.69 Counterrotating double scroll wave with the Monodomain–LR1 model on half ventricle
(500 500 80 elements). Distribution of v on the endocardium and intramural sections at times
t D 0:25, 25, 50, 100, 300, 500 ms. Viewpoint is inside the ventricle
Appendix A
Cardiac Simulation Projects, Software,
Libraries
1. Aliev, R.R., Panfilov, A.V.: A simple two-variable model of cardiac excitation. Chaos Solitons
Fract. 7, 293–301 (1996)
2. Alistair, Y., Frangi, A.: Computational cardiac atlases: from patient to population and back.
Exp. Physiol. 94(5), 578–596 (2009)
3. Amato, S., Bellettini, G., Paolini, M.: The nonlinear multidomain model: a new formal
asymptotic analysis. In: Chambolle, A., Novaga, M., Valdinoci, E. (eds.) Proceedings of
Geometric Partial Differential Equations, pp. 33–74. Edizioni della Normale, Pisa (2013)
4. Ambrosio, L., Colli Franzone, P., Savaré, G.: On the asymptotic behaviour of anisotropic
energies arising in the cardiac bidomain model. Interfaces Free Bound. 2(3), 213–266 (2000)
5. Ambrosio, L., Fusco, N., Pallara, D.: Functions of Bounded Variation and Free Discontinuity
Problems. Oxford University Press, Oxford (2000)
6. Ambrosio, L., Gigli, N., Savaré, G.: Gradient flows in metric spaces and in the space of
probability measures. Lectures in Mathematicsi, ETH Zuerich, 2nd edn. Birkhäuser, Basel
(2008)
7. Antzelevitch, C., Fish, J.: Electrical heterogeneity within the ventricular wall. Basic Res.
Cardiol. 96, 517–527 (2001)
8. Antzelevitch, C., Sicouri, S., Lukas, A., Nesterenko, V.V., Lu, D.-W., Di Diego, J.M.:
Regional differences in the electrophysiology of ventricular cells: physiological and clinical
implications. In: Zipes, D., Jalife, J. (eds.) Cardiac Electrophysiology, chap. 23, pp. 228–245.
W. B. Sauders, Philadelphia (1995)
9. Anyukhovsky, E.P., Sosunov, E.A., Rosen, M.R.: Regional differences in electrophysiological
properties of epicardium, midmyocardium and endocardium. Circulation 94, 1981–1988
(1996)
10. Anyukhovsky, E.P., Sosunov, E.A., Gainullin, R.Z., Rosen, M.R.: The controversial M cell. J.
Cardiovasc. Electrophysiol. 10, 244–260 (1999)
11. Arevalo, H., Rodriguez, B., Trayanova, N.A.: Arrhythmogenesis in the heart: multiscale mod-
eling of the effects of defibrillation shocks and the role of electrophysiological heterogeneity.
Chaos 17(1), 015103 (2007)
12. Arisi, G., Macchi, E., Corradi, C., Lux, R.L., Taccardi, B.: Epicardial excitation during
ventricular pacing. Relative independence of breakthrough sites from excitation sequence in
canine right ventricle. Circ. Res. 71, 840–849 (1992)
13. Arthurs, C.J., Bishop, M.J., Kay, D.: Efficient simulation of cardiac electrical propagation
using high order finite elements. J. Comput. Phys. 231(10), 3946–3962 (2012)
14. Arthurs, C.J., Bishop, M.J., Kay, D.: Efficient simulation of cardiac electrical propagation
using high order finite elements. II: adaptive p-version. J. Comput. Phys. 253, 443–470 (2013)
15. Asanov, G.S.: Finsler Geometry, Relativity and Gauge Theories. D. Reidel, Dordrecht (1985)
16. Ashihara, T., Trayanova, N.A.: Cell and tissue responses to electric shocks. Europace 7(s2),
S155–S165 (2005)
17. Aslanidi, O.V., Colman, M.A., Stott, J., Dobrzynski, H., Boyett, M.R., Holden, A.V.,
Zhang, H.: 3D virtual human atria: a computational platform for studying clinical atrial
fibrillation. Prog. Biophys. Mol. Biol. 107, 156–168 (2011)
18. Attin, M., Clusin, W.T.: Basic concepts of optical mapping techniques in cardiac electrophys-
iology. Biol. Res. Nurs. 11(2), 195–207 (2009)
19. Austin, T.M., Trew, M.L., Pullan, A.J.: Solving the cardiac Bidomain equations for discon-
tinuous conductivities. IEEE Trans. Biomed. Eng. 53(7), 1265–1272 (2006)
20. Axelsson, O.: Iterative Solution Methods. Cambridge University Press, Cambridge (1994)
21. Ayache, N., Delingette, H., Sermesant, M. (eds.): Proceedings of the 5th International
Conference on Functional Imaging and Modeling of the Heart, FIMH’09, Nice, 3–5 June
2009. LNCS, vol. 5528. Springer, Berlin (2009)
22. Azzouzi, A., Coudiere, Y., Turpault, R., Zemzemi, N.: A mathematical model of the Purkinje-
muscle junctions. Math. Biosci. Eng. 8(4), 915–930 (2011)
23. Backhvalov, N., Panasenko, G.: Homogenization: Averaging Processes in Periodic Media.
Kluwer Academic, Dordrecht (1990)
24. Balay, S., Buschelman, K., Gropp, W.D., Kaushik, D., Knepley, M., Curfman McInnes, L.,
Smith, B.F., Zhang, H.: PETSc home page. http://www.mcs.anl.gov/petsc (2001)
25. Balay, S., Buschelman, K., Gropp, W.D., Kaushik, D., Knepley, M., Curfman McInnes,
L., Smith, B.F., Zhang, H.: PETSc users manual. Tech. Rep. ANL-95/11 – Revision 2.1.5,
Argonne National Laboratory (2002)
26. Barcilon, V., Chen, D.-P., Eisenberg, R.S., Jerome, J.W.: Qualitative properties of steady-
state Poisson-Nernst-Planck systems: perturbation and simulation study. SIAM J. Appl. Math.
57(3), 631–648 (1997)
27. Barles, G.: Fully non-linear Neumann type boundary conditions for second-order elliptic and
parabolic equations. J. Differ. Equ. 106, 90–106 (1993)
28. Barr, R.C., Plonsey, R.: Propagation of excitation in idealized two-dimensional tissue.
Biophys. J. 45, 1191–1202 (1984)
29. Barr, R.C., Spach, M.S.: Inverse calculation of QRS-T epicardial potentials from body surface
potential distributions for normal and ectopic beats in the intact dog. Circ. Res. 42, 661–675
(1978)
30. Baruffi, S., Spaggiari, S., Stilli, D., Musso, E., Taccardi, B.: The importance of fiber
orientation in determining the features of cardiac electric field. In: Antaloczy, Z. (ed.) Modern
Electrocardiology, pp. 89–92. Excerpta Medica, Amsterdam (1978)
31. Baruscotti, M., Barbuti, A., Bucchi, A.: The cardiac pacemaker current. J. Mol. Cell. Cardiol.
48, 55–64 (2008)
32. Bassetti, F.: Variable time-step discretization of degenerate evolution equations in Banach
space. Numer. Funct. Anal. Optim. 24(3–4), 391–426 (2003)
33. Beeler, G.W., Reuter, H.T.: Reconstruction of the action potential of ventricular myocardial
fibers. J. Physiol. 268, 177–210 (1977)
34. Beg, M.F., Helm, P.A., McVeigh, E., Miller, M.I., Winslow, R.L.: Computational cardiac
anatomy using MRI. Magn. Reson. Med. 52,1167–1174 (2004)
35. Belhamadia, Y.: A time-dependent adaptive remeshing for electrical waves of the heart. IEEE
Trans. Biomed. Eng. 55(2), 443–452 (2008)
36. Belhamadia, Y., Fortin, A., Bourgault, Y.: Towards accurate numerical method for mon-
odomain models using a realistic heart geometry. Math. Biosci. 220(2), 89–101 (2009)
37. Bellettini, G., Paolini, M.: Anisotropic motion by mean curvature in the context of Finsler
geometry. Hokkaido Math. J. 25, 537–566 (1996)
38. Bellettini, G., Paolini, M., Venturini, S.: Some results on surface measures in calculus of
variations. Ann. Math. Pura Appl. 170, 329–359 (1996)
39. Bellettini, G., Colli Franzone, P., Paolini, M.: Convergence of front propagation for
anisotropic bistable reaction-diffusion equations. Asymptot. Anal. 15, 325–358 (1997)
References 369
40. Bellettini, G., Paolini, M., Pasquarelli, F.: Nonconvex mean curvature flow as a formal
singular limit of the nonlinear bidomain model. Adv. Differ. Equ. 18(9–10), 895–934 (2013)
41. Beltrami, A.P., et al.: Adult cardiac stem cells are multipotent and support myocardial
regeneration. Cell 114, 763–776 (2003)
42. Bendahmane, M., Karlsen, K.H.: Analysis of a class of degenerate reaction-diffusion systems
and the bidomain model of cardiac tissue. Netw. Heterog. Media 1, 185–218 (2006)
43. Bendahmane, M., Karlsen, K.H.: Convergence of a finite volume scheme for the bidomain
model of cardiac tissue. Appl. Numer. Math. 59(9), 2266–2284 (2009)
44. Bensoussan, A., Lions, J.-L., Papanicolaou, G.: Asymptotic Analysis for Periodic Structures.
North-Holland, Amsterdam (1978)
45. Berbari, E.J., Lander, P., Scherlag, B.J., Lazzara, R., Geselowitz, D.B.: Ambiguities of
epicardial mapping. J. Electrocardiol. 24(Suppl), 16–20 (1992)
46. Bergmann, O., et al.: Evidence for cardiomyocyte renewal in humans. Science 324, 98–102
(2009)
47. Bers, D.M.: Excitation-Contraction Coupling and Cardiac Contractile Force, 2nd edn.
Kluwer, Dordrecht (2001)
48. Bertran, G., Biagetti, M.O., Valverde, E., Arini, P.D., Quinteiro, R.A., Lack of effect of
conduction direction on action potential durations in anisotropic ventricular strips of pig heart.
J. Cardiovasc. Electrophysiol. 13(4), 380–387 (2002)
49. Biktashev, V.N., Holden, A.V., Mironov, S.F., Pertsov, A.M., Zaitsev, A.V.: Three-dimensional
organisation of re-entrant propagation during experimental ventricular fibrillation. Chaos
Solitons Fract. 13(8), 1713–1733 (2002)
50. Bishop, M.J., Gavahan, D.J., Trayanova, N.A., Rodriguez, B.: Photon scattering effects in
optical mapping of propagation and arrhythmogenesis in the heart. J. Electrocardiol. 40, S75–
S80 (2007)
51. Bishop, M.J., Boyle, P.M., Plank, G., Welsh, D.G., Vigmond, E.J.: Modeling the role of the
coronary vasculature during external field stimulation. IEEE Trans. Biomed. Eng. 57(10),
2335–2345 (2010)
52. Bishop, M.J., Plank, G., Burton, R., Schneider, J., Gavaghan, D., Grau, V., Kohl, P.: Develop-
ment of an anatomically-detailed MRI-derived rabbit ventricular model and assessment of its
impact on simulation of electrophysiological function. Am. J. Physiol. Heart Circ. Physiol.
298, H699–H718 (2010)
53. Bondarenko, V.E., Szigeti, G.P., Bett, G.C., Kim, S.: A computer model for the action
potential of mouse ventricular myocytes. Am. J. Physiol. Heart Circ. Physiol. 278, H1378–
H1403 (2004)
54. Bordas, R., Carpentieri, B., Fotia, G., Maggio, F., Nobes, R., Pitt-Francis, J., Southern, J.:
Simulation of cardiac electrophysiology on next-generation high-performance computers.
Philos. Trans. R. Soc. A 367(1895), 1951–1969 (2009)
55. Bordas, R.M., Gillow, K., Gavaghan, D., Rodriguez, B., Kay, D.: A Bidomain model of the
ventricular specialized conduction system of the heart. SIAM J. Appl. Math. 72(5), 1618–
1643 (2012)
56. Beuter, A., Glass, L., Mackey, M.C., Titcombe, M.S.: Nonlinear Dynamics in Physiology and
Medicine. Springer, New York (2003)
57. Bouchard, S., Jacquemet, V., Vinet, A.: Automaticity in acute ischemia: bifurcation analysis
of a human ventricular model. Phys. Rev. E 83, 011911-1-10 (2011)
58. Boulakia, M., Fernandez, M.A., Gerbeau, J.-F., Zemzemi, N.: Towards the Numerical
Simulation of Electrocardiograms. In: FIMH07, Salt Lake City, 7–9 June 2007. LNCS,
vol. 4466, pp. 240–249. Springer, Berlin (2007)
59. Boulakia, M., Fernandez, M.A., Gerbeau, J.-F., Zemzemi, N.: A coupled system of PDEs
and ODEs arising in electrocardiograms modelling. Appl. Math. Res. Exp. 2, 1–28 (2008).
doi:10.1093/amrx/abn002
60. Bourgault, Y., Coudiere, Y., Pierre, C.: Existence and uniqueness of the solution for the
bidomain model used in cardiac electrophysiology. Nonlinear Anal. Real World Appl. 10,
458–482 (2009)
370 References
61. Brenner, S.C., Scott, L.R.: The Mathematical Theory of Finite Element Methods. Texts in
Applied Mathematics, vol. 15, 2nd edn. Springer, New York (2002)
62. Brezis, H.: Analyse Fonctionnelle, Theory and Applications. Masson, Paris (1983)
63. Britton, N.F.: Reaction – Diffusion Equations and Their Applications to Biology. Academic,
London (1986)
64. Brooks, D.H., Ahmad, G.F., MacLeod, R.S., Maratos, G.M.: Inverse electrocardiography
by simultaneous imposition of multiple constraints. IEEE Trans. Biomed. Eng. 46(1), 3–18
(1999)
65. Bueno-Orovio, A., Cherry, E., Fenton, F.: Minimal model for human ventricular action
potentials in tissue. J. Theor. Biol. 253, 544–560 (2008)
66. Buist, M., Sands, G., Hunter, P., Pullan, A.: A deformable finite element derived finite
difference method for cardiac activation problems. Ann. Biomed. Eng. 31, 577–588 (2003)
67. Burgess, M.J., Steinhaus, B.M., Spitzer, K.W., Ershler, P.R.: Nonuniform epicardial activation
and repolarization properties of in vivo canine pulmonary conus. Circ. Res. 62(2), 233–246
(1988)
68. Burnes, J.E., Taccardi, B., Rudy, Y.: Noninvasive imaging modality for cardiac arrhythmias.
Circulation 102, 2152–2158 (2000)
69. Burnes, J.E., Taccardi, B., Ershler, P.R., Rudy, Y.: Noninvasive electrocardiographic imaging
of substrate and intramural ventricular tachycardia in infarcted hearts. J. Am. Coll. Cardiol.
38, 2071–2078 (2001)
70. Burton, F.L., Cobbe, S.M.: Dispersion of ventricular repolarization and refractory period.
Cardiovasc. Res. 50, 10–23 (2001)
71. Cabo, C., Rosenbaum, D.: Quantitative Cardiac Electrophysiology. Marcel Dekker, New York
(2002)
72. Cai, X.-C.: Additive Schwarz algorithms for parabolic convection-diffusion equations.
Numer. Math. 60(1), 41–61 (1991)
73. Cai, X.-C.: Multiplicative Schwarz methods for parabolic problems. SIAM J. Sci. Comput.
15(3), 587–603 (1994)
74. Camara, O., Pop, M., Rhode, K., Sermesant, M., Smith, N., Young, A. (eds.): Proceedings
of Statistical Atlases and Computational Models of the Heart, STACOM’10, Beijing, 20 Sept
2010. LNCS, vol. 6364. Springer, Berlin (2010)
75. Camara, O., Konukoglu, E., Pop, M., Rhode, K., Sermesant, M., Young, A. (eds.): Proceed-
ings of Statistical Atlases and Computational Models of the Heart, STACOM’11, Toronto,
22 Sept 2011. LNCS, vol. 7085. Springer, Berlin (2012)
76. Camara, O., Mansi, T., Pop, M., Rhode, K., Sermesant, M., Young, A. (eds.): Proceedings of
Statistical Atlases and Computational Models of the Heart, STACOM’12, Nice, 5 Oct 2012.
