You are on page 1of 58

Introduction to Cardiovascular Physiology

for
Students of Human Biology & Disease
Hesham A. Sadek MD, PhD
_______________________

Forward. This syllabus provides an introduction to basic


cardiovascular physiology. We will cover the materials in
Chapters 1 and 2 in the first Lecture and the materials in
Chapters 3 and 4 in the second Lecture. This material will
require study outside lecture to achieve the course objectives.

______________________

Chapter 1
Introduction to CV Physiology………………………………..………2

Chapter 2
Cardiac Mechanics & Excitation-Contraction Coupling…..…13

Chapter 3
Cardiac Electrical Activity………………………………………….…29

Chapter 4
Cardiac Control by the Autonomic Nervous ……………………..45

1
CHAPTER 1. INTRODUCTION TO CARDIOVASCULAR PHYSIOLOGY

This chapter introduces cardiovascular (CV) and cardiac function.

Major Concepts Introduced


1. Overall function and organization of the CV system
2. How the systemic and pulmonary systems are related.
3. The magnitude and variation of cardiac output.
4. Tthe role of valves in the CV system.
5. The pressures, volumes and flow rates in different parts of the CV system.
6. The relations between cardiac output, work, energy consumption & efficiency.

Figure 1. The two circuits of the cardiovascular system. Stippling indicates O2-rich blood.
Figure by Richard Moss (U. Wisc. Madison). With permission.

I. The Big Picture. In a nut-shell, the CV system rapidly circulates blood to within a few
micrometers of most cells in the body. The circulation is driven by the pump action of the heart,
whose failure is still the most common cause of death in the U.S. The CV system continuously
brings oxygen and nutrients to cells and removes CO2 and metabolic end-products. It overcomes the
limitation of diffusion to move substances over long distances, which would limit the growth of any
biologically active organism beyond a few microns in diameter. Nutrients are picked up in the
digestive tract, blood gases are exchanged to air in the respiratory tract, and the constituents of blood
are controlled and manipulated as blood passes through multiple organs, in particular the kidney, the
liver, the bone marrow, and the spleen. The circulation also provides a means to deliver hormones
quite rapidly to cells over long distances. The blood is moved in a unidirectional (one-way) fashion,
being driven by the pumping action of the heart, first through the arteries, then the capillaries, and
then the veins. This one-way movement of blood requires the presence of effective valves in the
heart and in the veins, which prevent backward blood flow.

In lower (primitive) animals, the heart operates as a single pump and has only one ventricle. Blood
may be pumped sequentially through the body (systemic system), a digestive tract, and a respiratory
network for oxygenation (see Fig. 1). In mammals, such as human beings, the heart consists of two
coordinated pumps that operate in series. One pump (the right side of the heart) moves blood

2
collected from the body to the lungs for gas exchange, and a second pump (the left side of the heart)
moves blood collected from the lungs out to the body. The two-pump arrangement allows us to have
greater metabolic rates and more effective delivery of blood throughout the body than a 1-pump
system could generate. The two pumps must operate in a highly coordinated fashion, and details are
forthcoming in this course. Here is the jist of things…

Figure 1

II. The Jist of CV Function. Figure 2 shows a block diagram of the total circulatory system. Blood
coming from the mass of the body enters the right atrium via the vena cavae, and from there it enters
the right ventricle. It is pumped to the lungs by the right ventricle and returns to the left atrium,
thereby completing one passage through the pulmonary circulatory system. From the left atrium,
blood passes to the left ventricle, and from there it is pumped into the aorta and passes through the

3
various capillary beds of the systemic circulatory system to return to the right atrium via the vena
cavae. Most of the blood returns to the heart after passing a single capillary bed. In some cases,
especially in the digestive tract, blood passes through two capillary beds.

DIASTOLE & SYSTOLE


The period of time during which the heart is contracting is called the systolic period, or simply
systole. The period between beats, beginning immediately when the cardiac contraction stops, is
called the diastolic period, or simply diastole.

Filling of the ventricles occurs in three parts. Most filling occurs as the ventricles relax. But some
filling continues more slowly between beats. A third phase, accounting for 10 to 15% of ventricle
filling, occurs in early systole as the two atria contract before the ventricle, thereby pushing
additional blood into the ventricles. Next, the two ventricles contract at nearly the same time. There
is NO mixing of blood between the right and left sides of the heart in normal people. When the right
ventricle contracts, the tricuspid valve closes (to the right atrium) and the pulmonary valve opens,
allowing (oxygen-poor) blood (from the body) to move into the pulmonary arteries. When the left
ventricle contracts, the mitral valve closes (to the left atrium) and the aortic valve opens (to the
aorta) so that (oxygen-rich) blood from the lungs can move to the body (‘systemic’ circulatory
system). During the entire diastolic period, the mitral and tricuspid valves are open to allow blood
to move into the ventricles. The return of blood from the venous side of the circulation is supported
by the presence of valves in the veins. When skeletal muscles contract, the ‘squeezing’ effect on
the veins ‘pressurizes’ the blood within them, opens the venous valves, and forces blood along the
veins toward the heart.

CARDIAC OUTPUT & STROKE VOLUME

The amount of blood pumped per minute by the heart is called the cardiac output. It must be equal
(on average) in the systemic and pulmonary systems. The cardiac output of a healthy 70 kg woman
is about 5 liters per minute, and her blood volume is also about 5 liters. Thus, it is evident that the
average RBC passes through the entire circulatory system at least once every min. With a heart rate
of 60 beats per minute, the average stroke volume of her heart (i.e. blood volume pumped per beat)
will be about 80 ml (which is more than 50% of the end-diastolic ventricular volume). During
exercise, as a reasonably conditioned woman, she can increase her heart rate to about 200 beats per
min and her cardiac output up to about 20 liters/min.
Cardiac Output = Stroke Volume X Heart Rate
NOTE: The volumes given here can be increased about 10% for a 70 kg man.

4
Figure 3. Overview of CV function.

Arteries are elastic (like a balloon), and veins are floppy (like a plastic bag).
Panel 5 of Fig.3 shows pressures in the different CV sections. In medical physiology, pressures are
given in millimeters of mercury (Hg), and you will be shown later how the pressures are measured.
In the heart and the large arteries, the pressures vary greatly with each heart cycle. These
oscillations depend first-and-foremost on the fact that the large arteries are elastic. You may think of
the aorta as being something like a balloon that gets a quantity of blood injected into it at each beat.
It expands with each injection, and the pressure within it increases. In order for the heart to inject
blood into the arterial trees, it must first develop enough pressure to overcome the pressure already

5
in the arteries. It is at that instant that the valves open and let blood out of the heart. The German
physiologists describe the elastic function of the large arteries as a ‘Windkessel’ (a wind-kettle; see
Fig. 4), and the principle is depicted in Fig 4. The heart moves blood into a compartment that has
elasticity, the aorta, which can expand to accommodate the blood. Thereby, the heart transfers
energy, invested to contract against arterial pressure, to the walls of the arteries. After the heart
relaxes, the arteries remain pressurized and drive blood into the small arteries and capillaries.