LNCS, vol. 7746. Springer, Berlin (2013)
77. Camelliti, P., Borg, T.K., Kohl, P.: Structural and functional characterization of cardiac
fibroblasts. Cardiovasc. Res. 65, 40–51 (2005)
78. Carmeliet, E., Vereecke, J.: Cardiac Cellular Electrophysiology. Kluwer, Dordrecht (2002)
79. Casten, R.G., Cohen, H., Lagerstrom, P.A.: Perturbation analysis of an approximation to the
Hodgkin-Huxley theory. Q. Appl. Math. 32(4), 365–402 (1975)
80. Cates, A.W., Pollard, A.E.: A model study of intramural dispersion of action potential
duration in the canine pulmonary conus. Ann. Biomed. Eng. 26, 567–576 (1998)
81. Cerbai, E., Barbieri, M., Mugelli, A.: Characterization of the hyperpolarization-activated
current, If in ventricular myocytes isolated from hypertensive rats. J. Physiol. 481(3), 585–
591 (1994)
82. Cerbai, E., Barbieri, M., Mugelli, A.: Occurrence and properties of the hyperpolarization-
activated current If in ventricular myocytes from normotensive and hypertensive rats during
aging. Circulation 94, 1674–1681 (1996)
83. Cerqueira, M.D., Weissman, N.J., Dilsizian, V., et al.: Standardized myocardial segmentation
and nomenclature for tomographic imaging of the heart – a statement for healthcare
professionals from the Cardiac Imaging Committee of the Council on Clinical Cardiology
of the American Heart Association. Circulation 105(4), 539–542 (2002)
References 371
84. Chauhan, V.S., Downar, E., Nanthakumar, K., Parker, J.D., Ross, H.J., Chan, W., Picton, P.:
Increased ventricular repolarization heterogeneity in patients with ventricular arrhythmia
vulnerability and cardiomyopathy: a human in vivo study. Am. J. Physiol. Heart Circ. Physiol.
290(1), H79–H86 (2006)
85. Chen, X.-Y.: Dynamics of interfaces in reaction diffusion systems. Hiroshima Math. J. 21,
47–83 (1991)
86. Chen, X.-Y.: Generation and propagation of interfaces in reaction-diffusion systems. Trans.
Am. Math. Soc. 334(2), 877–913 (1992)
87. Chen, P.-S., Moser, K.M., Dembitsky, W.P., Auger, W.R., Daily, P.O., Calisi, C.M., Jamieson,
S.W., Feld, G.K.: Epicardial activation and repolarization patterns in patients with right
ventricular hypertrophy. Circulation 83, 104–118 (1991)
88. Cheng, D.K., Tung, L., Sobie, E.A.: Nonuniform responses of transmembrane potential
during electric field stimulation of single cardiac cells. Am. J. Physiol. Heart Circ. Physiol.
277, H351–H362 (1999)
89. Cheng, L.K., Bodley, J.M., Pullan, A.: Comparison of potential- and activation-based
formulations for the inverse problem of electrocardiology. IEEE Trans. Biomed. Eng. 50(1),
11–22 (2003)
90. Cherry, E.M., Greenside, H.S., Henriquez, C.S.: A space-time adaptive method for simulating
complex cardiac dynamics. Phys. Rev. Lett. 84(6), 1343–1346 (2000)
91. Cherry, E.M., Greenside, H.S., Henriquez, C.S.: Efficient simulation of three-dimensional
anisotropic cardiac tissue using an adaptive mesh refinement method. Chaos 13, 853–865
(2003)
92. Ciarlet, P.G.: The finite element method for elliptic problems. Classics in Applied Mathemat-
ics, vol. 40. SIAM, Philadelphia (2002)
93. Clancy, C.E., Rudy, Y.: Linking a genetic defect to its cellular phenotype in a cardiac
arrhythmia. Nature 400 566–569 (1999)
94. Clancy, C.E., Rudy, Y.: NaC channel mutation that causes both Brugada and Long-QT
syndrome phenotypes: a simulation study of mechanism. Circulation 105, 1208–1213 (2002)
95. Clayton, R.H., Holden, A.V.: Propagation of normal beats and re-entry in a computational
model of ventricular cardiac tissue with regional differences in action potential shape and
duration. Prog. Biophys. Mol. Biol. 85(2–3), 473–499 (2004)
96. Clayton, R.H., Holden, A.V.: Dispersion of cardiac action potential duration and the
initiation of re-entry: a computational study. Biomed. Eng. 4, 11 (2005). Online: http://www.
biomedical-engineering-oline.com/content/4/1/11
97. Clayton, R.H., Panfilov, A.V.: A guide to modelling cardiac electrical activity in anatomically
detailed ventricles. Prog. Biophys. Mol. Biol. 96, 19–43 (2008)
98. Clayton, R.H., Zhuchkova, E.A., Panfilov, A.V.: Phase singularities and filaments: simplifying
complexity in computational models of ventricular fibrillation. Prog. Biophys. Mol. Biol. 90,
378–398 (2006)
99. Clements, J.C., Nenonen, J., Li, P.K.J., Horacek, B.M.: Activation dynamics in anisotropic
cardiac tissue via decoupling. Ann. Biomed. Eng. 32(7), 984–990 (2004)
100. Clerc, L.: Directional differences of impulse spread in trabecular muscle from mammalian
heart. J. Physiol. 255, 335–346 (1976)
101. Colli Franzone, P., Guerri, L.: Models of the spreading of excitation in myocardial tissue.
CRC Crit. Rev. Biomed. Eng. 20, 211–253 (1992); Pilkington, T.C., Loftis, B., Palmer, T.,
Budinger, T.F. (eds.) High-Performance Computing in Biomedical Research, pp. 359–401.
CRC, Boca Raton (1993)
102. Colli Franzone, P., Guerri, L.: Spread of excitation in 3-D models of the anisotropic cardiac
tissue I: validation of the eikonal approach. Math. Biosci. 113, 145–209 (1993)
103. Colli Franzone, P., Magenes, E.: On the inverse potential problem of electrocardiology.
Calcolo 16(4), 459–538 (1979)
104. Colli Franzone, P., Pavarino, L.F.: A parallel solver for reaction-diffusion systems in
computational electrocardiology. Math. Model Methods Appl. Sci. 14(6), 883–911 (2004)
372 References
105. Colli Franzone, P., Savaré, G.: Degenerate evolution systems modeling the cardiac electric
field at micro and macroscopic level. In: Lorenzi, A., Ruf, B. (eds.) Evolution Equations,
Semigroups and Functional Analysis, pp. 49–78. Birkhauser, Basel/Boston (2002)
106. Colli Franzone, P., Taccardi, B., Viganotti, C.: An approach to inverse calculation of epicardial
potentials from body surface maps. Adv. Cardiol. 21, 50–54 (1978)
107. Colli Franzone, P., Guerri, L., Taccardi, B., Viganotti, C.: The direct and inverse potential
problem in electrocardiology. Numerical aspects of some regularization methods and appli-
cation to data collected in isolated dog heart experiments. Tech. Rep. No. 222, IAN-CNR
(1979)
108. Colli Franzone, P., Guerri, L., Viganotti, C., Macchi, E., Baruffi, S., Spaggiari, S.,
Taccardi, B.: Potential fields generated by oblique dipole layer modeling excitation wave-
fronts in the anisotropic myocardium. Comparison with potential fields elicited by paced dog
hearts in a volume conductor. Circ. Res. 51(3), 330–346 (1982)
109. Colli Franzone, P., Guerri, L., Viganotti, C.: Oblique dipole layer potentials applied to
electrocardiology. J. Math. Biol. 17(1), 93–124 (1983)
110. Colli Franzone, P., Guerri, L., Magenes, E.: Oblique dipole layer potential for the direct and
inverse problems of electrocardiology. Math. Biosci. 68, 23–55 (1984)
111. Colli Franzone, P., Guerri, L., Viganotti, C., Taccardi, B.: Finite element approximation of
regularized solution of the inverse potential problem of electrocardiography and applications
to experimental data. Calcolo 12(1), 91–186 (1985)
112. Colli Franzone, P., Guerri, L., Rovida, S.: Wavefront propagation in an activation model of
the anisotropic cardiac tissue: asymptotic analysis and numerical simulations. J. Math. Biol.
28, 121–176 (1990)
113. Colli Franzone, P., Guerri, L., Tentoni, S.: Mathematical modeling of the excitation process
in myocardial tissue: influence of fiber rotation on wavefront propagation and potential field.
Math. Biosci. 101, 155–235 (1990)
114. Colli Franzone, P., Guerri, L., Taccardi, B.: Spread of excitation in a myocardial volume.
Simulation studies in a model of anisotropic ventricular muscle activated by point stimulation.
J. Cardiovasc. Electrophysiol. 4, 144–160 (1993)
115. Colli Franzone, P., Guerri, L., Pennacchio, M., Taccardi, B.: Spread of excitation in 3-D
models of the anisotropic cardiac tissue II: effects of fiber architecture and ventricular
geometry. Math. Biosci. 147, 131–171 (1998)
116. Colli Franzone, P., Guerri, L., Pennacchio, M., Taccardi, B.: Spread of excitation in 3-D
models of the anisotropic cardiac tissue III: effects of ventricular geometry and fiber structure
on the potential distribution. Math. Biosci. 151, 51–98 (1998)
117. Colli Franzone, P., Guerri, L., Pennacchio, M., Taccardi, B.: Anisotropic mechanisms for
multiphasic unipolar electrograms. Simulation studies and experimental recordings. Ann.
Biomed. Eng. 28, 1–17 (2000)
118. Colli Franzone, P., Pennacchio, M., Guerri, L.: Accurate computation of electrograms in the
left ventricular wall. Math. Model Methods Appl. Sci. 10(4), 507–538 (2000)
119. Colli Franzone, P., Guerri, L., Taccardi, B.: On the polyphasic character of simulated and
experimental electrograms. Biomed. Tech. 46(2), 16–19 (2001)
120. Colli Franzone, P., Pavarino, L.F., Taccardi, B.: A parallel solver for anisotropic cardiac
models. In: Proceedings of the IEEE Computers in Cardiology, Thessaloniki Chalkidiki,
21–24 Sept 2003. vol. 30, pp. 781–784 (2003)
121. Colli Franzone, P., Guerri, L., Taccardi, B.: Modeling ventricular excitation: axial and
orthotropic effects on wavefronts and potentials. Math. Biosci. 188, 191–205 (2004)
122. Colli Franzone, P., Pavarino, L.F., Taccardi, B.: Monodomain simulations of excitation and
recovery in cardiac blocks with intramural heterogeneity. In: Frangi, A.F., et al. (eds.):
FIMH05: Functional Imaging and Modeling of the Heart. LNCS, vol. 3504, pp. 267–277.
Springer, Berlin (2005)
123. Colli Franzone, P., Pavarino, L.F., Taccardi, B.: Simulating patterns of excitation, repolariza-
tion and action potential duration with cardiac Bidomain and Monodomain models. Math.
Biosci. 197, 35–66 (2005)
References 373
124. Colli Franzone, P., Deuflhard, P., Erdmann, B., Lang, J., Pavarino, L.F.: Adaptivity in space
and time for reaction-diffusion systems in electrocardiology. SIAM J. Sci. Comput. 28(3),
942–962 (2006)
125. Colli Franzone, P., Pavarino, L.F., Savarè, G.: Computational electrocardiology: mathematical
and numerical modeling. In: Quarteroni, A., et al. (eds.) Complex Systems in Biomedicine,
pp. 187–241. Springer, New York (2006)
126. Colli Franzone, P., Pavarino, L.F., Taccardi, B.: Effects of transmural electrical heterogeneities
and electrotonic interactions on the dispersion of cardiac repolarization and action potential
duration: a simulation study. Math. Biosci. 204(1), 132–165 (2006)
127. Colli Franzone, P., Pavarino, L.F., Scacchi, S., Taccardi, B.: Determining recovery times
from transmembrane action potentials and unipolar electrograms in normal heart tissue. In:
Sachse, F.B., Seemann, G. (eds.) FIMH’07, Salt Lake City, 7–9 June 2007. LNCS, vol. 4466,
pp. 139–149. Springer, Berlin (2007)
128. Colli Franzone, P., Pavarino, L.F., Scacchi, S., Taccardi, B.: Monophasic action potentials
generated by bidomain modeling as a tool for detecting cardiac repolarization times. Am. J.
Physiol. Heart Circ. Physiol. 293, H2771–H2785 (2007)
129. Colli Franzone, P., Pavarino, L.F., Scacchi, S.: Dynamical effects of myocardial ischemia in
anisotropic cardiac models in three dimensions. Math. Model Methods Appl. Sci. 17(12),
1965–2008 (2007)
130. Colli Franzone, P., Pavarino, L.F., Scacchi, S., Taccardi, B.: Modeling ventricular repolar-
ization: effects of transmural and apex-to-base heterogeneities in action potential durations.
Math. Biosci. 214, 140–152 (2008)
131. Colli Franzone, P., Pavarino, L.F., Scacchi, S.: Exploring anodal and cathodal make and break
cardiac excitation mechanisms in a 3D anisotropic bidomain model. Math. Biosci. 230(2),
96–114 (2011)
132. Colli Franzone, P., Pavarino, L.F., Scacchi, S.: Cardiac excitation mechanisms, wavefront
dynamics and strength – interval curves predicted by 3D orthotropic bidomain simulations.
Math. Biosci. 235(1), 66–84 (2012)
133. Conrath, C.E., Opthof, T.: Ventricular repolarization: an overview of (patho)physiology,
sympathetic effects and genetic aspects. Prog. Biophys. Mol. Biol. 92(3), 269–307 (2006)
134. Corbin II, L.V., Scher, A.M.: The canine heart as an electrocardiographic generator: depen-
dence on cardiac cell orientation. Circ. Res. 41, 58–67 (1977)
135. Coronel, R., de Bakker, J.M.T., Wilms-Schopman, F.J.G., Opthof, T., Linnenbank, A.C.,
Belterman, C.N., Janse, M.J.: Monophasic action potentials and activation recovery intervals
as measures of ventricular action potential duration: experimental evidence to resolve some
controversies. Heart Rhythm 3(9), 1043–1050 (2006)
136. Cortassa, S., Aon, M., B. O’Rourke, Jacques, R., Tseng, H., Marban, E., et al.: A compu-
tational model integrating electrophysiology, contraction, and mitochondrial bioenergetics in
the ventricular myocytes. Biophys. J. 91, 1564–1598 (2006)
137. Costa, K.D., K. May-Newman, Farr, D., O’Dell, W.G., McCulloch, A.D., Omens, J.H.: Three-
dimensional residual strain in midanterior canine left ventricle. Am. J. Physiol. Heart Circ.
Physiol. 42, H1968–H1976 (1997)
138. Costa, K.D., Holmes, J.W., McCulloch, A.D.: Modelling cardiac mechanical properties in
three dimensions. Philos. Trans. R. Soc. Lond. A 359(1783), 1233–1250 (2001)
139. Coudiere, Y., Pierre, C.: Stability, convergence of a finite volume method for two systems
of reaction-diffusion equations in electro-cardiology. Nonlinear Anal. Real World Appl. 7(4),
916–935 (2006)
140. Coveney, P., Diaz, V., Hunter, P., Viceconti, M.: Computational Biomedicine. Oxford
University Press, Oxford (2014)
141. Cronin, J.: Mathematical Aspects of Hodgkin-Huxley Neural Theory. Cambridge University
Press, Cambridge (1987)
142. Cuppen, J.J.M.: Calculating the isochrones of ventricular depolarization. SIAM J. Sci. Stat.
Comput. 5, 105–120 (1984)
374 References
143. Cuppen, J.J.M., van Oosterom, A.: Model studies with the inversely calculated isochrones of
ventricular depolarization. IEEE Trans. Biomed. Eng. 31(10), 652–659 (1984)
144. De Ambroggi, L., Musso, E., Taccardi, B.: Body-surface mapping. In: Macfarlane, P.W.,
Lawrie, T.D.V. (eds.) Comprehensive Electrocardiology, pp. 1015–1049. Pergamon, Oxford
(1989)
145. DeBruin, K.A., Krassowska, W.: Electroporation and shock-induced transmembrane potential
in a cardiac fiber during defibrillation strength shocks. Ann. Biomed. Eng. 26, 584–596 (1998)
146. Dekker, E.: Direct current make and break thresholds for pacemaker electrodes on the canine
ventricle. Circ. Res. 27, 811–823 (1970)
147. Deuflhard, P., Erdmann, B., Roitzsch, R., Lines, G.T.: Adaptive finite element simulation of
ventricular fibrillation dynamics. Comput. Vis. Sci. 12(5), 201–205 (2009)
148. di Bernardo, D., Murray, A.: Computer model for study of cardiac repolarization. J.
Cardiovasc. Electrophys. 11(8), 895–899 (2000)
149. Di Francesco, D., Noble, D.: A model of cardiac electrical activity incorporating ionic pumps
and concentration changes. Philos. Trans. R. Soc. Lond. B 307(1133), 353–398 (1985)
150. Doi, S., Inoue, J., Pan, Z., Tsumoto, K.: Computational Electrophysiology. Springer, Tokyo
(2010)
151. Dryja, M., Widlund, O.B.: Multilevel additive methods for elliptic finite element problems.
In: Hackbusch, W. (ed.) Parallel Algorithms for Partial Differential Equations – Proceedings
of the Sixth GAMM-Seminar, Kiel, 19–21 Jan 1990. Notes on Numerical Fluid Mechanics,
vol. 31, pp. 58–69. Vieweg, Braunschweig (1991). 3-528-07631-3
152. Dryja, M., Widlund, O.B.: Domain decomposition algorithms with small overlap. SIAM J.
Sci. Comput. 15(3), 604–620 (1994)
153. Dryja, M., Sarkis, M.V., Widlund, O.B.: Multilevel Schwarz methods for elliptic problems
with discontinuous coefficients in three dimensions. Numer. Math. 72(3), 313–348 (1996)
154. Ebihara, L., Johnson, E.A.: Fast sodium current in cardiac muscle: a quantitative description.
Biophys. J. 32 779–790 (1980)
155. Efimov, I.R., Ermentrout, B., Huang, D.T., Salama, G.: Activation and repolarization patterns
are governed by different structural characteristics of ventricular myocardium: experimental
study with voltage-sensitive dyes and numerical simulations. J. Cardiovasc. Electrophysiol.
7, 512–530 (1996)
156. Efimov, I.R., Gheng, Y., Van Eagoner, D.R., Mazgalev, T., Tchou, P.J.: Virtual electrode-
induced phase singularity: a basic mechanism of defibrillation failure. Circ. Res. 82, 918–925
(1998)
157. Efimov, I.R., Gray, R.A., Roth, B.J.: Virtual electrodes and deexcitation: new insights into
fibrillation induction and defibrillation. J. Cardiovasc Electrophysiol. 11, 339–353 (2000)
158. Efimov, I.E., Kroll, M.W., Tcho, P.J. (eds.): Cardiac Bioelectric Therapy. Springer, New York
(2009)
159. Einstein, A.: Eine neue Bestimmung der Moleküldimensionen. Ann. Phys. 324, 289–306
(1906)
160. Ejima, J., Martin, D., Engle, C., Sherman, Z., Kunimoto, S., Gettes, L.: Ability of activation
recovery intervals to assess action potential duration during acute no-flow ischemia in the in
situ porcine heart. J. Cardiovasc. Electrophysiol. 99, 832–844 (1998)
161. El-Sherif, N., Caref, E.B., Yin, H., Estivo, M.: The electrophysiological mechanism of
ventricular arrhythmias in the long QT syndrome. Tridimensional mapping of activation and
recovery patterns. Circ. Res. 79, 474–492 (1996)
162. El-Sherif, N., Chinushi, M., Caref, E.B., Restivo, M.: Electrophysiological mechanism of
the characteristic electrocardiographic morphology of torsade de points tachyarrhythmias in
the long-QT syndrome: detailed analysis of ventricular tridimensional activation patterns.
Circulation 96, 4392–4399 (1997)
163. Endresen, L.P., Hall, K., Hoye, J.S., Myrheim, J.: A theory for the membrane potential of
living cells. Eur. Biophys. J. 29, 90–103 (2000)
164. Entcheva, E., Eason, J., Efimov, I.R., Cheng, Y., Malkin, R., Clayton, F.: Virtual electrode
effects in transvenous defibrillation-modulation by structure and interface: evidence from
bidomain simulations an optical mapping. J. Cardiovasc. Electrophysiol. 9, 949–961 (1998)
References 375
165. Ermentrout, B.: Simulating, analyzing, and animating dynamical systems: a guide to XPPAUT
for researchers and students. SIAM, Philadelphia (2002)
166. Ethier, M., Bourgault, Y.: Semi-implicit time-discretization schemes for the Bidomain model.
SIAM J. Numer. Anal. 46(5), 2443–2468 (2008)
167. Evans, L.C., Gariepy, R.F.: Measure Theory and Fine Properties of Functions. Studies in
Advanced Mathematics. CRC, New York (1997)
168. Fall, C.P., Marland, E.S., Wagner, J.M., Tyson, J.J.: Computational Cell Biology, 3rd edn.
Springer, New York (2005)
169. Fast, V.G., Kléber, A.G.: Role of wavefront curvature in propagation of cardiac impulse.
Cardiovasc. Res. 33, 258–271 (1997)
170. Fenton, F.H., Karma, A.: Vortex dynamics in three-dimensional continuous myocardium with
fiber rotation: filament instability and fibrillation. Chaos 8, 20–47 (1998)
171. Fenton, F.H., Cherry, E.M., Hastings, H.M., Evans, S.J.: Multiple mechanisms of spiral wave
breakup in a model of cardiac electrical activity. Chaos 12(3), 852–892 (2002)
172. Fernandez, M.A., Zemzemi, N.: Decoupled time-marching schemes in computational cardiac
electrophysiology and ECG numerical simulation. Math. Biosci. 226, 58–75 (2010)
173. Fife, P.C.: Mathematical Aspect of Reacting and Diffusing Systems. Springer, Berlin (1979)
174. Fife, P.C.: Dynamics of Internal Layers and Diffusive Interfaces. CBMS-NSF Regional
Conference Series in Applied Mathematics, vol. 53. SIAM, Philadelphia (1988)
175. Fife, P.C., McLeod, J.B.: The approach of solutions of nonlinear diffusion equations to
travelling front solutions. Arch. Ration. Mech. Anal. 65, 335–361 (1977)
176. Fish, J.M., Di Diego, J.M., Nesterenko, V., Antzelevich, C.: Epicardial activation of left ven-
tricular wall prolongs QT interval and transmural dispersion of repolarization. Implications
for biventricular pacing. Circulation 109, 2136–2142 (2004)
177. Fischer, G., Tilg, B., Modre, R., Huiskamp, G.J.M., Fetzer, J., Rucker, W., Wach, P.: A
bidomain model based BEM-FEM coupling formulation for anisotropic cardiac tissue. Ann.
Biomed. Eng. 28(10), 1229–1243 (2000)
178. FitzHugh, R.: Impulses and physiological states in theoretical models of nerve membrane.
Biophys. J. 1, 445–466 (1961)
179. FitzHugh, R.: Mathematical models of excitation and propagation in nerve. In: Schwan, H.P.
(ed.) Biological Engineering, pp. 1–85. MacGraw-Hill, New York (1969)
180. Frangi, A.F., Radeva, P.I., Santos, A., Hernandez, M. (eds.): Proceedings of the Third
International Workshop on Functional Imaging and Modeling of the Heart, FIMH’05,
Barcelona, 2–4 June 2005. LNCS, vol. 3504. Springer, Berlin (2005)
181. Franz, M.R.: Monophasic Action Potentials: Bridging Cells to Bedside, pp. 19–45. Futura
Publishing Company, Armonk (2000)
182. Franz, M.R.: What is a monophasic action potential recorded by Franz contact electrode?
Cardiovasc. Res. 65, 940–941 (2005)
183. Franz, M.R., et al.: Monophasic action potential mapping in human subjects with normal
electrocardiograms: direct evidence for the genesis of the T wave. Circulation 75, 379–386
(1987)
184. Frazier, D.W., Krassowska, W., Chen, P.-S., Wolf, P.D., Danieley, N.D., Smith, W.M., Ideker,
R.E.: Transmural activation and stimulus potentials in three-dimensional anisotropic canine
myocardium. Circ. Res. 63, 135–146 (1988)
185. Frazier, D.W., Wolf, P.D., Wharton, J.M., Tang, A.S.L., Smith, W.M., Ideker, R.E.: Stimulus-
induced critical point: mechanism for electrical initiation of reentry in normal canine
myocardium. J. Clin. Invest. 83, 1039–1052 (1989)
186. Fu, Z., Kirby, R.M., Whitaker, R.T.: A fast iterative method for solving the eikonal equation
on tetrahedral domains. SIAM J. Sci. Comput. 35(5), C473–C494 (2013)
187. Fuller, M.S., Sandor, G., Punske, B., Taccardi, B., MacLeod, R.S., Ershler, P.R., Green, L.S.,
Lux, R.L.: Estimates of repolarization dispersion from electrocardiographic measurements.
Circulation 102(6), 685–691 (2000)
188. Fuller, M.S., Sandor, G., Punske, B., Taccardi, B., MacLeod, R.S., Ershler, P.R., Green, L.S.,
Lux, R.L.: Estimates of repolarization and its dispersion from electrocardiographic
376 References
measurements: direct epicardial assessment in the canine heart. J. Electrocardiol. 33(2), 171–
180 (2000)
189. Fuster, V., Walsh, R.A., Harrington, R.A.: Hurst’s the Heart, 13th edn. MacGraw-Hill,
New York (2011)
190. Garbern, J.C., Lee, R.T.: Cardiac stem cell therapy and the promise of heart regeneration. Cell
Stem Cell 12(6), 689–698 (2013)
191. Garfinkel, A., Kim, Y.-H., Voroshilovsky, O., Qu, Z., Kil, J.R., Lee, M.-H.,
Karagueuzian, H.S., Weiss, J.N., Chen, P.-S.: Preventing ventricular fibrillation by flattening
cardiac restitution. Proc. Natl. Acad. Sci. U S A 97(11), 6061–6066 (2000)
192. Gaudesius, G., Miragoli, M., Thomas, S.P., Rohr, S.: Coupling of cardiac electrical activity
over extended distances by fibroblasts of cardiac origin. Circ. Res. 93, 421–428 (2003)
193. Gauthier, L.D., Greenstein, J.L., Winslow, R.L.: Toward an integrative computational model
of the guinea pig cardiac myocyte. Front. Physiol. 3, 244 (2012)
194. Gepstein, L., Hayam, G., Ben-Haim, S.A.: Activation-recovery coupling in the normal swine
endocardium. Circulation 96(11), 4036–4043 (1997)
195. Gerardo Giorda, L., Mirabella, L., Nobile, F., Perego, M., Veneziani, A.: A model-based
block-triangular preconditioner for the Bidomain system in electrocardiology. J. Comput.
Phys. 228(10), 3625–3639 (2009)
196. Gerardo Giorda, L., Perego, M., Veneziani, A.: Optimized Schwarz coupling of Bidomain and
Monodomain models in electrocardiology. Math. Model. Numer. Anal. 45, 309–334 (2011)
197. Geselowitz, D.B.: On the theory of the electrocardiogram. Proc. IEEE 77, 857–876 (1989)
198. Geselowitz, D.B.: Description of cardiac sources in anisotropic cardiac muscle. Application
of bidomain model. J. Electrocardiol. 25, 65–67 (1992)
199. Geselowitz, D.B., Miller, W.T.: A bidomain model for anisotropic cardiac muscle. Ann.
Biomed. Eng. 11, 191–206 (1983)
200. Ghosh, S., Rudy, Y.: Application of L1-norm regularization to epicardial potential solutions
of the inverse electrocardiography problem. Ann. Biomed. Eng. 37(5), 902–912 (2009)
201. Giaquinta, M., Hildebrandt, S.: Calculus of Variations I. Volume 310 of Grundlehren der
mathematischen Wissenschaften. Springer, Berlin (1996)
202. Giga, Y., Goto, S.: Motion of hypersurfaces and geometric equations. J. Math. Soc. Jpn. 44(1),
99–111 (1992)
203. Giga, Y., Sato, M.-H.: Neumann problem for singular degenerate parabolic equations. Differ.
Integral Equ. 6(6), 1217–1230 (1993)
204. Giga, Y., Goto, S., Ishii, H., Sato, M.-H.: Comparison principle and convexity preserving
properties for singular degenerate parabolic equations on unbounded domains. Ind. Univ.
Math. J. 40(2), 443–469 (1991)
205. Gima, K., Rudy, Y.: Ionic current basis of electrocardiographic waveforms. A model study.
Circ. Res. 90, 889–896 (2002)
206. Goldsmith, E.C., et al.: Organization of fibroblasts in the heart. Dev. Dyn. 230(4), 787–794
(2004)
207. Goto, M., Brooks, C.: Membrane excitability of the frog ventricle examined by long pulses.
Am. J. Physiol. Heart Circ. Physiol. 217, H1236–H1245 (1969)
208. Gotoh, M., Uchida, T., Fan, W., Fishbein, M.C., Karagueuzian, H.S., Chen, P.-S.: Anisotropic
repolarization in ventricular tissue. Am. J. Physiol. (Heart Circ. Physiol.) 41, 107–113 (1997)
209. Grandi, E., Pasqualini, F.S., Bers, D.M.: A novel computational model of the human
ventricular action potential and Ca transient. J. Mol. Cell. Cardiol. 48, 112–121 (2010)
210. Greensite, F.: The mathematical basis for imaging cardiac electrical function. CRC Crit. Rev.
Biomed. Eng. 22, 347–399 (1994)
211. Greensite, F., Huiskamp, G.: An improved method for estimating epicardial potentials from
the body surface. IEEE Trans. Biomed. Eng. 45, 98–104 (1998)
212. Griffith, B.E., Peskin, C.S.: Electrophysiology. Commun. Pure. Appl. Math. 66, 1837–1913
(2013)
213. Guan, S., Lu, Q., Huang, K.: A discussion about the Di Francesco-Noble model. J. Theor.
Biol. 189, 27–32 (1997)
References 377
214. Gulrajani, R.M.: Models of the electrical activity of the heart and computer simulation of the
electrocardiogram. CRC Crit. Rev. Biomed. Eng. 16, 1–66 (1988)
215. Gulrajani, R.M.: Bioelectricity and Biomagnetism. Wiley, New York (1998)
216. Gulrajani, R.M., Roberge, F.A., Savard, P.: The inverse problem of electrocardiography. In:
Macfarlane, P.W., Lawrie, T.T.V. (eds.) Comprehensive Electrocardiology, I: chap. 9, pp. 237–
288. Pergamon, Oxford (1989)
217. Han, C., Pogwizd, S.M., Killingsworth, C.R., He, B.: Noninvasive reconstruction of the three-
dimensional ventricular activation sequence during pacing and ventricular tachycardia in the
canice heart. Am. J. Physiol. Heart Circ. Physiol. 302, H244–H252 (2012)
218. Harrild, D.M., Henriquez, C.S.: A finite volume model of cardiac propagation. Ann. Biomed.
Eng. 28(2), 315–334 (1997)
219. Harrild, D.M., Henriquez, C.S.: A computer model of normal conduction in the human atria.
Circ. Res. 87, e25–e36 (2000)
220. Harrild, D.M., Penland, R., Henriquez, C.: A flexible method for simulating cardiac conduc-
tion in three-dimensional complex geometries. J. Electrocardiol. 33(3), 241–251 (2000)
221. Haws, C.W., Lux, R.L.: Correlation between in vivo transmembrane action potential durations
and activation–recovery intervals from electrograms. Circulation 81, 281–288 (1990)
222. He, B., Li, G., Zhang, X.: Noninvasive imaging of cardiac transmembrane potentials within
three-dimensional myocardium by means of a realistic geometry anisotropic heart model.