Figure 4. During systole, the stroke


volume ejected by the ventricle results in
some forward flow of blood toward the
small arterioles. But most of the ejected
volume is stored in the elastic large
arteries. During diastole, the elastic
recoil of the arterial walls maintains
capillary flow throughout the remainder
of the cardiac cycle.
Figure by Richard Moss (U. Wisc. Madison).
With permission.

The Concept of Compliance (or ‘Distensibility’)


‘Compliance’ describes how easily a vessel extends or expands and fills when it experiences a
pressure difference. If the aorta were a steel pipe (i.e. were completely non-compliant), the heart
would have to develop much more force to eject blood into it. And the pressure developed by the
heart would be felt instantly at the opposite end of the pipe. The easier it is for the heart to eject
blood into the aorta, the less work the heart has to do. One reason for decreasing CV function with
aging is that arteries (and the heart itself) tend to become stiff. The speed at which the arterial
pressure wave moves along the aorta and the peripheral arteries becomes greater as arteries become
stiffer (non-compliant). Some new therapies of high blood pressure reduce not only the actual total
peripheral resistance, but also (over longer periods of time) can reduce arterial stiffness.

On the one hand, you can see from this situation that arterial pressure becomes the driving force to
move blood through the CV system during diastole. To be effective in a long complex CV system, a
good deal of arterial pressure is essential. BUT, the heart struggles against this pressure to move
blood into the aorta at each beat. So, excess pressure in the arterial system is a serious strain on the
heart.

The pulmonary circulation pathway is much shorter than the systemic pathway, and therefore the
pressure needed in the ‘Windkessel’ driving blood through the pulmonary path can be much less. As
expected from this, pulmonary artery pressure oscillations are much smaller than those of the
systemic arteries, and the right ventricle does not need to develop as much force as the left ventricle.
As you know already, this is why the right ventricular wall is much thinner than the left ventricular
wall. The right ventricle does much less work than the left, although it moves just as much blood as
the left. Remember: Work = Force x Distance = Volume X Pressure!

6
What is ‘The Law of Laplace’?

The ‘Law of Laplace’ describes mathematically the physical stress occurring in the wall of a
hollow organ in dependence on its dimensions and the pressure gradient across its wall:

Wall Stress = P ● r / w

where P is pressure, r is radius, and w is wall thickness. You will need this physical principle many
times in the coming lectures. With reference to the left and right ventricles, you will find that wall
stress is approximately equal in the left and right ventricles during contraction. Why? Developed
pressure (P) is about 4 times higher in the left versus the right ventricle during systole, but the wall
thickness (w) is about 4 times less in the right ventricle. You will hear in later lectures that the heart
is capable of adapting via hypertrophy (or atrophy) to its experience of high (or low) pressure over
prolonged periods of weeks and months.

The oscillatory function of the arterial system is most pronounced in the larger arteries, and the
pressure oscillations become damped and disappear in the small arteries. Over the same arterial CV
stretch, the average (or mean) pressure in the arterial system is falling off steeply. The pressure
profiles and the blood flow become smooth in the smallest arterioles, and blood flow is completely
smooth in capillaries. The pressure in the CV system continues to drop through the capillaries and
venous sections on retuning to the heart.

The decline of pressure along the small arterioles is an important clue! Pressure falls off most
steeply in those regions that provide the most resistance to blood flow. Thus, it is clear that the
small arteries (arterioles) are generating the largest resistance to blood flow (see panel 5 of Fig. 3,
and Fig. 5). The resistance to blood flow can be adjusted in these arterioles by the contraction of
smooth muscle, and that is how blood flow to different organs and capillary beds is regulated. You
will hear much more about the regulation of blood flow by arterioles in later lectures. You will see
that a very small change of arteriole diameter has very big effect on blood flow through that
arteriole.

7
Figure 5. Normal pressures in the CV system

Over on the venous side, things are quiet and placid. There are only very small pressure
oscillations, which are UNRELATED to the pressure waves on the arterial side. The amount of
pressure in the final large veins is just enough so that blood flows rapidly into the ventricles when
they relax and expand. As noted with regard to Fig. 3, the majority of blood in the body is found in
the venous CV section of the systemic circulation. At the same time, the veins are extremely
expandable (called better compliant). While the arteries can be thought to function something like
a balloon being injected with liquid, the veins function more like a plastic bag. The veins can be
filled with very different amounts of liquid without generating much pressure at all. In this light,
we can think of the venous part of the systemic circulation as a lake (or possibly better as a swamp).
There is a large volume in this lake, and the lake is placid. This lake (or swamp) drains into the
heart. Some CV physiologists liken the return of blood into the ventricles to a continuous seepage
of water into a basement. The heart must continuously function as an effective sump-pump to keep
water out of the basement by moving it back above ground

Figure 6. The cardiac valves. Figure by Richard Moss (U. Wisc. Madison). With permission.

8
III. Getting to the heart of it. With this background, cardiac function can be placed in the context
of overall CV function: The return of blood to the ventricles defines a ‘preload’ for the heart. The
delivery of blood into the arterial sections of the CV system requires that the ventricles develop
more pressure than already exists in the large arteries. Thus, the arterial pressure determines an
‘afterload’ against which it must develop pressure. As already mentioned, blood fills the heart from
the venous side and exists to the arterial side because the four cardiac valves enforce this flow.
Valves in the veins also insure that blood moves toward the heart when the veins are compressed by
the contraction of skeletal muscle. Fig. 6 points out the function, placement, & anatomy of the four
heart valves.
Figure 7 gives a first look at the details of the cardiac cycle. From bottom to top, we have a record
of the electrocardiogram, the ‘ECG’, (that reflects the electrical activity of the heart), the two major
heart sounds (S1 and S2, that associate with the closing of the mitral and tricuspid valves, and the
aortic and pulmonary valves, respectively), the left ventricular volume, and the pressures in the left
ventricle, the left atrium and the aorta. You will be seeing many diagrams such as these in the
course of the CV lectures. Many of the details are obvious, if you remember the foregoing. Here,
we will outline only the major events of the cycle:

Figure 7. Records of atrial, aortic and


ventricular pressure waves in relation to aortic
flow, ventricular volume, heart sounds,
venous pulse and the ECG during a complete
cardiac cycle. This is often called a
“Wigger’s Diagram” after the physiologist
who first published it. Figure by Richard Moss (U.
Wisc. Madison) with permission.