IEEE Trans. Biomed. Eng. 50(10), 1190–1202 (2003)
223. Heidenreich, E.A., Rodriguez, J.F., Gaspar, F.J., Doblaré, M.: Fourth-order compact schemes
with adaptive time step for monodomain reaction–diffusion equations. J. Comput. Appl.
Math. 216(1), 39–55 (2008)
224. Heijman, J., Volders, P.G.A., Westra, R.L., Rudy, Y.: Local control of ˇ-adrenergic stimu-
lation: effects on ventricular myocyte electrophysiology and Ca2C -transient. J. Mol. Cell.
Cardiol. 50(5), 863–871 (2011)
225. Henriquez, C.S.: Simulating the electrical behavior of cardiac tissue using the bidomain
model. CRC Crit. Rev. Biomed. Eng. 21, 1–77 (1993)
226. Henriquez, C.S., Muzikant, A.L., Smoak, C.K.: Anisotropy, fiber curvature, and bath loading
effects on activation in thin and thick cardiac tissue preparations: simulations in a three-
dimensional bidomain model. J. Cardiovasc. Electrophysiol. 7(5), 424–444 (1996)
227. Herron, T.J., Lee, P., Jalife, J.: Optical imaging of voltage and calcium in cardiac cells and
tissues. Circ. Res. 110, 609–623 (2012)
228. Hille, B.: Ionic Channels of Excitable Membranes, 2nd edn. Sinauer Associates Inc.,
Sunderland (1982)
229. Hodgkin, A., Huxley, A.: A quantitative description of membrane current and its application
to conduction and excitation in nerve. J. Physiol. (Lond.) 117, 500–544 (1952)
230. Hoff, D.: Stability and convergence of finite difference methods for systems of nonlinear
reaction-diffusion equations. SIAM J. Numer. Anal. 15, 1161–1177 (1978)
231. Holland, R.P., Brooks, H.: Precordial end epicardial surface potentials during Myocardial
ischemia in the pig. A theoretical and experimental analysis of the TQ and ST segments.
Circ. Res. 37, 471–480 (1975)
232. Holzapfel, G.A., Ogden, R.W.: Constitutive modelling of passive myocardium: a structurally
based framework for material characterization. Philos. Trans. R. Soc. A 13(367), 3445–3475
(2009)
233. Hooke, N.: Efficient simulation of action potential propagation in a bidomain. Ph.D. thesis,
Department of Computer Science, Duke University (1992)
234. Hooks, D.A., Tomlinson, K.A., Mardsen, S.G., LeGrice, I.J., Smaill, B.H., Pullan, A.J.,
Hunter, P.J.: Cardiac microstructure. Implications for electrical propagation and defibrillation
in the heart. Circ. Res. 91, 331–338 (2002)
235. Hopenfeld, B., Stinstra, J.G., MacLeod, R.S.: Mechanism for ST depression associated with
contiguous subendocardial ischemia. J. Cardiovasc. Electrophysiol. 29, 1200–1206 (2004)
236. Hormander, L.: The Analysis of Linear Partial Differential Operators. Springer, Berlin (1983)
378 References
237. Hoyt, R.H., Cohen, M.L., Saffitz, J.E.: Distribution and three-dimensional structure of
intercellular junctions in canine myocardium. Circ. Res. 64, 563–574 (1989)
238. Hsiao, G.C., Wendland, W.L.: Boundary Integral Equations. Springer, Berlin (2008)
239. http://commons.wikimedia.org/wiki/File:Wiggers_Diagram.png
240. http://www.texasheartinstitute.org/HIC/Anatomy/images/fig1_crosslg.jpg
241. Huiskamp, G., Greensite, F.: A new method for myocardial activation imaging. IEEE Trans.
Biomed. Eng. 44, 433–446 (1997)
242. Humphrey, J.D.: Cardiovascular Solid Mechanics. Cells, Tissues, and Organs. Springer,
New York (2002)
243. Hund, T.J., Rudy, Y.: Rate transient and regulation of action potential and calcium transient
in a canine cardiac ventricular cell model. Circulation 110, 3168–3174 (2004)
244. Hund, T.J., Kucera, J.P., Otani, N.F., Rudy, Y.: Ionic charge conservation and long-term steady
state in hte Luo–Rudy dynamic cell model. Biophys. J. 81, 3324–3331 (2001)
245. Hunter, P.J., Nash, M.P., Sands, G.B.: Computational electromechanics of the heart. In:
Panfilov, A.V., Holden, A.V. (eds.) Computational Biology of the Heart. Wiley, New York
(1997)
246. Hunter, P.J., McCulloch, A.D., ter Keurs, H.: Modelling the mechanical properties of cardiac
muscle. Prog. Biophys. Mol. Biol. 69(2–3), 289–331 (1998)
247. Hunter, P.J., et al.: A vision and strategy for the virtual physiological human in 2010 and
beyond. Philos. Trans. R. Soc. A 368, 2595–2614 (2010)
248. Hunter, P.J., et al.: A vision and strategy for the virtual physiological human: 2012 update.
Interface Focus 3, 1–9 (2013)
249. Isakov, V.: Inverse Problems for Partial Differential Equations, 2nd edn. Springer, Berlin
(2006)
250. Jack, J.J.B., Noble, D., Tsien, R.W.: Electric Current Flow in Excitable Cells. Clarendon,
Oxford (1983)
251. Jacquemet, V.: An eikonal approach for the initiation of reentrant cardiac propagation in
reaction-diffusion models. IEEE Trans. Biomed. Eng. 57(9), 2090–2098 (2010)
252. Jacquemet, V., Henriquez, C.: Finite volume stiffness matrix for solving anisotropic cardiac
propagation in 2-D and 3-D unstructured meshes. IEEE Trans. Biomed. Eng. 52(8), 1490–
1492 (2005)
253. Jacquemet, V., Kappenberger, L., Henriquez, C.S.: Modeling atrial arrhythmias: impact on
clinical diagnosis and therapies. IEEE Rev. Biomed. Eng. 1, 94–114 (2008)
254. Jafri, S., Rice, J.J., Winslow, R.L.: Cardiac Ca2C dynamics: the role of ryanodine receptor
adaptation and sarcoplasmic reticulum load. Biophys. J. 74, 1149–1168 (1998)
255. Janks D.L., Roth B.J.: Quaterfoil reentry caused by bursting pacing. J. Cardiovasc. Electro-
physiol. 17, 1362–1368 (2006)
256. Janks, D.L., Roth, B.J.: The bidomain theory of pacing. In: Efimov I.R., Kroll M.W., Tchou, J.
(eds.) Cardiac Bioelectric Therapy, chap. 2.1, pp. 63–83. Springer, New York (2009)
257. Janse, M.J.: ST segment mapping and infarct size. Cardiovasc. Res. 45, 190–193 (2000)
258. Janse, M., Sosunov, E.A., Corornel, R., Opthof, T., Anyukhovsky, E.P., de Bakker J.M.T.,
Plotnikov, A.N., Shlapakova, I.N., Danilo, P., J.Tijssen, G.P., Rosen, M.R.: Repolarization
gradients in the canine left ventricle before and after induction of short-term cardiac memory.
Circulation 112, 1711–1718 (2005)
259. Jerome, J.W.: Convergence of successive iterative semidiscretizations for FitzHugh-Nagumo
reaction systems. SIAM J. Numer. Anal. 27, 2054–1065 (1984)
260. Jikov, V.V., Kozlov, S.M., Oleinik, O.A.: Homogenization of Differential Operators and
Integral Functionals. Springer, Berlin (1994)
261. Johnston, P.: Computational Inverse Problems in Electrocardiography. WIT, Southampton
(2001)
262. Johnston, P.R., Kilpatrick, D.: The effect of conductivity values on ST segment shift in
subendocardial ischaemia. IEEE Trans. Biomed. Eng. 50, 150–158 (2003)
263. Johnston, P.R., Kilpatrick, D., Li, C.Y.: The importance of anisotropy in modeling ST segment
shift in subendocardial ischaemia. IEEE Trans. Biomed. Eng. 48, 1366–1376 (2001)
References 379
264. Joyner, R.W.: Modulation of repolarization by electrotonic interactions. Jpn. Heart J. 27, 167–
183 (1986)
265. Jungschleger, J.G., Vos, M.A.: Hybrid action potential etiology. J. Cardiovasc. Electrophysiol.
11(8), 946–948 (2000). (Letter to the Editor)
266. Katila, T., Magnin, I.E., Clarysse, P., Montagnat, J., Nenonen, J. (eds.): Proceedings of the
First International Workshop on Functional Imaging and Modeling of the Heart, FIMH’01,
Helsinki, 15–16 Nov 2001. LNCS, vol. 2230. Springer, Berlin (2001)
267. Katz, A.M.: Physiology of the Heart. Wolters Kluwer, Philadelphia (2011)
268. Keener, J.P.: An eikonal-curvature equation for action potential propagation in myocardium.
J. Math. Biol. 29, 629–651 (1991)
269. Keener, J.P.: Direct activation and defibrillation of cardiac tissue. J. Theor. Biol. 178, 313–324
(1996)
270. Keener, J.P., Bogar, K.: A numerical method for the solution of the bidomain equations in
cardiac tissue. Chaos 8(1), 234–241 (1998)
271. Keener, J.P., Panfilov, A.V.: Three–Dimensional propagation in the heart: the effects of
geometry and fiber orientation on propagation in myocardium. In: Zipes, D.P., Jalife, J. (eds.)
Cardiac Electrophysiology: From Cell to Bedside, pp. 335–347. W. B. Sounders, Philadelphia
(1995)
272. Keener, J.P., Panfilov, A.V.: The effects of geometry and fibre orientation on propagation and
extracellular potentials in myocardium. In: Panfilov, A.V., Holden, A.V. (eds.) Computational
Biology of the Heart, chap. 8, pp. 235–258. Wiley, New York (1997)
273. Keener, J.P., Sneyd, J.: Mathematical Physiology, 2nd edn. Springer, New York (2008)
274. Kerckhoffs, R.C.P. (ed.): Patient-Specific Modeling of the Cardiovascular System:
Technology-Driven Personalized Medicine. Springer, New York (2010)
275. Kerckhoffs, R.C.P., Bovendeerd, B.H.M., Kotte, J.C.S., Prinzen, F.W., Smits, K., Arts, T.:
Homogeneity of cardiac contraction despite physiological asynchrony of depolarization: a
model study. Ann. Biomed. Eng. 31, 536–547 (2003)
276. Kleber, A.G.: ST-segment elevation in the electrocardiogram: a sign of myocardial ischemia.
Cardiovasc. Res. 45, 111–118 (2000)
277. Kleber, A.G., Riegger, C.B.: Electrical constants of arterially perfused rabbit papillar muscle.
J. Physiol. 385, 307–324 (1987)
278. Kleber, A.G., Rudy, Y.: Basic mechanisms of cardiac impulse propagation and associated
arrhythmias. Physiol. Rev. 84(2), 431–488 (2004)
279. Kleber, A.G., Janse, M.J., van Capelle, F.J.L., Durrer, D.: Mechanism and time course of
S-T and T-Q segment changes during acute regional myocardial ischemia in the pig heart
determined by extracellular and intracellular recordings. Circ. Res. 42(5), 603–613 (1978)
280. Kneller, J., Ramirez, R.J., Chartier, D., M.Courtemanche, Nattel, S.: Time-dependent tran-
sients in an ionically based mathematical model of the canine atrial action potential. Am. J.
Physiol. Heart. Circ. Physiol. 282, H1437–H1451 (2002)
281. Knisley, S.B.: Transmembrane voltage changes during unipolar stimulation of rabbit ventri-
cle. Circ. Res. 77(6), 1229–1239 (1995)
282. Koch, C.: Biophysics of Computation. Oxford University Press, New York (1999)
283. Kogan, B.Y.: Introduction to Computational Cardiology. Springer, New York (2010)
284. Kohl, P., Sachs, F., Franz, M.R.: Cardiac Mechano-Electric Coupling and Arrhythmias.
Oxford University Press, Oxford (2011)
285. Kondo, M., Nesterenko, V., Antzelevitch, C.: Cellular basis for the monophasic action
potential. Which electrode is the recording electrode? Cardiovasc. Res. 63, 635–644 (2004)
286. Kjekshus, J.K., Maroko, P.R., Sobel, B.E.: Distribution of myocardial injury and its relation
to epicardial ST-segment changes after coronary artery occlusion in the dog. Cardiovasc. Res.
6, 490–499 (1972)
287. Krogh-Masden, T., Cristini, D.J.: Nonlinear dynamics in cardiology. Ann. Rev. Biomed. Eng.
14, 179–203 (2012)
288. Kuznetsov, Y.: Elements of Applied Bifurcation Theory. Springer, New York (2000)
289. Laflamme, M.A., Murry, C.E.: Heart regeneration. Nature 473 326–335 (2011)
380 References
290. Lang, J.: Adaptive Multilevel Solution of Nonlinear Parabolic PDE Systems. Theory,
Algorithms, and Applications. LNCSE, vol. 16. Springer, Berlin (2000)
291. Leeson, P.: Cardiovascular Imaging. Oxford University Press, Oxford (2011)
292. LeGrice, I.J., Smaill, B.H., Chai, L.Z., Edgar, S.G., Gavin, J.B., Hunter, P.J.: Laminar
structure of the heart: ventricular myocyte arrangement and connective tissue architecture
in the dog. Am. J. Physiol. Heart Circ. Physiol. 269(38), H571–H582 (1995)
293. LeGrice, I.J., Smaill, B.H., Hunter, P.J.: Laminar structure of the heart: a mathematical model.
Am. J. Physiol. Heart Circ. Physiol. 272(41), H2466–H2476 (1997)
294. LeGrice, I.J., Hunter, P.J., Young, A., Smaill, B.H.: The architecture of the heart: a data-based
model. Philos. Trans. R. Soc. Lond. A 359, 1217–1232 (2001)
295. Leon, L.J., Horacek, B.M.: Computer model of excitation and recovery in the anisotropic
myocardium, I: rectangular and cubic arrays of excitable elements. II: excitation in the
simplified left ventricle III: arrhythmogenic conditions in the simplified left ventricle. J.
Electrocardiol. 14, 1–15, 17–31, 33–41 (1991)
296. Leri, A., Kajstura, J., Anversa, P.: Role of cardiac stem cells in cardiac phatophysiology: a
paradigm shift in human myocardial biology. Circ. Res. 109, 941–961 (2011)
297. Lesh, M.D., Spear, J.F., Simson, M.B.: A computer model of the electrogram: what causes
fractionation? J. Electrocardiol. 21(Suppl), S69–S73 (1988)
298. Li, D., Li, C.Y., Yong, A.C., Johnston, P.R., Kilpatrick, D.: Epicardial ST depression in acute
myocardial ischemia. Circ. Res. 85, 959–964 (1999)
299. Li, L., Niederer, S., et al.: A mathematical model of the murine ventricular myocyte: a data-
driven biophysically based approach applied to mice overexpressing the canine NCX isoform.
Am. J. Physiol. HC 299, H1045–H1063 (2010)
300. Lindemans, F.W., van der Gon, J.J.D.: Current threshold and liminal size in excitation of heart
muscle. Cardiovasc. Res. 12(8), 477–485 (1978)
301. Lindemans, F.W., Heethaar, R.M., van der Gon, J.J.D., Zimmerman, A.N.E.: Site of initial
excitation and current threshold as a function of electrode radius in heart muscle. Cardiovasc.
Res. 9, 95–104 (1975)
302. Lines, G.T., Grottum, P., Tweito, A.: Modeling the electric activity of the heart: a bidomain
model of the ventricles embedded in a torso. Comput. Vis. Sci. 5, 195–213 (2003)
303. Linge, S., Sundnes, J., Hanslien, M., Lines, G.T., Tveito, A.: Numerical solution of the
bidomain equations. Philos. Trans. R. Soc. A 367(1895), 1931–1950 (2009)
304. Lions, J.L.: Optimal Control of Systems Governed by Partial Differential Equations. Springer,
Berlin (1971)
305. Lions, J.L., Magenes, E.: Nonhomogeneous boundary value problems and applications.
I. Springer, Berlin (1972)
306. Livshitz, L.M., Rudy, Y.: Regulation of Ca2C and electrical alternans in cardiac myocytes:
role of CAMKII and repolarizing currents. Am. J. Physiol. Heart Circ. Physiol. 292, H2854–
H2866 (2007)
307. Livshitz, L., Rudy, Y.: Uniqueness and stability of action potential models during rest, pacing,
and conduction using problem-solving environment. Biophys. J. 97, 1265–1276 (2009)
308. Luo, C., Rudy, Y.: A model of the ventricular cardiac action potential: depolarization,
repolarization, and their interaction. Circ. Res. 68(6), 1501–1526 (1991)
309. Luo, C., Rudy, Y.: A dynamic model of the cardiac ventricular action potential. I. Simulations
of ionic currents and concentration changes. Circ. Res. 74(6), 1071–1096 (1994)
310. Luo, C., Rudy, Y.: A dynamic model of the cardiac ventricular action potential. II. Afterde-
polarizations, triggered activity, and potentiation. Circ. Res. 74(6), 1097–1113 (1994)
311. Lux, R.L., Gettes, L., Mason, J.W.: Understanding proarrhythmic potential in therapeutic drug
development: alternate strategies for measuring and tracking repolarization. J. Electrocardiol.
39, S161–S164 (2006)
312. Macfarlane, P.W., van Oosterom, A., Janse, M., Kligfield, P., Camm, J., Pahlm, O. (eds.):
Basic Electrocardiology. Cardiac Electrophysiology, ECG Systems and Mathematical Mod-
eling. Springer, New York (2012)
References 381
313. MacLachlan, M.C., Sundnes, J., Lines, G.T.: Simulation of ST segment changes during
subendocardial ischemia using a realistic 3-D cardiac geometry. IEEE Trans. Biomed. Eng.
52, 799–807 (2005)
314. MacLeod, R.S., Punske, B., Yilmaz, B., Shome, S., Taccardi, B.: The role of heart rate in
myocardial ischemia from restricted coronary perfusion. J. Electrocardiol. 34, 43–51 (2001)
315. Mahajan, A., Shiferaw, Y., Sato, D., Baher, A., Olcese, R., Xie, L.H., Yang, M.J., Chen, P.S.,
Restrepo, J.G., Karma, A., Garfinkel, A., Qu, Z., Weiss, J.N.: A rabbit ventricular action
potential model replicating cardiac dynamics at rapid heart rates. Biophys. J. 94(2), 392–410
(2008)
316. Magnin, I.E., Montagnat, J., Clarysse, P., Nenonen, J., Katila, T. (eds.): Proceedings of the
Second International Workshop on Functional Imaging and Modeling of the Heart, FIMH’03,
Lyon, 5–6 June 2003. LNCS, vol. 2674. Springer, Berlin (2003)
317. Malmivuo, J., Plonsey, R.: Bioelectromagnetism. Oxford University Press, Oxford (1995)
318. Mardal, K.A., Sundnes, J., Langtangen, H.P., Tveito, A.: Systems of PDEs and block
preconditioning. In: Langtangen, H.P., Tveito, A. (eds.) Advanced Topics in Computational
Partial Differential Equations. LNCSE, vol. 33, chap. 5, pp. 200–236. Springer, Berlin (2004)
319. Mardal, K.-A., Nielsen, B.F., Cai, X., Tveito, A.: An order optimal solver for the discretized
bidomain equations. Numer. Linear Algebr. Appl. 14(2), 83–98 (2007)
320. Markowich, P.A.: The Stationary Semiconductor Device Equations. Springer, Berlin (1986)
321. Mascagni, M.: The backward euler method for numerical solution of the Hodgkin–Huxley
equations of nerve conduction. SIAM J. Numer. Anal. 27(4), 941–962 (1990)
322. Matta, R.J., Verrier, R.L., Lown, B.: Repetitive extrasystole as an index of vulnerability to
ventricular fibrillation. Am. J. Physiol. 230, 1469–1473 (1976)
323. McAllister, R.E., Noble, D., Tsien, R.W.: Reconstruction of the electrical activity of cardiac
Purkinje fibres. J. Physiol. 251, 1–59 (1975)
324. Mehra, R., Furman, S.: Comparison of cathodal, anodal, and bipolar strength-interval curves
with temporary and permanent pacing electrodes. Br. Heart J. 41, 468–476 (1979)
325. Messnarz, B., Seger, M., Modre, R., Fischer, G., Hanser, F., Tilg, B.: A comparison of
noninvasive reconstruction of epicarial versus transmembrane potentials in consideration of
the null space. IEEE Trans. Biomed. Eng. 51(9), 1609–1618 (2004)
326. Messnarz, B., Tilg, B., Modre, R., Fischer, G., Hanser, F.: A new spatiotemporal regularization
approach for reconstruction of cardiac transmembrane potential patterns. IEEE Trans.
Biomed. Eng. 51(2), 273–281 (2004)
327. Metaxas, D.N., Axel, L. (eds.): Proceedings of the 6th International Conference on Functional
Imaging and Modeling of the Heart, FIMH’11, New York City, 25–27 May 2011. LNCS,
vol. 6666. Springer, Berlin (2011)
328. Millar, C.K., Kralios, F.A., Lux, R.L.: Correlation between refractory periods and activation-
recovery intervals from electrograms – effects of rate and adrenergic interventions. Circula-
tion 72, 1372–1379 (1985)
329. Miller, W.T., Geselowitz, D.B.: Simulation studies of the electrocardiogram I. The normal
heart. Circ. Res. 43(2), 301–315 (1978)
330. Miragoli, M., Gaudesius, G., Rohr, S.: Electrotonic modulation of cardiac impulse conduction
by myofibroblasts. Circ. Res. 98, 801–810 (2006)
331. Mirams, G.R., et al.: Chaste: an open source C plus plus library for computational physiology
and biology. PLoS Comput. Biol. 9(3), e100297 (2013)
332. Miranda, C.: Partial Differential Equations of Elliptic Type. Springer, Berlin (1970)
333. Mitchell, C.C., Schaeffer, D.G.: A two-current model for the dynamics of cardiac membrane.
Bull. Math. Biol. 65(5), 767–793 (2003)
334. Miura, R.H.: Accurate computation of stable solitary waves for the FitzHugh–Nagumo
equations. J. Math. Biol. 13, 247–269 (1982)
335. Modre, R., Tilg, B., Fischer, G., Wach, P.: Noninvasive myocardial activation time imaging:
a novel inverse algorithm applied to clinical ECG mapping data. IEEE Trans. Biomed. Eng.
49(10), 1153–1161 (2002)
382 References
336. Moore, P.K.: An adaptive finite element method for parabolic differential systems: some
algorithmic considerations in solving in three space dimensions. SIAM J. Sci. Comput. 21(4),
1567–1586 (2000)
337. Morris, C., Lecar, H.: Voltage oscillations in the barnacle giant muscle. Biophys. J. 35, 193–
213 (1981)
338. Munteanu, M.: Overlapping additive Schwarz methods for nonlinear parabolic reaction-
diffusion problems. Ph.D. thesis, Department of Mathematics, University of Milano (2008)
339. Munteanu, M., Pavarino, L.F.: Implicit parallel solvers in computational electrocardiology.
In: Carja, O., Vrabie, I.I. (eds.) Applied Analysis and Differential Equations, pp. 255–266.
World Scientific, Singapore (2007)
340. Munteanu, M., Pavarino, L.F.: Decoupled Schwarz algorithms for implicit discretization of
nonlinear Monodomain and Bidomain systems. Math. Model Methods Appl. Sci. 19(7),
1065–1097 (2009)
341. Munteanu, M., Pavarino, L.F., Scacchi, S.: A scalable Newton-Krylov-Schwarz method for
the Bidomain reaction-diffusion system. SIAM J. Sci. Comput. 31(5), 3861–3883 (2009)
342. Murillo, M., Cai, X.: A fully implicit parallel algorithm for simulating the nonlinear electrical
activity of the heart. Numer. Linear Algebr. Appl. 11, 261–277 (2004)
343. Murthy, M.K.V., Stampacchia, G.: Le problème de Dirichlet pour les équations elliptiques du
second ordreà coefficients discontinus. Ann. Ist. Fourier XV, 189–258 (1965)
344. Muzikant, A., Hsu, E.W., Wolf, P.D., Henriquez, C.S.: Region specific modeling of cardiac
muscle: comparison of simulated and experimental potentials. Ann. Biomed. Eng. 30, 867–
883 (2002)
345. Nayak, A.R., Shajahan, T.K., Panfilov, A.V., Pandit, R.: Spiral-wave dynamics in a mathemat-
ical model of human ventricular tissue with myocytes and fibroblasts. PLoS One 8(9), e72950
(2013)
346. Nelson, C.V., Gezelowitz, D.B.: The Theoretical Basis of Electrocardiology. Clarendon,
Oxford (1976)
347. Nesterenko, V.V., Kondo, M., Antzelevitch, C.: Biophysical basis for monophasic action
potential. Cardiovasc. Res. 65, 942–944 (2005)
348. Neu, J.S., Krassowska, W.: Homogenization of syncitial tissues. CRC Crit. Rev. Biomed. Eng.
21(2), 137–199 (1993)
349. Neunlist, M., Tung, L.: Spatial distribution of cardiac transmembrane potentials around an
extracellular electrode: dependence on fiber orientation. Biophys. J. 68, 2310–2311 (1995)
350. Ni, Q., MacLeod, R.S., Punske, B.B., Taccardi, B.: Computing and visualizing electric
potentials and current pathways in the Thorax. J. Electrocard. 33, 189–197 (2000)
351. Niederer, S., Smith, N.: A mathematical model of the slow force response to stretch in rat
ventricular myocytes. Biophys. J. 92, 4030–4044 (2007)
352. Nielsen, I.J., Le Grice, P.M.F., Hunter, P.J., Smaill, B.H.: Mathematical model of geometry
and fibrous structure of the heart. Am. J. Physiol. Heart Circ. Physiol. 260, H1365–H1378
(1991)
353. Nielsen, B.F., Cai, X., Sundnes, J., Tveito, A.: Toward a computational method for imaging
the extracellular potassium concentration during regional ischemia. Math. Biosci. 220, 118–
130 (2009)
354. Nielsen, B.F., Lysaker, M., Grøttum, P.: Computing ischemic regions in the heart with the
bidomain model: first step toward validation. IEEE Trans. Med. Imag. 32(6), 1085–1096
(2013)
355. Noble, D.: A modification of the Hodgkin-Huxley equations applicable to Purkinje fibre
action and pacemaker potentials. J. Physiol. 160, 317–352 (1962)
356. Noble, D., Rudy, Y.: Models of cardiac ventricular action potentials: iterative interaction
between experiment and simulation. Philos. Trans. R. Soc. Lond. A 359, 1127–1142 (2001)
357. Noble, D., Noble, S., Bett, C., Earm, Y.E., Ko, W.K., So, I.K.: The role of sodium-calcium
exchange during the cardiac action potential. Ann. NY Acad. Sci. 639, 334–354 (1991)
References 383
358. Noble, D., Varghese, A., Kohl, P., Noble, P.J.: Improved guinea-pig ventricular cell model
incorporating a diadic space, iKr & iKs, and lenght- & tension-dependent processes. Can.
J. Cardiol. 14, 123–134 (1998)
359. Nochetto, R.H., Savarè, G., Verdi, C.: A posteriori error estimates for variable time-step
discretizations of nonlinear evolution equations. Commun. Pure Appl. Math. 53(5), 525–589
(2000)
360. Nygren, A., Fiset, C., Firek, L., Clark, J.W., Lindblad, D.S., Clark, R.B., Giles, W.R.:
Mathematical Model of an Adult Human Atrial Cell. Circ. Res. 82, 63–81 (1998)
361. O’Hara, T., Virag, L., Varro, A., Rudy, Y.: Simulation of the undiseased human cardiac
ventricular action potential: model formulation and experimental validation. PLoS Comput.
Biol. 7(5), 1–29 (2011)
362. Oleinik, O.A., Shamaev, A.S., Yosifian, G.A.: Mathematical Problems in Elasticity and
Homogenization. North-Holland, Amsterdam (1992)
363. Opthof, T., Coronel, R., Wilms-Schopman, F.J.G., Plotnikov, A.N., Shlapakova, I.N.,
Danilo, P., Rosen, M.R., Janse, M.J.: Dispersion of repolarization in canine ventricle and
the electrocardiographic T wave: Tpe interval does not reflect transmural dispersion. Heart
Rhythm 4, 341–348 (2007)
364. Osaka, T., Kodama, I., Tsuboi, N., Toyama, J., Yamada, K.: Effects of activation sequence and
anisotropic cellular geometry on the repolarization phase of action potential of dog ventricular
muscle. Circulation 76(1), 226–236 (1987)
365. Osher, S., Fedkin, R.: Level Set Methods and Dynamic Implicit Surfaces. Applied Mathemat-
ical Sciences, vol. 153. Springer, New York (2003)
366. Ourselin, S., Rueckert, D., Smith, N. (eds.): Proceedings of the 7th International Conference
on Functional Imaging and Modeling of the Heart, FIMH’13, London, 20–22 June 2013.
LNCS, vol. 7945. Springer, Berlin (2013)
367. Pandit, S.V., Clark, R.B., Giles, W.R., Demir, S.S.: A mathematical model of action potential
heterogeneity in adult rat left ventricular myocytes. Biophys. J. 81(6), 3029–3051 (2001)
368. Panfilov, A.V.: Spiral breakup as a model of ventricular fibrillation. Chaos 8, 57–64 (1998)
369. Panfilov, A.V., Holden, A.V.: Computational Biology of the Heart. Wiley, New York (1997)
370. Park, J.-H., Jerome, J.W.: Qualitative properties of steady-state Poisson-Nernst-Planck sys-
tems: mathematical study. SIAM J. Appl. Math. 57(3), 609–630 (1997)
371. Pavarino, L.F., Colli Franzone, P.: Parallel solution of cardiac reaction-diffusion models. In:
Kornhuber, R., et al. (eds.) Domain Decomposition Methods in Science and Engineering.
LNCSE, vol. 40, pp. 669–676. Springer, Berlin (2004)
372. Pavarino, L.F., Scacchi, S.: Multilevel additive Schwarz preconditioners for the Bidomain
reaction-diffusion system. SIAM J. Sci. Comput. 31(1), 420–443 (2008)
373. Pavarino, L.F., Scacchi, S.: Parallel multilevel Schwarz and block preconditioners for the
Bidomain parabolic-parabolic and parabolic-elliptic formulations. SIAM J. Sci. Comput.