9
The ECG, which will be discussed in much more detail in two forthcoming lectures, shows the earliest
signal of the cycle. That is the ‘P’ wave, which corresponds to the depolarization of the atria. The QRS
corresponds to the depolarization of the ventricles, and the T-wave corresponds to the T-wave of the
ventricles. This already indicates that the ventricular action potential must be very long. The
ventricular contraction begins during the QRS wave, and at this time pressure begins to build up in the
left ventricle. Associated with this rise of pressure, the S1 sounds indicate that the valves have closed
to the left and right atrium. Please note that the opening of valves does NOT cause heart sounds.
When ventricular pressure exceeds the aortic pressure, the aortic valve opens and ventricular volume
begins to decrease as blood is ejected. Whereas ventricular pressure matches venous pressure in the
diastolic period, ventricular pressure matches aortic pressure during the ejection phase when
ventricular volume is decreasing. As the heart stops contracting and ejecting blood, pressures begin
to fall and the aortic valve closes when ventricular pressure falls below aortic pressure. Associated
with this closure, comes the second heart sound(s). The left ventricle remains ‘isovolumetric’ (i.e.
‘isometric’) for about 50 to 100 ms, as left ventricular pressure continues to fall until pressure in the
ventricle matches or undershoots pressure in the left atrium. Then, the mitral valve opens and the rapid
phase of ventricular filling and distension begins. As the ventricle ejects blood, pressure rises in the
aorta. Since the load is increasing as the blood is ejected, this is not really a purely ‘isotonic’
contraction phase. We call it ‘auxotonic’, equivalent to contraction against a spring that becomes
tenser as it is pulled.
We will not discuss the left atrial pressure wave form: It is of small magnitude and considerable
complexity. And it is of great clinical importance. Therefore, it will be explained later with
appropriate care to detail by Dr. Hill. Also, we will not consider here the details of the arterial
pressure waveforms. For now, just note that the elasticity of the blood vessels gives rise to rather
complex hemodynamic phenomena that include ‘resonances’ and ‘reflections’ of pressure waves.

Figure 8. The Pressure-Volume Loop of the Left


Ventricle. The ‘P-V’ loop summarizes the changes in
left ventricular pressure and volume throughout the
cardiac cycle. The element of time is not evident in
such loops. Diastolic filling starts at “ESV” (end-
systolic volume) and terminates at ‘EDV’ (end-
diastolic volume), when the mitral valve closes.
Please note that there is only a small increase in
pressure with the increase in ventricular volume
during diastole. During isovolumetric contraction
(‘EDV’ to ‘Aortic Valve Opens’) there is a steep rise
Cardiac in pressure with no changes in ventricular volume.
Work When the aortic valve opens, ejection begins and
proceeds across the top of the loop from right to left
(‘Aortic valve opens’ to ‘Aortic valve closes’).
Pressure continues to rise briefly during ejection, and
there is a large reduction in volume. There is a small
decrease in ventricular pressure before valve closure.
Following aortic valve closure, this is isovolumetric
relaxation (‘Aortic valve closed’ to ‘ESV’), which is
characterized by a sharp drop in pressure with no
change in volume. The mitral valve then opens to
complete one cardiac cycle. Figure by Hilgemann, 2006

10
Cardiac Work & Efficiency. Figure 8 shows an extension of the ‘Wigger diagram’ that is very
useful to analyze cardiac function. This is the ‘Pressure-Volume Loop’. In this plot, the X-axis is
the Left Ventricular Volume and the Y-axis is the Left Ventricular Pressure. Imagine that you plot
these two variables against one another during a cardiac cycle. The cardiac cycle corresponds to
movement around the loop in a counter-clockwise direction (see figure legend for details). The
area formed by this loop corresponds to the work performed by the heart during the cardiac cycle
(Volume x Pressure = Force x Distance = Work). Since these loops will be used again in later
lectures (and again later in Cardiology), you should try to visualize how each of the 4 sides of this
loop correspond to the different parts of the cardiac cycle in Fig. 7. Also, try to visualize the
relationship of this diagram to the length-tension relations of skeletal muscle, as presented by Dr.
Stull. The lower part of the loop (ESV to EDV) corresponds to the ‘resting length-tension’. The
heart must be able to develop GREATER pressure than the pressure occurring when the aortic valve
opens. So, the peak ‘isovolumetric pressure’ would be higher at all filling volumes that the pressure
needed to open the aortic valve.

CARDIAC OUTPUT, WORK, OXYGEN CONSUMPTION & EFFICIENCY


Cardiac Output = Stroke Volume x Heart Rate
Cardiac Work = Cardiac Output x Mean Arterial Pressure
Cardiac Energy Consumption ~ Cardiac Oxygen Consumption
Cardiac Energy Consumption = Cardiac Work + Other Energy Expenditures
Cardiac Efficiency = Cardiac Work / Cardiac Energy Consumption = 25-35%

Cardiac output is the major ‘job’ of the CV system and it is extremely important to realize that the
same cardiac output can require different amounts of work, depending on blood pressures, aortic
compliance and blood volume. But there is still one more step: Finally, ‘energetics’ is central to
understanding cardiac function. The heart extracts a greater percent of oxygen from the blood
(between 10 and 30%!) than any other organ because its pump activity requires vigorous and
continuous oxygen consumption. In this light, the question of ‘cardiac efficiency’ becomes key.
How much energy must be consumed to carry out the work of the heart? The highest estimates of
cardiac efficiency are in the range of 35%. In other words, at least three times more energy is used
by the heart than would be needed to carry out its work, if were a perfect engine. You will learn in
subsequent lectures that cardiac contractility, heart rate and wall stress are the three main
determinants of cardiac oxygen consumption. Clinically, you will be able to influence all of the
variables that determine cardiac work and cardiac efficiency!

11
5.
4. heart rate
blood volume

3.
systemic
vascular
resistance

2.
1.
aorta compliance
cardiac contractility

Figure 9. Overview of major factors affection cardiovascular function. Figure modified from
Richard Moss (U. Wisc. Madison). With permission.

Summary & outlook: Figure 9 highlights major variables that influence CV function and that are
highly important in the practice of medicine. You will need to know how changes of each of these
variables changes CV function and elicits important reflexes that control CV function. If the aorta is
stiff (i.e. has low compliance) or if peripheral resistance is high (i.e. arterioles are contracted), the
heart must pump blood against higher pressures and therefore must do more work to maintain an
adequate cardiac output. Also, if the total blood volume is high, the CV system is stressed because
blood pressures are increased and there is a bigger volume to circulate (i.e. more work to do). As
you will know, the heart can be stimulated to do more work by increasing cardiac frequency and
contractility, namely with activation of the sympathetic nervous system. But it can be dangerous in
patients. The heart cannot do more work if it cannot supply the necessary energy to do so.
‘Whipping’ the heart to work harder is not good therapy for most cardiac patients.