33(4), 1897–1919 (2011)
374. Payne, L.E.: Improperly Posed Problems in Partial Differential Equations. PA Saunders,
Philadelphia (1975)
375. Penland, R., Harrild, D., Henriquez, C.: Modeling impulse propagation and extracellular
potential distributions in anisotropic cardiac tissue using a finite volume element discretiza-
tion. Comput. Vis. Sci. 4, 215–226 (2002)
376. Pennacchio, M.: A nonconforming domain decomposition method for the cardiac potential
problem. In: Proceedings of IEEE Computers in Cardiology, Rotterdam, 23–26 Sept 2001,
vol. 28, pp. 537–540 (2001)
377. Pennacchio, M.: The mortar finite element method for the cardiac “bidomain” model of
extracellular potential. J. Sci. Comput. 20(2), 191–210 (2004)
378. Pennacchio, M., Savarè, G., Colli Franzone, P.: Multiscale modeling for the electrical activity
of the heart. SIAM J. Math. Anal. 37(4), 1333–1370 (2006)
379. Pennacchio, M., Simoncini, V.: Efficient algebraic solution of reaction-diffusion systems for
the cardiac excitation process. J. Comput. Appl. Math. 145, 49–70 (2002)
384 References
380. Pennacchio, M., Simoncini, V.: Substructuring preconditioners for mortar discretization of
degenerate evolution problem. J. Sci. Comput. 36, 391–419 (2008)
381. Pennacchio, M., Simoncini, V.: Algebraic multigrid preconditioners for the bidomain
reaction–diffusion system. Appl. Numer. Math. 59, 3033–3050 (2009)
382. Pennacchio, M., Simoncini, V.: Fast structured AMG preconditioning for the bidomain model
in electrocadiology. SIAM J. Sci. Comput. 33, 721–745 (2011)
383. Peskin, C.S.: Mathematical Aspects of Heart Physiology. Lecture Notes of the Courant
Institute of Mathematical Sciences, New York University, New York (1975)
384. Peskin, C.S.: Numerical analysis of blood flow in the heart. J. Comput. Phys. 25(3), 220–252
(1977)
385. Peskin, C.S., McQueen, D.M.: Cardiac fluid dynamics. Crit. Rev. Biomed. Eng. 20, 451–459
(1992)
386. Pilkington, T.C., Plonsey, R.: Engineering Contributions to Biophysical Electrocardiography.
IEEE, New York (1982)
387. Pilkington, T.C., Loftis, B., Palmer, T., Budinger, T.F.: High-Performance Computing in
Biomedical Research. CRC, Boca Raton (1993)
388. Plank, G., Burton, R.A.B., Hales, P., Bishop, M., Mansoori, T., Bernabeu, M.O., Garny,
A., Prassl, A.J., Bollendorsff, C., Mason, F., Mahmood, F., Rodriguez, B., Grau, V.,
Schneider, J.E., Gavaghan, D., Kohl, P.: Generation of histo-anatomically representative
models of the individual heart: tools and application. Philos. Trans. R. Soc. A 367(1895),
2257–2292 (2009)
389. Plank, G., Liebmann, M., Weber dos Santos, M.O., Vigmond, E.J., Haase, G.: Algebraic
Multigrid Preconditioner for the Cardiac Bidomain Model. IEEE Trans. Biomed. Eng. 54(4),
585–596 (2007)
390. Plank, G., Prassl, A., Hofer, E., Trayanova, N.A.: Evaluating intramural virtual electrodes in
the myocardial wedge preparation: simulations of experimental conditions. Biophys. J. 94,
1904–1915 (2008)
391. Plonsey, R.: Bioelectric sources arising in excitable fibers (Alza lecture). Ann. Biomed. Eng.
16, 519–546 (1988)
392. Plonsey, R., Barr, R.C.: Current flow patterns in two-dimensional anisotropic bisyncitia with
normal and extreme conductivities. Biophys. J. 45, 557–571 (1984)
393. Plonsey, R., Barr, R.C.: Bioelectricity: A Quantitative Approach. Springer, New York (2007)
394. Plonsey, R., Heppner, D.: Consideration of quasi-stationarity in electrophysiological systems.
Bull. Math. Biophys. 29, 657–664 (1967)
395. Poelzing, S., Rosenbaum, D.S.: Heterogeneous connexin43 expression produces electrophys-
iological heterogeneities across ventricular wall. Am J. Physiol. Heart Circ. Physiol. 286,
H2001–H2009 (2004)
396. Poelzing, S., Rosenbaum, D.S.: Altered connexin43 expression produces arrhythmia substrate
in heart failure. Am. J. Physiol (Heart Circ. Physiol) 287, H1762–H1770 (2004)
397. Pollard, A.E., Hooke, N., Henriquez, C.S.: Cardiac propagation simulation. CRC Crit. Rev.
Biomed. Eng. 20(3–4), 171–210 (1992)
398. Pollard, A.E., Burgess, M.J., Spitzer, K.W.: Computer simulations of three-dimensional
propagation in ventricular myocardium. Circ. Res. 72(4), 744–756 (1993)
399. Potse, M., Dubè, B., Richer, J., Vinet, A., Gulrajani, R.: A comparison of Monodomain and
Bidomain reaction–diffusion models for action potential propagation in the human heart.
IEEE Trans. Biomed. Eng. 53(12), 2425–2434 (2006)
400. Potse, M., Vinet, A., Opthof, T., Coronel, R.: Validation of a simple model for the morphology
of the T wave in unipolar electrograms. Am. J. Physiol. Heart Circ. Physiol. 297(2), H792–
H801 (2009)
401. Pressler, M.L., Munster, P.N., Huang, X.-D.: Gap junction distribution in the heart: functional
relevance. In: Zipes, D., Jalife, J. (eds.) Cardiac Electrophysiology, chap. 16, pp. 144–151.
W. B. Sauders, Philadelphia (1995)
402. Priebe, L., Beuckelmann, D.J.: Simulation study of cellular electrical properties in heart
failure. Circ. Res. 82, 1206–1223 (1998)
References 385
403. Prinzmetal, M., Toyoshima, A., Ekmekci, Y., Mizumo, Y., Nagaya, T.: Myocardial ischemia.
Nature of ischemic electrocardiographic patterns in the mammalian ventricles as determined
by intracellular electrographic and metabolic changes. Am. J. Cardiol. 8, 493–503 (1961)
404. Prior, P., Roth, B.J.: Calculation of optical signal using three-dimensional bidomain/diffusion
model reveals distortion of the transmembrane potential. Biophys. J. 95, 2097–2102 (2008)
405. Puglisi, J.L., Bers, D.M.: LabHEART: an interactive computer model of rabbit ventricular
myocyte ion channels and Ca transport. Am. J. Physiol. Cell Physiol. 281(6), C2049–C2060
(2001)
406. Pullan, A.J., Buist, M.L., Cheng, L.K.: Mathematically Modelling the Electrical Activity of
the Heart. World Scientific, Singapore (2005)
407. Pullan, A.J., Cheng, L.K., Nash, M.P., Ghodrati, A., MacLeod, R.S., Brooks, D.H.: The
inverse problem of electrocardiography. In: Comprehensive Electrocardiology, pp. 299–344.
Springer, New York (2010)
408. Punske, B.B., Ni, Q., Lux, R.L., MacLeod, R.S., Ershler, P.R., Dustman, T.J., Allison, M.J.,
Taccardi, B.: Spatial methods of epicardial activation time determination in normal hearts.
Ann. Biomed. Eng. 31(7), 781–792 (2003)
409. Puwal, S., Roth, B.J.: Forward Euler stability of the bidomain model of cardiac tissue. IEEE
Trans. Biomed. Eng. 54(5), 951–953 (2007)
410. Qu, Z., Garfinkel, A.: An advanced algorithm for solving partial differential equation in
cardiac conduction. IEEE Trans. Biomed. Eng. 46(9), 1166–1168 (1999)
411. Qu, Z., Kill, J., Xie, F., Garfinkel, A., Weiss, J.N.: Scroll wave dynamics in a three-
dimensional cardiac tissue model: roles of restitution, thickness, and fiber rotation. Biophys.
J. 78, 2761–2775 (2000)
412. Qu, Z., Xie, F., Garfinkel, A., Weiss, J.N.: Origins of spiral wave meander and breakup in a
two-dimensional cardiac tissue model. Ann. Biomed. Eng. 28, 755–771 (2000)
413. Quan, W., Evans, S.J., Hastings, H.M.: Efficient integration of a realistic two-dimensional
cardiac tissue model by domain decomposition. IEEE Trans. Biomed. Eng. 45, 372–385
(1998)
414. Quarteroni, A., Valli, A.: Numerical Approximation of Partial Differential Equations.
Springer, Berlin (1994)
415. Ramanathan, C., Ghanem, R.N., Jia, P., Ryu, K., Rudy, Y.: Noninvasive electrocardiographic
imaging for cardiac electrophysiology and arrhythmia. Nat. Med. 10(4), 422–428 (2004)
416. Ranjan, R., Tomaselli, G.F., Marban, E.: A novel mechanism of anode-break stimulation
predicted by bidomain modeling. Circ. Res. 84, 153–156 (1999)
417. Rappel, W.J.: Filament instability and rotational tissue anisotropy: a numerical study using
detailed cardiac models. Chaos 11(1), 71–80 (2001)
418. Rasmusson, R.I., Clark, J.W., Giles, W.R., Robinson, K., Clark, R.B., Shibata, E.F., Campbell,
D.L.: A mathematical model of electrophysiological activity in a bullfrog atrial cell. Am. J.
Physiol. 259, H370–H389 (1990)
419. Rinzel, J., Ermentrout, B.: Analysis of neural excitability and oscillations. In: Kock, C., et al.
(eds.) Methods in Neuronal Modelling: From Synapses to Networks. MIT, Boston (1998)
420. Roberts, D., Scher, A.M.: Effect of tissue anisotropy on extracellular potential fields in canine
myocardium in situ. Circ. Res. 50, 342–351 (1982)
421. Roberts, D., Hersch, L.T., Scher, A.M.: Influence of cardiac fiber orientation on wave front
voltage, conduction velocity and tissue resistivity. Circ. Res. 44, 701–712 (1979)
422. Rogers, J.M., McCulloch, A.D.: A collocation-Galerkin finite element model of cardiac action
potential propagation. IEEE Trans. Biomed. Eng. 41, 743–757 (1994)
423. Rohr, S.: Role of gap junctions in the propagation of the cardiac action potential. Cardiovasc.
Res. 62, 309–322 (2004)
424. Rohr, S.: Cardiac fibroblasts in cell culture systems: myofibroblasts all along? J. Cardiovasc.
Pharmacol. 57(4), 389–399 (2011)
425. Romero, D., Sebastian, R., Bijnens, B.H., et al.: Effects of the Purkinje system and cardiac
geometry on biventricular pacing: a model study. Ann. Biomed. Eng. 38(4), 1388–1398
(2010)
386 References
426. Roth, B.J.: The electrical potential produced by a strand of cardiac muscle: a bidomain
analysis. Ann. Biomed. Eng. 16, 609–637 (1988)
427. Roth, B.J.: A mathematical model of make and break electrical stimulation of cardiac tissue
by a unipolar anode or cathode. IEEE Trans. Biomed. Eng. 42, 1174–1184 (1995)
428. Roth, B.J.: Strength-Interval curve for cardiac tissue predicted using the bidomain model. J.
Cardiovasc. Electrophysiol. 7, 722–737 (1996)
429. Roth, B.J.: Nonsustained reentry following successive stimulation of cardiac tissue through a
unipolar electrode. J. Cardiovasc. Electrophysiol. 8, 768–778 (1997)
430. Roth, B.J., Chen, J.: Mechanism of anode break excitation in the heart: the relative influence
of membrane and electrotonic factors. J. Biol. Syst. 7(4), 541–552 (1999)
431. Roth, B.J., Krassowska, W.: The induction of reentry in cardiac tissue. The missing link: how
electric fields alter transmembrane potential. Chaos 8, 204–220 (1998)
432. Roth, B.J., Patel, S.G.: Effects of elevated extracellular potassium ion concentration on anodal
excitation of cardiac tissue. J. Cardiovasc. Electrophysiol. 14, 1351–1355 (2003)
433. Roth, B.J., Pertsov, A.M.: Hybrid modeling of electrical and optical behavior in the heart.
Physica D 238, 1019–1027 (2009)
434. Roth, B.J., Wikswo, J.P., Jr.: A bidomain model for the extracellular potential and magnetic
field of cardiac tissue. IEEE Trans. Biomed. Eng. 33(4), 467–469 (1986)
435. Roth, B.J., Wikswo, J.P., Jr.: Electrical stimulation of cardiac tissue: a bidomain model with
active membrane properties. IEEE Trans. Biomed. Eng. 41(3), 232–240 (1994)
436. Roth, B.J., Lin, S.-F., Wikswo, J.P., Jr.: Unipolar stimulation of cardiac tissue. J. Electrocar-
diol. 31(Suppl), 6–12 (1998)
437. Roux, B., Allen, T., Berneche, S.: Theoretical and computational models of biological ion
channels. Q. Rev. Biophys. 37(1), 15–103 (2001)
438. Rubinstein, I.: Electro-Diffusion of Ions. SIAM, Philadelphia (1990)
439. Rudy, Y.: The electrocardiogram and its relationship to excitation of the heart. In: Sperelakis,
N. (ed.) Physiology and Pathophysiology of the Heart, 3rd edn., chap. 11, pp. 201–239.
Kluwer Academic, Dordrecht (1995)
440. Rudy, Y.: Noninvasive electrocardiographic imaging of arrhythmogenic substrates in humans.
Circ. Res. 112, 863–874 (2013)
441. Rudy, Y., Messinger-Rapport, B.J.: The inverse problem in electrocardiography: solutions in
terms of epicardial potentials. CRC Crit. Rev. Biomed. Eng. 16(3), 215–268 (1988)
442. Rudy, Y., Oster, H.S.: The electrocardiographic inverse problem. CRC Crit. Rev. Biomed.
Eng. 20, 25–45 (1992)
443. Rudy, Y., Silva, J.R.: Computational biology in the study of cardiac ion channels and cell
electrophysiology. Q. Rev. Biophys. 39(1), 57–116 (2006)
444. Rushmer, R.F.: Structure and Function of the Cardiovascular System, 2nd edn.
W. B. Saunders, Philadelphia (1976)
445. Saad, Y.: Iterative Methods for Sparse Linear Systems, 2nd edn. SIAM, Philadelphia (2003)
446. Sachse, F.B.: Computational Cardiology. Modeling of Anatomy, Electrophysiology, and
Mechanics. LNCS, vol. 2966. Springer, Berlin (2004)
447. Sachse, F.B., Seemann, G. (eds.): Proceedings of the 4th International Conference on
Functional Imaging and Modeling of the Heart, FIMH’07, Salt Lake City, 7–9 June 2007.