12
CHAPTER 2. CARDIAC MECHANICS & EXCITATION-
CONTRACTION COUPLING: Intrinsic Control Mechanisms of the Heart

This chapter introduces cardiac function by considering the heart as a muscle specialized to be a
pump. Specifically, we will consider how cardiac excitation-contraction coupling and mechanics are
similar and different from skeletal muscle. Then we will consider how cardiac function is modulated
by the major parameters that are changing constantly during our daily life, independent of neuronal
reflexes. These variables are 1) ‘contractile state’ or ‘contractility’ of the cardiac myocytes, 2) the
‘preload’ of the heart determined by its filling, 3) the ‘afterload’ of the heart as determined primarily
by the arterial pressure which the heart must overcome to eject blood, and 4) the cardiac frequency
and excitation sequence.

Lecture Outline & Key Concepts

1. Methods used to study cardiac function


- Heart-lung preparation
- Isolated perfused heart
- ‘Superfused’ muscle strips and papillary muscles
- Patch clamp of isolated myocytes

2. Cardiac calcium homeostasis determines the cardiac contractile state (=contractility).


- Cardiac cells have long action potentials and refractory periods (i.e. no tetanus!)
- Ca influx by Ca channels contributes to activation of the myofillaments
- Ca extrusion takes place by Na/Ca exchange
- Cardiac contractility increases when Ca influx increases
- Cardiac contractility increases when cytoplasmic Na rises
- What are dihydropyridines (e.g. nifedipine) and heart glycosides (e.g. ouabain)?

3. Autoregulation of cardiac output (stroke volume) by cardiac preload (=enddiastolic filling)


- Frank-Starling Mechanism: Stroke volume increases when LV filling is increased

4. The influence of afterload (arterial pressure) on cardiac function


- Stroke volume decreases when arterial resistance increases

5. The molecular components of cardiac excitation contraction coupling


- Ca influx activates Ca release from the sarcoplasmic reticulum (SR)
- The cardiac SR Ca release mechanism undergoes prolonged inactivation
- What is ryanodine?

6. The influences of excitation pattern and frequency on cardiac function


- Contractility at a premature beat is decreased because Ca release is refractory.
- Contractility at a delayed beat is increased.
- What is post extrasystolic potentiation and ‘treppe’?

13
Dog Heart-Lung Isolated Perfused
Preparation Heart

i
pA

Isolated “Superfused” Patch Clamp of Single


Papillary Muscle Isolated Cardaic Myocytes
Fig 1. Methods to study cardiac function. Fig. by Hilgemann, 2006

I. Methods. Cardiac function depends on many variables at any moment. To study cardiac function,
it is essential to control and manipulate (1) sympathetic and vagal nerve activity, (2) cardiac
preload (i.e. filling), (3) cardiac afterload (i.e. arterial pressure), and (4) cardiac frequency. Fig. 1
shows four classical methods that allow controlled studies of cardiac function.
Dog heart-lung preparation. Vessels to the heart are cannulated and a pump is used to control
precisely cardiac filling and thereby the ventricular volumes.
Isolated perfused heart. The heart is isolated and major vessels are cannulated. Venous pressures
and arterial pressures are precisely controlled and measured.
Isolated ‘superfused’ muscle. Thin papillary muscles or strips of cardiac muscle are mounted in an
oxygenated muscle bath. Frequency is controlled by electrical shocks, if the muscle does not beat
spontaneously. If it does beat spontaneously, frequency can be monitored electrically.
Patch clamp of isolated myocytes. Cardiac myocytes are isolated by digesting the extracellular
matrix of the heart with proteases. Composition of the cytoplasm can be controlled through the
pipette. Ion channels and transporters that generate electrical current can be studied in detail.

14
Sinus node

‘ Working '
atrium

A - V node

Ventricle

Purkinje
fiber

Fig. 2 The heart is a functional syncytium.


Excitation travels from cell to cell.

Fig. 3 Structure of gap junctions. Fig. 4 Gap junctions couple myocytes at intercalated disks.
Figs. by R. Moss, U. Wisc. With permission

II. Introduction to cardiac excitation-contraction coupling (ECC). You will hear many times:
The heart is a ‘functional syncytium.’ What does this mean? The heart functions as a single unit.
Every cardiac contraction is preceded by an action potential that spreads from one myocyte to the
next. The excitation wave begins spontaneously in the right atrium, spreads across the atria, rapidly
downward along the septum, and then upward toward the heart base along the ventricles (see arrows
in Fig. 2). The heart can continue to beat when it is removed from an animal, at least for some time.
In short, it functions autonomously.

Electrical connections between myocytes are formed by ‘gap junctions’, which are ion channels,
composed of protein complexes called connexins. The large-diameter pores of these channels link
the cytoplasms of two neighboring cells (Fig. 3). Each myocyte generates one half of each channel,
called ‘hemi’-channels, and each half contains six subunits. Hemichannels remain closed until they
couple to a hemi-channel on a neighboring cell. When open they allow ions and solutes up to
molecular weights of about 700 to diffuse between cells. These channels close if the cytoplasm

15
becomes acidic or has high Ca, which can occur in cardiac ischemia. By closing, gap junctions
‘uncouple’ myocytes and stop electrical excitation from spreading. Gap junctions are mostly at the
ends of the myocytes in the ‘intercalated disk’ region (see Fig. 4). EXAM QUERY: Which other
organ contains cells with gap junctions at high density?

Fig 5. Cardiac excitation-contraction cycle. Fig. 6. Peak cytoplasmic Ca (free and total) and
contraction. Figs by Hilgemann, 2006

Cardiac contractility (the ability to contract) is controlled by calcium. Skeletal muscles must be
able to lift and hold a load. In other words, they must develop ‘tetanus’ (continuous) contraction. A
tetanus contraction in heart would negate its function as a pump, because the heart must relax to fill
and pump again. It is therefore important that the heart cannot tetanize. One mechanism that
prevents tetanus in heart is that action potentials are very long, typically 250 ms in ventricular
myocytes (see Fig. 5). This prolonged depolarization, usually above 0 mV, is called an action
potential ‘plateau.’ During the plateau, electrical shocks cannot evoke another action potential; this
is the ‘absolute refractory period’. During the second half of repolarization, a large electrical shock
can evoke partial action potentials. This is the ‘relative refractory period’.

Since all myocytes contract at each beat, cardiac contraction cannot be regulated by changing
the number of activated cells, as in skeletal muscle. Cardiac contraction is regulated by
changing the amount of Ca made available to activate the myofillaments in heart (see Fig. 6),
and as discussed later, catecholamines strongly increase the cardiac Ca transient. For this to
be possible, Ca transients must be much smaller in the heart than in skeletal muscle, so that a
small change of the Ca transient changes significantly the degree of activation and thereby the
magnitude of the contraction.