LNCS, vol. 4466. Springer, Berlin (2007)
448. Saffitz, J.E., Kanter, H.L., Green, K.G., Tolley, T.K., Beyer, E.C.: Tissue-specific determinants
of anisotropic conduction velocity in canine atrial and ventricular myocardium. Circ. Res. 74,
1065–1070 (1994)
449. Saleheen, H.I., Ng, K.T.: A new three-dimensional finite-difference bidomain formulation for
inhomogeneous anisotropic cardiac tissues. IEEE Trans. Biomed. Eng. 45(1), 15–25 (1998)
450. Sambelashvili, A., Efimov, I.R.: Dynamics of virtual electrode-induced scroll-wave reentry in
a 3D bidomain model. Am. J. Physiol Heart Circ. Physiol. 287, H1570–H1581 (2004)
451. Sambelashvili, A., Nikolsky, V.P., Efimov, I.R.: Virtual electrode theory explains pacing
threshold increase caused by cardiac tissue damage. Am. J. Physiol Heart Circ. Physiol. 286,
H2183–H2194 (2004)
References 387
452. Sampson, K.J., Henriquez, C.S.: Electrotonic influences on action potential duration disper-
sion in small hearts: a simulation study. Am. J. Physiol. Heart Circ. Physiol. 289, H350–H360
(2005)
453. Samson, W., Scher, A.: Mechanism of ST-segment alteration during acute myocardial injury.
Circ. Res. 8, 780–787 (1960)
454. Sanchez-Palencia, E., Zaoui, A.: Homogenization Techniques for Composite Media. Lectures
Notes in Physics, vol. 272. Springer, Berlin (1987)
455. Sanfelici, S.: Convergence of the Galerkin approximation of a degenerate evolution problem
in electrocardiology. Numer. Methods Part. Differ. Equ. 18, 218–240 (2002)
456. Sanfelici, S.: Numerical simulations of fractioned electrograms and pathological cardiac
action potential. J. Theor. Med. 4(3), 167–181 (2002)
457. Savaré, G.: Weak solutions and maximal regularity for abstract evolution inequalities. Adv.
Math. Sci. Appl. 6, 377–418 (1996)
458. Scacchi, S.: A hybrid multilevel Schwarz method for the bidomain model. Comput. Methods
Appl. Mech. Eng. 197(45–48), 4051–4061 (2008)
459. Scacchi, S.: A multilevel hybrid Newton-Krylov-Schwarz method for the Bidomain model of
electrocardiology. Comput. Methods Appl. Mech. Eng. 200(5–8), 717–725 (2011)
460. Scacchi, S., Colli Franzone, P., Pavarino, L.F., Taccardi, B.: A reliability analysis of cardiac
repolarization time markers. Math. Biosci. 219(2), 113–128 (2009)
461. Scacchi, S., Colli Franzone, P., Pavarino, L.F., Taccardi, B.: Computing cardiac recovery
maps from electrograms and monophasic action potentials under heterogeneous and ischemic
conditions. Math. Model Methods Appl. Sci. 20(7), 1089–1127 (2010)
462. Scher, A.M.: Excitation of the heart. In: Nelson, C.V., Geselowitz, D.B. (eds.) The Theoretical
Basis of Electrocardiology, pp. 44–67. Clarendon, Oxford (1976)
463. Schmitt, O.H.: Biological information processing using the concept of interpenetrating
domains. In: Leibovich, K.N. (ed.) Information Processing in the Nervous System, pp. 325–
331. Springer, New York (1969)
464. Schuss, Z., Nadler, B., Eisenberg, R.S.: Derivation of Poisson and Nernst-Planck equations in
a bath and channel from a molecular model. Phys. Rev. E 64, 036116 (2001)
465. Scollan, D.F., Holmes, A., Zhang, J., Winslow, R.L.: Reconstruction of cardiac ventricular
geometry and fiber orientation using magnetic resonance imaging. Ann. Biomed. Eng. 28(8),
934–944 (2000)
466. Seemann, G., Hoeper, C., Sachse, F.B., Doessel, O., Holden, A.V.: Heterogeneous three-
dimensional anatomical and electrophysiological model of human atria. Philos. Trans. R. Soc.
A 364, 1465–1481 (2006)
467. Seemann, G., Sachse, F.B., Weiss, D.L., Dossel, O.: Quantitative reconstruction of cardiac
electromechanics in human myocardium: regional heterogeneity. J. Cardiovasc. Electrophys-
iol. 14(10), S219–S228 (2003)
468. Seger, M., Fischer, G., Modre, R., Messnarz, B., Hanser, F., Tilg, B.: Lead field computation
for the electrocardiographic inverse problem – finite element versus boundary elements.
Comput. Methods Prog. Biomed. 77, 241–252 (2005)
469. Sepulveda, N.G., Roth, B.J., Wikswo, J.P., Jr.: Current injection into a two-dimensional
anisotropic bidomain. Biophys. J. 55, 987–999 (1989)
470. Sermesant, M., Konukoglu, E., Delingette, H., Coudiere, Y., Khinchapatnam, P., Rhode,
K.S., Razzavi, R., Ayache, N.: An anisotropic multi-front fast marching method for real-time
simulation in cardiac electrophysiology. In: Sachse, F.B., Seemann, G. (eds.) FIMH’07, Salt
Lake City, 7–9 June 2007. LNCS, vol. 4466, pp. 160–169. Springer, Berlin (2007)
471. Sethian, J.A.: Level Set Methods: Evolving Interfaces in Geometry, Fluid Mechanics,
Computer Vision and Material Sciences, 2nd edn. Cambridge University Press, Cambridge
(1999)
472. Severi, S., Fantini, M., Charawi, L.A., Di Francesco, D.: An updated computational model
of rabbit sinoatrial action potential to investigate the mechanisms of heart rate modulation.
J. Physiol. 590, 4483–4499 (2012)
388 References
473. Shannon, T.R., Wang, F., Puglisi, J., Weber, C., Bers, D.M.: A mathematical treatment of
integrated Ca dynamics within the ventricular myocyte. Biophys. J. 87(5), 3351–3357 (2004);
Erratum Biophys. J. 102(8), 1996–2001 (2012)
474. Shaw, R.M., Rudy, Y.: Ionic mechanisms of propagation in cardiac tissue. Role of the sodium
and L-type calcium currents during reduced excitability and decreased gap junction coupling.
Circ. Res. 81(5), 727–741 (1997)
475. Shenasa, M., Hindricks, G., Borggrefe, M., Breithardt, G., Josephson, M.E.: Cardiac Map-
ping, 4th edn. Wiley-Blackwell, Chichester (2012)
476. Shibata, N., Chen, P.S., Dixon, E.G., Wolf, P.D., Danieley, N.D., Smith, W.M., Ideker, R.E.:
Influence of shock strength and timing on induction of ventricular arrhythmias in dogs. Am.
J. Physiol. 255, H891–H901 (1988)
477. Shou, G., Xia, L., Jiang, M.: Total variation regularization in electrocardiographic mapping.
In: Li, K., et al. (eds.) Life System Modeling and Intelligence Computing. LNMI, vol. 6330,
pp. 51–59. Springer, Berlin (2010)
478. Skipa, O., Nalbach, M., Sachse, F., Werner, C., Dossel, O.: Transmembrane potential
reconstruction in anisotropic heart model. Int. J. Bioelectromagn. 4(2), 17–18 (2002)
479. Sicouri, S., Antzelevich, C.: A subpopulation of cells with unique electrophysiological
properties in the deep subepicardium of the canine ventricle. The M cell. Circ. Res. 68, 1729–
1741 (1991)
480. Sidorov, V.Y., Woods, M.C., Baudenbacher, P., Baudenbacher, F.: Examination of stimulation
mechanism and strength-interval curve in cardiac tissue. Am. J. Physiol. Heart Circ. Physiol.
289, H2602–H2615 (2005)
481. Sigg, D.C., Iaizzo, P.A., Xiao, Y.-F., He, B.: Cardiac Electrophysiology Methods and Models.
Springer, New York (2010)
482. Silva, J.R., Pan, H., Wu, D., Nekouzadeh, A., Decker, K.F., Cui, J., Baker, N.A., Sept, D.,
Rudy, Y.: A multiscale model linking ion-channel molecular dynamics and electrostatics to
the cardiac action potential. Proc. Nat. Acad. Sci. 106(27), 11102–11106 (2009)
483. Simms, H.D., Geselowitz, D.B.: Computation of heart surface potentials using the surface
source model. J. Cardiovasc. Electrophysiol. 6, 522–531 (1995)
484. Singer, A., Gillespie, D., Norbury, J., Eisenberg, R.S.: Singular perturbation analysis of the
steady-state Poisson-Nernst-Planck system: applications to ion channels. Eur. J. Appl. Math.
19, 541–560 (2008)
485. Skouibine, K., Krassowska, W.: Increasing the computational efficiency of a bidomain model
of defibrillation using a time-dependent activating function. Ann. Biomed. Eng. 28, 772–780
(2000)
486. Skouibine, K., Trayanova, N., Moore, P.: Anode/cathode make and break phenomena in a
model of defibrillation. IEEE Trans. Biomed. Eng. 46(7), 769–777 (1999)
487. Skouibine, K., Trayanova, N., Moore, P.: A numerically efficient model for the simulation
of defibrillation in an active bidomain sheet of myocardium. Math. Biosci. 166(1), 85–100
(2000)
488. Smaill, B.H., Zhao, J., Trew, M.L.: Three-dimensional impulse propagation in myocardium.
Arrhythmogenic mechanisms at the tissue level. Circ. Res. 112, 834–848 (2013)
489. Smith, B.F., Bjørstad, P., Gropp, W.D.: Domain Decomposition: Parallel Multilevel Methods
for Elliptic Partial Differential Equations. Cambridge University Press, Cambridge (1996)
490. Smith, N.P., Pullan, A.J., Hunter, P.J.: An anatomically based model of transient coronary
blood flow in the heart. SIAM J. Appl. Math. 62(3), 990–1018 (2002)
491. Smith, N.P., Nickerson, D.P., Crampin, E.J., Hunter, P.J.: Multiscale computational modelling
of the heart. Acta Numer. 371–431 (2004)
492. Soravia, J.P., Souganidis, P.E.: Phase-field theory for Fitzhugh-Nagumo type systems. SIAM
J. Math. Anal. 27(5), 1341–1359 (1996)
493. Spach, M.S., Dolber, P.C.: Relating extracellular potentials and their derivatives to anisotropic
propagation at microscopic level in human cardiac muscle. Evidence for electrical uncoupling
of side-to-side fiber connections with increasing age. Circ. Res. 58, 356–371 (1986)
References 389
494. Spiteri, R.J., Dean, R.C.: On the performance of an implicit-explicit Runge-Kutta method in
models of cardiac electrical activity. IEEE Trans. Biomed. Eng. 55(5), 1488–1495 (2008)
495. Stampacchia, G.: Boundary value problems for some degenerate-elliptic operators. Ann.
Math. Pura Appl. LXXX(IV), 1–122 (1968)
496. Steinhaus, B.M.: Estimating cardiac transmembrane activation and recovery times from
unipolar and bipolar extracellular electrograms: a simulation study. Circ. Res. 64(3), 449–
462 (1989)
497. Steinhaus, B.M., Spitzer, K.W., Isomura, S.: Action potential collision in heart tissue.
Computer simulations and tissue experiments. IEEE Trans. Biomed. Eng. 32(10), 731–742
(1985)
498. Streeter, D.: Gross morphology and fiber geometry in the heart. In: Berne, R.M. (ed.)
Handbook of Physiology, vol. 1, sect. 2, pp. 61–112. Williams & Wilkins, Philadelphia (1979)
499. Sundnes, J., Lines, G.T., Mardal, K., Tveito, A.: Multigrid block preconditioning for a coupled
system of partial differential equations modeling the electrical activity in the heart. Comput.
Methods Biomech. Biomed. Eng. 5, 397–409 (2002)
500. Sundnes, J., Lines, G., Lines, G.T., Grottum, P., Tveito, A.: Electrical activity in the human
heart. In: Langtangen, H.P., Tveito, A. (eds.) Advanced Topics in Computational Partial
Differential Equations. LNCSE, vol. 33, chap. 10, pp. 401–449. Springer, Berlin (2004)
501. Sundnes, J., Lines, G.T., Tveito, A.: An operator splitting method for solving the bidomain
equations coupled to a volume conductor model for the torso. Math. Biosci. 194(2), 233–248
(2005)
502. Sundnes, J., Lines, G.T., Cai, X., Nielsen, B.F., Mardal, K.-A., Tveito, A.: Computing the
Electrical Activity of the Heart. Springer, Berlin (2006)
503. Sundnes, J., Nielsen, B.F., Mardal, K.A., Lines, G.T., Mardal, K.A., Tveito, A.: On the
computational complexity of the bidomain and the monodomain models of electrophysiology.
Ann. Biomed. Eng. 34(7), 1088–1097 (2006)
504. Taccardi, B., Punske, B.B.: Body surface potential mapping. In: Zipes, D., Jalife, J. (eds.)
Cardiac Electrophysiology. From cell to Bedside, 4th edn., pp. 803–811. W. B. Saunders,
Philadelphia (2004)
505. Taccardi, B., de Ambroggi, L., Viganotti, C.: Body surface mapping of heart potentials. In:
Nelson, C.V., Geselowitz, D.B. (eds.) The Theoretical Basis of Electrocardiology, pp. 436–
466. Clarendon, Oxford (1976)
506. Taccardi, B., Lux, R.L., Ershler, P.R., MacLeod, R.S., Vyhmeister, Y.: Effect of myocardial
fiber direction on 3-D shape of excitation wavefronts and associated potential distributions in
ventricular walls. Circulation 86, I-752 (1992)
507. Taccardi, B., Macchi, E., Lux, R.L., Ershler, P.R., Spaggiari, S., Baruffi, S., Vyhmeister,
Y.: Effect of myocardial fiber direction on epicardial potentials. Circulation 90, 3076–3090
(1994)
508. Taccardi, B., Lux, R.L., Ershler, P.R., MacLeod, R.S., Vyhmeister, Y.: Modern views on the
spread of excitation in anisotropic heart muscle. Jpn. Heart J. 35, 31–35 (1994)
509. Taccardi, B., Lux, R.L., Ershler, P.R., MacLeod, R.S., Dustman, T.J., Ingebrigtsen, N.:
Anatomical architecture and electrical activity of the heart. Acta Cardiol. 52, 91–105 (1997)
510. Taccardi, B., Veronese, S., Colli Franzone, P., Guerri, L.: Multiple components in the
unipolar electrocardiogram: a simulation study in a three-dimensional model of ventricular
myocardium. J. Cardiovasc. Electrophysiol. 9, 1062–1084 (1998)
511. Taccardi, B., Punske, B., Lux, R., MacLeod, R., Ershler, P., Dustman, T., Vyhmeister, Y.:
Useful lesson from body surface mapping on body. J. Cardiovasc. Electrophysiol. 9(7), 773–
786 (1998)
512. Taccardi, B., Punske, B., Helie, F., MacLeod, R., Lux, R., Ershler, P., Dustman, T.,
Vyhmeister, Y.: Epicardial recovery sequences and excitation recovery intervals during paced
beats. Role of myocardial architecture. PACE 22(4), part II: 833 (1999)
513. Taccardi, B., Punske, B.B., MacLeod, R.S., Ni, Q.: Visualization, analysis and physiological
interpretation of three-dimensional cardiac electric fields. In: Proceedings of the 2nd Joint
EMBS/BMSE Conference, Houston, Oct 2002. vol. 2, pp. 1366–1367 (2002)
390 References
514. Taccardi, B., Punske, B.B., Colli Franzone, P.: Cardiac potential mapping. In: Proceedings
of EMBS/25th IEEE Annual International Conference, Cancun, 17–21 Sept 2003. vol. 4,
pp. 3749–3752 (2003)
515. Taccardi, B., Punske, B.B., Sachse, F., Tricoche, X., Colli Franzone, P., Pavarino, L.F.,
Zabawa, C.: Intramural activation and repolarization sequences in canine ventricles. Experi-
mental and simulation studies. J. Electrocardiol. 38, 131–137 (2005)
516. Taccardi, B., Punske, B., Macchi, E., MacLeod, R.S., Ershler, P.R.: Epicardial and intramural
excitation during ventricular pacing: effects of myocardial structure. Am. J. Physiol. Heart.
Circ. Physiol. 294, H1753–H1766 (2008)
517. Taggart, P., Sutton, P., Opthof, T., Coronel, R., Trimlett, R., Pugsley, W., Kallis, P.: Transmural
repolarization in the left ventricle in humans during normoxia and ischemia. Cardiovasc. Res.
50, 454–462 (2001)
518. Tarkhanov, N.N.: The Cauchy Problem for Solutions of Elliptic Equations. Akademic Verlag
Gmbh, Berlin (1995)
519. ten Tusscher, K.H.W.J., Panfilov, A.V.: Alternans and spiral breakup in a human ventricular
tissue model. Am. J. Physiol. Heart Circ. Physiol. 291(3), H1088–H1100 (2006)
520. ten Tusscher, K.H.W.J., Panfilov, A.V.: Modelling of the ventricular conduction system. Prog.
Biophys. Mol. Biol. 96(1–3), 152–170 (2008)
521. ten Tusscher, K., Noble, D., Noble, P.J., Panfilov, A.V.: A model for human ventricular tissue.