16
Isolated Electrically Stimulated Figure 7. Cardiac contraction strength depends
‘Superfused’ Papillary Muscle strongly on the cytoplasmic Ca transient. In contrast to
skeletal muscle, a substantial amount of Ca enters the
Ou abain cardiac myocyte at each beat and is extruded during
Contraction relaxation. Ca influx occurs via Ca channels that open
Force
Control immediately when the cell depolarizes, and Ca
extrusion occurs via Na/Ca exchange during
High-d ose repolarization and in early diastole. As illustrated in
Electrical
Stimulus
Dihydropyrid ine Fig. 7, drugs that block Ca channels (e.g.
dihydropyridines) can decrease cardiac Ca transients
and therewith cardiac contraction (see Fig. 7). In the
Free 0.2 s opposite direction, drugs like ouabain (a ‘heart
Cytoplasmic
Calcium
glycoside’) cause cytoplasmic Ca transients to increase
and thereby cause cardiac contraction to increase.
When Ca transients and contraction are decreased or
increased, we speak of a ‘negative inotropic’ or
‘positive inotropic’ effect, respectively.
Fig. by Hilgemann, 2006

Neither dihydropyridines nor ouabain has any effect on the contractile function of skeletal muscle at
the concentrations needed to induce inotropic effects in heart muscle. In contrast to heart, skeletal
muscle relies entirely on release of intracellular Ca to activate contraction.

a -a -
b AP upstroke via
Action Potential Na channels
c
-b - Figure 8. Cardiac excitation-
Ca influx by Ca channels contraction coupling.
& Na/Ca exchange
activate Ca-induced Ca It was discovered in the late
[Ca2+] i d
Contraction release from SR 19th century that the heart stops
-c- beating within a few beats when
0.1 s Ca uptake by SR Ca
pump
extracellular Ca is removed.
The reason is that Ca influx via
-d-
Ca extrusion via Na/Ca Ca channels provides a
exchange significant part of the Ca needed
for activation of the contraction,
Na/Ca exchanger
2+ (NCX) probably between 10 and 25%
Ca 2+
Ca 2+ 3Na
in man. In addition, Ca influx
‘Long’ (L) -type
calcium channel directly activates the Ca release
Calcium release
channels in the SR. Thus, Ca
3Na+ Ca2+ channel release is inhibited when there
=ryanodine
SR receptor is no Ca influx.
Sarcoplasmic Fig. by Hilgemann, 2006
Reticulum Calcium
Pump = SERCA

17
Review of the Players in Cardiac Excitation-Contraction Coupling and Their Function

Voltage-gated Ca channels. 10 to 25% of Ca needed to activate contraction in cardiac muscle comes into myocytes at
each beat via voltage-gated Ca channels. These channels activate (open) upon depolarization and inactivate slowly
during the plateau. Thus, they are called ‘slow’ Ca channels. We will discuss them more in the next cardiac lecture.
Na/Ca exchangers. The Na/Ca exchanger plays a central (but somewhat complex) role in heart. It exchanges 3 Na for 1
Ca. Thus, it makes an electrical current. 3 Na corresponds to 3 charges, whereas 1 Ca has only two charges. For this
reason, it is affected by the membrane potential. The energies of the ion gradients and membrane potential determine that
the exchanger brings Ca into cardiac cells at the beginning of the action potential and out of cardiac cells during
relaxation and between beats. To understand this, remember that positive membrane potential favors outward movement
of positive charges, and the exchanger moves 3 Na ions out when the cell depolarizes. At this time cytoplasmic Ca is
very low, so the positive potential creates a ‘driving force’ to briefly move Ca into the cell and Na out. If you are
confused, just remember that the exchanger is doing exactly what the cell needs…first, Ca influx and then Ca extrusion
during the action potential. When cytoplasmic Ca rises and repolarization has begun, the driving forces switch so that
the exchanger extrudes Ca.

For those who are comfortable with equations, try this: The exchanger always tries to move
cytoplasmic FREE Ca to an equilibrium concentration, given by the following equation:

[Ca] i = [Ca] o • ([Na] i / [Na] o) 3 • exp (Em/27mV)

If free cytoplasmic Ca is higher than the calculated value, the system will extrude Ca. If free
cytoplasmic Ca is lower than calculated, the exchanger moves Ca into cardiac cells. When
cytoplasmic Na increases or when the cell depolarizes, the system brings Ca into cells. When the cell
repolarizes, the system brings Ca out.

Cytoplasmic Na competes with cytoplasmic Ca, so a small increase of cytoplasmic Na strongly


inhibits Ca extrusion. No matter how you remember it, the operation of Na/Ca exchange explains
two basic findings about cardiac muscle, which are very different from skeletal muscle. First, in
experiments with isolated hearts, cardiac contraction strength increases dramatically when the
extracellular Na concentration is decreased. Second, cardiac contraction strength increases
dramatically when the cytoplasmic Na concentration increases. As you should remember, heart
glycosides such as ouabain and strophantidin inhibit the Na/K pumps, and the inhibition of just 5 or
10% of the pumps can have a strong effect on contraction. The cytoplasmic Na concentration may
rise only from 8 to 10 or 12 mM to have a large effect. In both of the cases, the Na gradient is
decreased, and the exchanger is less effective in extruding Ca. Thus, Ca accumulates in cells, it is
taken up by the SR, and more Ca becomes available for release to activate contraction. If
cytoplasmic Na reaches values of more than 15 mM, the heart cannot fully relax, and it develops a
continuous contraction, called a ‘contracture.’ Contracture leads quickly to myocyte death!

18
Na/Ca exchange in cardiac ischemia. The exchangers' function is important to understand
pathologies of the heart. For example, when circulation is compromised, as during a heart attack,
cardiac cells become acidic because lactate and other acids accumulate in the cytoplasm. In this
situation, the Na/H exchanger becomes activated and moves so much Na into heart cells (to extrude
protons) that heart cells become Na-loaded. When cardiac cells are Na loaded, the exchanger cannot
extrude Ca, and it rather moves Ca into the cells. We say that cells become 'overloaded' with Ca
when Ca cannot be extruded, and they develop a continuous contraction, called a 'contracture.' The
continuous presence of Ca, and therefore contraction, can lead quickly to cardiac cell death.
SR Ca release channels (RyR, ryanodine receptors). The cardiac sarcoplasmic reticulum (SR) Ca release channels
are activated by Ca influx, and they release a large portion of the Ca stored in the SR at each beat. Their function
explains why the cardiac contraction cannot summate to a tetanus. It would be catastrophic if the heart were to develop
a maintained (tetanus) contraction, as in skeletal muscle, since it would no longer rhythmically pump blood. Two factors
hinder development of a tetanus. First, as described in the previous lecture, the ‘refractory period’ of cardiac myocytes is
very long because the action potentials are long. It is not possible to evoke a new action potential until repolarization is
completed. Second, the cardiac SR Ca release channels undergo ‘inactivation’ with each cardiac cycle. It is the rise of
cytoplasmic Ca that induces inactivation. The inactivation is very strong. After Ca is released once, the release process is
refractory. It takes up to 1 or 2 seconds to completely recover and be able to release Ca effectively again. Half-time for
recovery from inactivation is about 0.5 seconds, whereas the voltage-gated channels are ready to open again almost as
soon as repolarization is complete.

SR Ca pumps. As in skeletal muscle, the cardiac sarcoplasmic reticulum contains a high density of Ca pumps that
transport into the reticulum at the expense of ATP hydrolysis. The relaxation rate of the heart is determined mostly by
the rate at which the SR Ca pump removes Ca from the cytoplasm. Remember, that most of the cytoplasmic Ca is
pumped back into the SR at relaxation; 10 to 25% is extruded from the myocytes by the Na/Ca exchanger. At a
molecular level, the SR Ca pump is similar to the Na/K pump of the surface membrane. It is a 'P'-type pump. It binds
ATP, and it phosphorylates itself during each cycle of Ca uptake. It moves 2 Ca ions into the SR for each ATP
hydrolyzed.

Na/K pumps. Don’t forget the old gray mare...the Na/K pump! Of course, the Na/K pump is also an integral part of
cardiac excitation-contraction coupling. It generates the Na and K gradients used to make action potentials and used to
extrude Ca and protons from the heart. Small changes of its activity have a big effect, as just described in connection with

19
Na/Ca exchange.

SNEAK PREVIEW: We will discuss how cardiac Ca handling is regulated, to regulate contractility, in a later lecture. For
now, please take note that changes of the myocyte Ca transients regulate more than just contraction. Ca transients also
regulate glycolysis and oxidative phosphorylation in the mitochondria. The mitochondria sense the cardiac activity by
taking up about 10% of the Ca released to the cytoplasm at each beat and slowly returning it to the cytoplasm between
beats. It is a channel-type mechanism lets Ca into the mitochondria, and a Na/Ca exchange system lets it back out
between beats. Finally, it is impressive that Ca transients also strongly regulate important transcription factors in the
heart, which control cardiac growth. Thus, cardiac Ca homeostasis is important in all aspects of myocyte life.

What are dihydropyridines? Specific inhibitors of ‘L-type’ Ca channels.


Dihydropyridines are members of a class of drugs that inhibit ‘L-type’ Ca channels of smooth
muscle and cardiac muscle, but not the ’N-type’ Ca channels of neurons. ‘L-type’ indicates that
these Ca channels open for relatively Long times. Dihydropyridines were developed by Bayer
corporation to be powerful vasodilators with few or no other effects. The expectation was that these
drugs would increase coronary blood flow in ischemic heart disease and decrease blood pressure in
hypertensive diseases. ‘Nifedipine’ was the lead compound. These drugs can be used to inhibit Ca
channels in experiments with cardiac myocytes, but the concentrations needed to induce a negative
inotropic effect in heart about 10 times higher than needed to induce relaxation of arterial smooth
muscle. In an animal, blood pressure would be fatally reduced by a concentration of nifedipine that
would have a substantial negative effect on contractility of the heart (i.e. a ‘negative inotropic
effect).

What are heart glycosides?


As discussed in a previous lecture, heart glycosides such as ouabain are unusual steroidal
compounds that inhibit the Na/K pumps of all animal cells. They are synthesized by a number of
plants, the most famous plant being ‘fox glove’, and may serve as natural insecticides. When
administered carefully, heart glycosides increase cardiac contractility with few other effects. They
are however extremely toxic at concentrations just few fold higher than needed to increase cardiac
contraction.

20
How do heart glycosides increase cardiac contraction strength?
When a small fraction of the cardiac Na/K pumps is inhibited by heart glycosides, cytoplasmic Na may rise from 7
to 10 or 12 mM. This small rise is enough to inhibit Ca extrusion by the Na/Ca exchanger. Less Ca is then extruded
during each contraction cycle, and Ca accumulates in the sarcoplasmic reticulum. As a result, more Ca is released at
each beat, and the cardiac contraction strength increases. Similar small changes of cytoplasmic Na have
comparatively small effects on the function of other cell types. Thus, it would seem that the heart is much more
dependent on Na/Ca exchange to control its cytoplasmic Ca than most (or all) other tissues.

What is ryanodine? Ryanodine is a biological insecticide found in the bark of some exotic trees in South America. It
was used quite extensively in garden agriculture in the past. It kills insects by putting their flight muscles into
contracture (i.e. continuous contraction) by permanently opening their SR Ca release channels. Mammalian skeletal
muscle also goes into contracture upon exposure to ryanodine. The heart does not go into contracture because the Na/ Ca
exchanger can powerfully extrude the Ca released from the SR when ryanodine is applied. Thus, in an isolated cardiac
muscle preparation, ryanodine is a strong negative inotropic agent.

Isolated Electrically Stimulated


‘Superfused' Papillary Muscle

High Preload
% of
Force Maximal Skeletal
Control Isometric (active)
Tension
‘Resting Cardiac
Low (active)
Tension' Prelo ad
induced by
stre tch
Cardiac
(passive) Skeletal
(passive)
0.2 s
Free
Cytoplasmic Sarcomer length (microns)
Calcium
(Modified)

Fig. 9. Effect of resting length of a cardiac Fig. 10. Comparison of length tension relations for a skeletal muscle
papillary muscle Fig. by Hilgemann, 2006 fiber and a cardiac papillary muscle

III. Introduction to cardiac mechanics. Like skeletal muscle, cardiac muscle contracts more forcefully when stretched
(distended) over the range of sarcomer lengths occurring physiologically. The effect is even more pronounced in heart.
Fig. 9 illustrates how contraction of an isolated papillary muscle changes when is stretched by 10 and 15 % from its
‘resting length.’ Contractions are strongly increased as the ‘resting’ (=passive) force is increased by stretch with little or
no change of the cytoplasmic Ca transient.

21
Cardiac length-tension relations. As outlined in Fig. 10, there are substantial differences between the length-tension
relations of skeletal and cardiac muscle. Physiologically, skeletal muscle is limited in how far it can be stretched by the
skeleton, namely by the physical restrictions of joints. When isolated, skeletal muscle can be stretched readily to a point
where thin and thick filaments no longer overlap (see Skeletal (passive)). Cardiac muscle must provide its own
resistance to stretch as it fills. Thus, it must be much stiffer. The ‘passive’ resting force rises steeply as a cardiac muscle
is stretched beyond sarcomer lengths of 2.2 microns (Cardiac (passive)). Most of the ‘stiffness’ is accounted for by a
giant protein that is linked to the Z-bands and traverses the entire length of the sarcomer between filaments. This protein
is (not surprisingly) called ‘Titan.’

The ‘active’ curves in Fig. 10 give the peak isometric force developed when these muscles are excited to contract at the
corresponding lengths. The active tension of skeletal muscle rises steeply in the range of 1.5 to 1.8 micron sarcomer
lengths, and this probably corresponds to the degree of overlap of thick and thin filaments to form active cross bridges.
In cardiac muscle, developed force is in some way inhibited until sarcomer lengths of more than 1.8 microns are
achieved. Between 2 and 2.3 microns, the developed force increased very steeply with stretch (see vertical arrows).
This rise of force with stretch falls just in the range of physiological stretch experienced by myocytes in the ventricles as
the ventricles are filling. The implication and mechanism of this effect are described next.