Am. J. Phys. Heart. Circ. Physiol. 286(4), H1573–H1589 (2004)
522. Tomlinson, K.A., Hunter, P.J., Pullan, A.J.: A finite element method for an eikonal equation
model of myocardial excitation wavefront propagation. SIAM J. Appl. Math. 63(1), 324–350
(2002)
523. Toselli, A., Widlund, O.B.: Domain Decomposition Methods: Algorithms and Theory.
Computational Mathematics, vol. 34. Springer, Berlin (2004)
524. Transgenstein, J.A., Kim, C.: Operator splitting and adaptive mesh refinement for the Luo–
Rudy I model. J. Comput. Phys. 196, 645–679 (2004)
525. Trayanova, N.A.: Defibrillation of the heart: insights into mechanisms from modelling
studies. Exp. Physiol. 91(2), 323–337 (2006)
526. Trayanova, N., Eason, J., Aguel, F.: Computer simulations of cardiac defibrillation: a look
inside the heart. Comput. Vis. Sci. 4, 259–270 (2002)
527. Trayanova, N.A., Constantino, J., Gurev, V.: Electromechanical models of the ventricles. Am.
J. Physiol. Heart Circ. Physiol. 301(2), H279–H286 (2011)
528. Trew, M., Le Grice, I., Smaill, B., Pullan, A.: A finite volume method for modeling
discontinuous electrical activation in cardiac tissue. Ann. Biomed. Eng. 33(5), 590–602
(2005)
529. Trew, M., Smaill, B., Bullivant, D., Hunter, P., Pullan, A.: A generalized finite difference
method for modeling cardiac electrical activation on arbitrary, irregular computational
meshes. Math. Biosci. 198(2), 169–189 (2005)
530. Trew, M.L., Caldwell, B.J., Sands, G.B., Hooks, D.A., Tai, D.C.-S., Austin, T.M., LeGrice,
I.J., Pullan, A.J., Smaill, B.H.: Cardiac electrophysiology and tissue structure: bridging the
scale gap with a joint measurement and modelling paradigm. Exp. Physiol. 91(2), 355–370
(2006)
531. Tung, L.: A bidomain model for describing ischemic myocardial D.C. potentials. Ph.D.
dissertation, MIT, Cambridge, MA (1978)
532. Tyson, J.J., Keener, J.P.: Singular perturbation theory of traveling waves in excitable media.
(A review). Physica D 32, 327–361 (1988)
533. Ueda, N., Zipes, D.P., Wu, J.: Functional and transmural modulation of M cell behavior in
canine ventricular wall. Am. J. Physiol. (Heart Circ. Physiol) 287, H2569–H2575 (2004)
534. van Dam, P.M., Oostendorp, T.F., Linnenbank, A.C., van Oosterom, A.: Non-invasive imaging
of cardiac activation and recovery. Ann. Biomed. Eng. 37(9), 1739–1756 (2009)
535. van Oosterom, A.: Cell models–Macroscopic source descriptions. In: Macfarlane, P.W.,
Lawrie, T.D.V. (eds.) Comprehensive Electrocardiology, pp. 155–179. Pergamon, Oxford
(1989)
References 391
536. van Oosterom, A.: Forward and inverse problems in electrocardiography. In: Panfilov, A.V.,
Holden, A.V. (eds.) Computational Biology of the Heart. Wiley, New York (1997)
537. van Oosterom, A.: Genesis of the T wave as based on an equivalent surface source model. J.
Electrocardiol. 34(Suppl), 217–227 (2001)
538. van Oosterom, P.: Genesis of the T wave as based on an equivalent surface source model. J.
Electrocardiol. 34, 217–227 (2001)
539. Varghese, A., Sell, G.R.: A conservation principle and its effect on the formulation of Na–Ca
exchanger current in cardiac cells. J. Theor. Biol. 189, 33–40 (1997)
540. Veneroni, M.: Reaction-diffusion systems for the microscopic cellular model of the cardiac
electric field. Math. Methods Appl. Sci. 29, 1631–1661 (2006)
541. Veneroni, M.: Reaction-Diffusion systems for the macroscopic Bidomain model of the cardiac
electric field. Nonlinear Anal. Real World Appl. 10(2), 849–868 (2009)
542. Vigmond, E.J.: The electrophysiologic basis of MAP recordings. Cardiovasc. Res. 68, 502–
503 (2005)
543. Vigmond, E.J., Leon, L.J.: Computational efficient model for simulating electrical activity in
cardiac tissue with fiber rotation. Ann. Biomed. Eng. 27, 160–170 (1999)
544. Vigmond, E.J., Aguel, F., Trayanova, N.A.: Computational techniques for solving the
bidomain equations in three dimensions. IEEE Trans. Biomed. Eng. 49(11), 1260–1269
(2002)
545. Vigmond, E.J., Weber dos Santos, R., Prassl, A.J., Deo, M., Plank, G.: Solvers for the cardiac
bidomain equations. Prog. Biophys. Mol. Biol. 96, 3–18 (2008)
546. Viswanathan, P.C., Shaw, R.M., Rudy, Y.: Effects of IKr and IKs heterogeneity on action
potential duration and its rate dependence. A simulation study. Circulation 99(18), 2466–
2474 (1999)
547. Wagner, J., Keizer, J.: Effects of rapid buffers on Ca2C diffusion and CaC2 oscillations.
Biophys. J. 67, 447–456 (1994)
548. Wang, Y., Rudy, Y.: Action potential propagation in inhomogeneous cardiac tissue: safety
factor considerations and ionic mechanism. Am. J. Physiol. Heart Circ. Physiol. 278, H1019–
H1029 (2000)
549. Wang, D., Kirby, R.M., Johnson, C.R.: Finite-element-based discretization and regularization
strategies for 3-D inverse electrocardiography. IEEE Trans. Biomed. Eng. 58(6), 1827–1838
(2011)
550. Wang, L., Wong, K.C., Zhang, H., Liu, H., Shi, P.: Noninvasive computational imaging
of cardiac electrophysiology for 3-D infarct. IEEE Trans. Biomed. Eng. 58(4), 1033–1043
(2011)
551. Wang, L., Dawoud, F., Yeung, S.-K., Shi, P., Wong, K., Lardo, A.: Transmural imaging of
ventricular action potential and post-infarction scars in swine hearts. IEEE Trans. Med. Image
32(4), 731–747 (2013)
552. Weber dos Santos, R., Plank, G., Bauer, S., Vigmond, E.J.: Parallel multigrid preconditioner
for the cardiac bidomain model. IEEE Trans. Biomed. Eng. 51(11), 1960–1968 (2004)
553. Weissenburger, J., Nesterenko, V., Antzelevitch, C.: Transmural heterogeneity of ventricular
repolarization under baseline and long QT conditions in the canine heart in vivo: torsade de
Pointes develops with halotane but not pentobarbital anesthesia. J. Cardiovasc. Electrophys-
iol. 11, 290–304 (2000)
554. White, C.S., Haramati, L.B., Jen-Sho Chen, J., Levsky, J.M.: Cardiac Imaging. Oxford
University Press, Oxford (2014)
555. Whiteley, J.P.: An efficient numerical technique for the solution of the monodomain and
bidomain equations. IEEE Trans. Biomed. Eng. 53(11), 2139–2147 (2006)
556. Whiteley, J.P.: Physiology driven adaptivity for the numerical solution of the bidomain
equations. Ann. Biomed. Eng. 35(9), 1510–1520 (2007)
557. Wikswo, J.P., Jr.: Tissue anisotropy, the cardiac bidomain, and the virtual cathode effect.
In: Zipes D.P., Jalife J. (eds.) Cardiac Electrophysiology: From Cell to Bedside, 2nd edn.,
pp. 348–361. W. B. Saunders, Philadelphia (1994)
392 References
558. Wikswo, J.P., Jr., Roth, B.J.: Virtual electrode theory of pacing. In: Efimov, I.R., Kroll, M.W.,
Tchou, P.J. (eds.) Cardiac Bioelectric Therapy, Chap. 4.3, pp. 283–330. Springer, New York
(2009)
559. Wikswo, J.P., Jr., Lin, S.-F., Abbas, R.A.: Virtual electrodes in cardiac tissue: a common
mechanism for anodal and cathodal stimulation. Biophys. J. 69, 2195–2210 (1995)
560. Wilson, L.D., Jeyaraj, D.: Controversies in measuring repolarization using extracellular
recordings: why should be care. Heart Rhythm 3(9), 1051–1052 (2006)
561. Winfree, A.T.: Sudden cardiac death: a problem in topology. Sci. Am. 248, 144–161 (1983)
562. Winfree, A.T.: The Geometry of Biological Time, 2nd edn. Springer, New York (2001)
563. Winslow, R.L., Rice, J., Jafri, S., Marban, E., O’Rourke, B.: Mechanisms of altered excitation-
contraction coupling in canine tachycardia-induced heart failure, II – model studies. Circ. Res.
84(5), 571–586 (1999)
564. Wit, A.L., Janse, M.J.: The Ventricular Arrhythmias of Ischemia and Infarction: Electrophys-
iological Mechanisms. Futura Publishing Co., New York (1993)
565. Wolferth, C.C., Bettet, S., Livezey, M.M., Murphy, F.: Negative displacement of the RS-T seg-
ment in the electrocardiogram and its relationships to positive displacement: an experimental
study. Am. Heart J. 29, 220–244 (1945)
566. Wyatt, R.P.: Comparison of estimates of activation and recovery times from bipolar and
unipolar electrograms to in vivo transmembrane action potential durations. In: Proceedings
of IEEE Engineering in Medicine and Biology Society, 2nd Annual Conference, Washington,
DC, pp. 22–25 (1980)
567. Xia, Y., Kongstad, O., Hertvig, E., Li, Z., Holm, M., Olsson, B., Yuan, S.: Activation
recovery time measurements in evaluation of global sequence and dispersion of ventricular
repolarization. J. Electrocardiol. 38, 28–35 (2005)
568. Xie, F., Qu, Z.L., Yang, J., Baher, A., Weiss, J.N., Garfinkel, A.: A simulation study of the
effects of cardiac anatomy in ventricular fibrillation. J. Clin. Invest. 113, 686–693 (2004)
569. Xu, A., Guevara, M.R.: Two forms of spiral-wave reentry in an ionic model of ischemic
ventricular myocardium. Chaos 8(1), 157–174 (1998)
570. Yamashita, Y., Geselowitz, D.B.: Source–field relationships for cardiac generators on the heart
surface based on their transfer coefficients. IEEE Trans. Biomed. Eng. 32, 964–970 (1985)
571. Yan, G.-X., Antzelevitch, C.: Cellular basis for the normal T wave and the electrocardio-
graphic manifestations of the Long-QT syndrome. Circulation 98, 1928–1936 (1998)
572. Yan, G.X., Shimizu, W., Antzelevitch, C.: Characteristics and distribution of M cells in
arterially perfused canine left ventricular wedge preparations. Circulation 98(18), 1921–1927
(1998)
573. Yehia, A.R., Jeandupeaux, D., Alonso, F., Guevara, M.R.: Hysteresis and bistability in the
direct transition from 1:1 to 2:1 rhythm in periodically driven single ventricular cells. Chaos
9, 916–931 (1999)
574. Ying, W.J., Rose, D.J., Henriquez, C.S.: Efficient fully implicit time integration methods for
modeling cardiac dynamics. IEEE Trans. Biomed. Eng. 55(12), 2701–2711 (2008)
575. Young, R.J., Panfilov, A.: Anisotropy of wave propagation in the heart can be modeled by a
Riemannian electrophysiological metric. Proc. Natl. Acad. Sci. 107(34), 15063–15068 (2010)
576. Young, A.A., LeGrice, I.J., Young, M.A., Smaill, B.H.: Extended confocal microscopy of
myocardial laminae and collagen network. J. Microscop. 192, 139–150 (1998)
577. Yu, H.: Solving parabolic problems with different time steps in different regions in space
based on domain decomposition methods. Appl. Numer. Math. 30(4), 475–491 (1999)
578. Yu, H.: A local space-time adaptive scheme in solving two-dimensional parabolic problems
based on domain decomposition methods. SIAM J. Sci. Comput. 23(1), 304–322 (2001)
579. Yu, H., Chang, F., Cohen, I.S.: Pacemaker i(f) in adult canine cardiac ventricular myocytes. J.
Physiol. 485, 469–483 (1995)
580. Yue, A.M., Betts, T.R., Roberts, P.R., Morgan, J.M.: Global dynamic coupling of activation
and repolarization in human ventricle. Circulation 112, 2592–2601 (2005)
581. Zampini, S.: Balancing Neumann-Neumann methods for the cardiac Bidomain model. Numer.
Math. 123, 363–393 (2013)
References 393
582. Zampini, S.: Dual-primal methods for the cardiac bidomain model. Math. Model Methods
Appl. Sci. 24(4), 667–696 (2014)
583. Zeng, J., Laurita, K.R., Rosenbaum, D.S., Rudy, Y.: Two Components of the delayed rectifier
K C current in ventricular myocytes of the guinea pig type. Theoretical formulation and their
role in repolarization. Circ. Res. 77, 140–152 (1995)
584. Zhang, X.: Multilevel Schwarz methods. Numer. Math. 63(4), 521–539 (1992)
585. Zhang, H., Holden, A.V., Kodama, I., Honjo, H., Lei, M., Varghese, T., Boyett, M.R.:
Mathematical models of action potentials in the periphery and center of the rabbit sinoatrial
node. Am. J. Physiol. Heart. Circ. Physiol. 279, H397–H421 (2000)
586. Zipes, D., Jalife, J.: Cardiac Electrophysiology, 2nd edn. W. B. Saunders, Philadelphia (1995)
587. Zipes, D., Jalife, J.: Cardiac Electrophysiology, 3rd edn. W. B. Saunders, Philadelphia (2000)
588. Zipes, D., Jalife, J.: Cardiac Electrophysiology: From Cell to Bedside, 4th edn.
W. B. Saunders, Philadelphia (2004)
589. Zipes, D., Jalife, J.: Cardiac Electrophysiology, 5th edn. W. B. Saunders, Philadelphia (2009)
590. Zipes, D., Jalife, J.: Cardiac Electrophysiology: From Cell to Bedside, 6th edn. W. B.
Saunders, Philadelphia (2013)
591. Zozor, S., Blanc, O., Jacquemet, V., Virag, N., Vesin, J., Pruvot, E., Kappenberger, L.,
Henriquez, C.: A numerical scheme for modeling wavefront propagation on a monolayer
of arbitrary geometry. IEEE Trans. Biomed. Eng. 50(4), 412–420 (2003)
Index
Fiber architecture, 7
Oblique dipole layer, 168
Fibroblasts, 5
Optical mapping, 19
FitzHugh-Nagumo (FHN) model, 60
phase-plane analysis, 62
Frequency diagram, 71 P wave, 13
FHN, 72 Parallel Bidomain solvers, 227
Poisson-Nernst-Planck (PNP) equation, 29
Precordial leads, 13
Gap junctions, 6 Purkinje fibers, 4
Gating variables, 36
Goldman-Hodgkin-Katz equation, 24
QRS complex, 14
QT interval, 15
Heart surface source model, 157
Hodgkin-Huxley (HH) model, 41
Homogenization, 98 Reentry phenomena, 348
Hybrid monophasic action potential (HMAP), anatomical reentry, 348
325 circus movement, 348
Hybrid Schwarz preconditioners, 234 critical point theory, 352
cross gradient stimulation, 353
figure-8 reentry, 358
Infarction, 17 functional reentry, 348
Inverse problem of Electrocardiology, 175 pinwheel experiment, 348
in terms of potential, 176 rotors, 351
in terms of wavefront, 187 scroll wave, 352
Ionic channels, 21 scroll wave breakup, 352, 353
Ionic currents, 36 spiral tip, 353
Ionic models, 40 spiral wave, 353
Relaxed Monodomain model, 132
Ischemia, 17
Roger-McCulloch model, 60
M cell, 11 T wave, 15
Magnetic resonance imaging (MRI), 19 morphology, 315
Mitchell-Schaeffer model, 60 Transmembrane action potential (TAP), 323
Monodomain model activation markers, 324
space discretization, 193 recovery markers, 324
variational formulation, 192 Transmembrane potential, 21
Monophasic action potential (MAP), 325 Traveling waves, 86
recovery markers, 325 fronts, 86
Morris-Lecar model, 60 pulses, 89
Index 397
Series Editors:
Editor at Springer:
Francesca Bonadei
[email protected]