“Vmax”

High
Contractility

Low

High
Contractility

Low

0 Load (gm)
Figure 12

Cardiac force-velocity relations. Just as in skeletal muscle, cardiac myocytes lift a load more quickly (i.e. can eject
blood more quickly) when the load is small (see Fig. 12). The fastest and most extensive contraction occurs with no load
at all, and the velocity of contraction in this ‘unloaded’ state is called the ‘V-max.’ When the load is larger than the
myocytes can lift, there is no shortening, and in this state the myocytes contract in an isometric fashion. In between
Vmax and the isometric state, myocytes will contract isometrically and develop force up to the point at which load can be
lifted. Then, while the load is being lifted, they contract isotonically.

22
When thinking of the intact heart, remember that ‘isometric’ really means ‘isovolumetric’. The heart contracts
‘isovolumetrically’ up to the point that the valves open. The ‘load’, which we call ‘afterload’ in the cardiac cycle,
corresponds (approximately) to the pressure in the aorta against which the heart ejects blood. When the aortic pressure is
higher, the heart ejects blood less vigorously and less blood is ejected. Thus, when arterial resistance is increased, the
cardiac output is at first decreased. To compensate the decrease of cardiac output, cardiovascular reflexes must be
rapidly activated to increase the contractility of the heart (see the upper force-velocity relation in Fig. 12) and the cardiac
frequency. These mechanisms are the subject of later lectures. For now please note that the work the heart performs
(stroke volume x pressure) increases as the arterial pressure increases, while the cardiac output (approximately
proportional to contraction velocity) decreases. As the afterload increases further, the work performed by the heart
decreases and would be null if the aortic pressure were so high that the cardiac contraction could not generate an
equivalent pressure in the ventricle.

Immediate effects of changing cardiac preload,


afterload, & contractility in the absence of reflexes
Changes of Preload Changes of Afterload Contractility Changes
170 170
170

120 120
120

70 70
70
160 160
160

80 0.2 s 0.2 s
80 0.2 s 80

0 0
0
150 150
150

100 100 100

50 50 50
Figure 13. Immediate effects of preload, afterload and contractility changes on cardiac function.
Fig. by Hilgemann, 2006

IV. Immediate effects of preload, afterload and contractility changes on cardiac function in the absence of reflexes.
We now consider how cardiac function is altered when the three parameters just outlined change (see Fig. 13).

Effects of preload. The left panel of Fig. 13 describes the effects of increasing ventricular filling
(=preload) on left ventricular (LV) volume, LV pressure and arterial pressure. Over the
physiological range of filling volumes, the ventricle contracts more forcefully and thereby ejects
more blood when the filling volume at the onset of the contraction is larger. Thus, the end-systolic
volume remains nearly constant. Peak pressures are somewhat increased. This is a very important
auto-regulatory mechanism of the heart that we call the ‘Frank-Starling mechanism.’ In short, stroke
volume of the heart tends to increase in proportion to an increase of the preload (i.e. filling), so that
the end-systolic volume remains rather constant.

23
Frank- Starling Mechanism for the beating heart.
The increase of cardiac contraction strength that occurs with increasing end-diastolic volume is called the ‘Frank-
Starling relation’, after its discoverers in the early 20th century. In skeletal muscle, the activation of contraction by
stretch probably reflects an increased overlap of thick and thin filaments. In heart, the major effect occurs at longer
sarcomer lengths. A simple explanation is the following: As the myofilaments are stretched, they become compressed with
respect to one another. As lateral spacing becomes restricted, cross bridge formation is enhanced and activation by
calcium binding to Troponin C becomes more effective.

Effects of afterload. The middle panel of Fig. 13 shows the immediate effect of increasing cardiac afterload (i.e. aortic
pressure) on cardiac function. With higher aortic pressure, the heart must develop more force before the aortic valve
opens, and it must eject blood against a larger resistance (pressure). Thus, the stroke volume is decreased, as the cardiac
function is forced to right along the force-velocity curves described in Fig. 12. This is a situation that must be
compensated in the body. To a certain extent it will be compensated at subsequent beats because the end-diastolic volume
will be increases when the heart ejects less blood (i.e. stroke volume is decreased). Why? Naturally, the Frank-Starling
mechanism will come into play to increase the output when the diastolic volume is increased.

Effects of contractility. The right panel of Fig. 13 shows the immediate effects of changing contractility on cardiac
function. When the heart contracts more forcefully (e.g. when a patient is treated with ouabain), more blood is ejected
into the aorta. The end-systolic volume is decreased and the peak LV and arterial pressures are somewhat increased.

Immediate effects of changing cardiac preload,


afterload, & contractility in the absence of reflexes
Figure 14. Plot of major
Changes of Preload Changes of Afterload Contractility Changes parameters from Fig.13.
170
170 systolic Lower plots, end-diastolic
170
(green) and end-systolic
(yellow) LV volumes.
120
120 120

70
70 diastolic Middle plots: Peak and
diastolic LV pressures.
70

160 160 systolic


Upper plots: Systolic and
160

80 80
80
diastolic arterial
0 diastolic pressures. Note how each
parameter change (red
0 0
150
end-
150 150 diastolic
arrows) affects other
100 100
100
Stroke
Volume parameters.
Fig. by Hilgemann, 2006
50 50 50 end-
systolic

24
For better clarity, the major parameters from the records in Fig. 13 are plotted in Fig. 14. Arrows indicate the variable
that is assumed to change. Preload changes affect first end-diastolic volume. Afterload changes affect first arterial
pressures. Contractility changes affect first end-systolic ventricular pressure. You should now be able to explain all of
the changes described here and the major underlying mechanisms.

The ‘Anrepp’ effect.


One ‘take-home’ message of this lecture is that the heart must be able to adjust its output to match the needs of the body.
Much of the feed back control involves reflexes of the autonomic nervous system that are the subject of later lectures.
We stress here that the heart has an important, but limited, intrinsic ability to increase its contractility by increasing the
cytoplasmic Ca transient, when either cardiac preload or afterload is increased.

Cardiac myocytes experience physical stress whenever preload or afterload is increased. If you think about it, it will be
clear that those elements of the myocytes that transfer force to their neighboring myocytes will experience a stretch
during either an increase of preload or afterload. With both an increase of preload and afterload, a mechanism is activated
that can increase the cytoplasmic Ca transient by 10 to 20%. This response takes 20 to 60 seconds to develop, and it is
called the ‘Anrepp’ effect. The mechanism appears to involve the activation of stretch-activated nonspecific ion channels
that will tend to load myocytes with Na and Ca.

Figure 15

V. Intrinsic effects of cardiac frequency and excitation pattern.


Many of us have an occasional extrasystolic beat because an ectopic excitation originates
somewhere in the ventricles outside of the regular conduction pathway. As discussed earlier, it is
important that contractions cannot summate in the heart. The contraction strength of the extra beat at
a shortened interval is small (see Fig. 15) because the Ca release mechanism of the SR is inactivated
and cannot release Ca. The cardiac release mechanism of the SR inactivates with each Ca release
cycle, and the recovery takes place with a time constant of about 1 s.

25
Delayed beats. It may also happen that conduction to the ventricles occasionally fails, and a beat is missed. In this case
there is an abrupt lengthening of the interval between beats. When this occurs, the resulting ‘post-rest’ or ‘delayed’ beat
is large. We can also explain this by the inactivation of the SR-Ca release mechanism, or more accurately by its recovery
from inactivation. At the usual cardiac contraction frequencies of 70 to 140 per min, the release mechanism does not
have time to recover completely from the inactivation induced by the previous Ca release. Under normal conditions, the
SR is releasing most, but not all of its Ca at each beat. There is some reserve Ca accumulated in the SR, and this reserve
becomes available when a beat is missed and the contraction interval is longer. At the delayed beat, it is important to
remember that cardiac filling will be larger than normal, because some additional venous return has occurred in the
absence of contraction. The extra Ca release supports ejection of the extra blood together with the Frank-Starling
mechanism that is also brought into play as result of the increased ventricular filling.

Figure 16. Post-extrasystolic


Frequency ‘inotropy’ or ‘Treppe’ potentiation and frequency
or ‘Bowditch Staircase’ inotropy. Fig. by Hilgemann, 2006

0.5 Hz 2 Hz 0.5 Hz

Post-extrasystolic potentiation. Now, imagine that several extra beats occur briefly at a high rate (see Fig. 16). Ca
enters the heart cells at each extra beat, but the SR does not release Ca. Thus, the SR becomes very loaded with Ca.
When the regular beat pattern continues, the next contractions are strongly enhanced or 'potentiated' because more Ca can
be released. This phenomenon is called 'post-extrasystolic' potentiation.
‘Frequency inotropy’ or ‘treppe’ or ‘Bowditch staircase.’ Finally, imagine that the heart is paced
at a steady, controlled frequency. When this is carried out, one finds that the contraction strength
increases over the course of 10 to 100 beats when the frequency is increased to a new higher steady
state. And the contraction strength again decreases when the stimulation frequency is decreased (see
lower record in Fig. 16). What is the mechanism? We have already seen that the Na gradient is a
very important determinant of cardiac contractility. Cytoplasmic Na will naturally increase with
increase of frequency. At each beat, Na enters cardiac cells via channels and Na/Ca exchange. It is
extruded slowly between beats by the Na/K pumps. Thus, an increase of frequency will result in an
increased Na influx per unit time, and extrusion by the pump does not keep up. As Na increases, Ca
extrusion will be decreased, because Na/Ca exchange cannot extrude Ca so effectively. And
contraction increases for the reasons already outlined.

26
AFFICIONADO”S APPENDIX

Many of you will be on pins and needles to see how Preload, Afterload, and Contractility changes affect the P-V
Loop’s of the Left Ventricle.

Isolated Perfused Working Heart


Changes of Preload Isovolumetric
End-Systolic
170
300 Pressure-Volume
Curve
120

240
70

160
180
0.2 s
80

0
120

150
60

100 15
0 60 120 180
50 LV Volume (ml)
Figure 17. An increase of preload shifts the isovolumetric ejection phase to the right (to higher pressures). Owing to the
Frank-Starling Mechanism, the heart contracts more vigorously and nearly the same end-systolic volume is achieved.

27
Isolated Perfused Working Heart
Changes of Afterload Isovolumetric
End-Systolic
170
Pressure-Volume
300
Curve
120

240
70

160
180
80 0.2 s

120
0

150
60

100 15
0 60 120 180
50 LV Volume (ml)

Fig. 18. An increase of afterload forces the heart to develop more pressure before the aortic valve
opens. Contracting against a larger pressure, less blood is ejected.

Isolated Perfused Working Heart


Isov olumetric
Contractility Changes E nd- Sy stolic
160 P ress ure-V olume
300 Curv es

120

240

80

160
180
0.2 s
80

120
0

150
60

100 15
Ouabain
Control 0 60 120 180
50 Dihydropyridine LV Volume (ml)
Overdos e

Fig. 19. Changes of contractility change the vigor with which blood is ejected. The end-systolic volume
is decreased. The isovolumetric end-systolic pressure-volume curve is shifted upward. In other words,
contractility changes would change the peak isometric contraction force, IF the heart were contracting
isometrically against a load that it cannot hold.

28
CHAPTER 3. CARDIAC ELECTRICAL ACTIVITY

This chapter describes the electrical activity of the heart, which initiates the heart beat. We
introduce the different shapes of cardiac action potentials in different regions, the cardiac
electrocardiogram (ECG), the major ion channel types, and the specific roles of those ion channels
in generating cardiac action potentials and cardiac pacemaking in the sinoatrial (SA) and
atrioventricular (AV) node cells.

KEY CONCEPTS
1. To recognize the shapes of the cardiac action potentials from different regions.
2. To know which cardiac cells determine heart rate.
3. To know which cardiac cells have the greatest propensity to become pacemakers.
4. To know how the electrocardiogram is generated.
5. To know which ion channels are most important and how they generate cardiac action
potentials.
6. To describe the electrophysiological basis of sinus node pacemaking.
7. To describe the basis of cardiac Ca homeostasis

Review of Basic Electrophysiological Principles

1. Ion channels determine membrane potential as a result of their selective permeability to ions.
When ions diffuse through a channel, they leave ions of opposite charge behind and thereby
charge the inside of the cell oppositely to their charge. When potassium goes out, the inside
becomes more negative.

2. Ion channels with perfect selectivity to one ion have a reversal potential equal to the Nernst
potential for that ion.

3. The current generated by each ion channel, when it is open, tends to move membrane potential to
the reversal potential of that channel.

4. The current generated by each channel, when open, depends on the 'driving forces' of ion
gradients and membrane potential. This ‘force’ is given by the difference between the membrane
potential and the reversal potential of that channel (Em - Erev).

5. Voltage-gated Na and K channels are activated (opened) by depolarization. We will now also see
examples of ion channels that are activated (opened) by repolarization and deactivated (closed)
by depolarization.

6. The RATE OF CHANGE of membrane potential is directly proportional to the total


membrane current. If the membrane is depolarizing, there is a net influx of positive charges
or expulsion of negative charges.

29
I. Cardiac electrical activity

Figure 1. Cardiac electrical activity (© J.A. Hill, 2004)

The heart initiates its own electrical activity. As shown in Fig. 1, the action potentials in all
myocytes are much longer than those of nerve axons or skeletal muscle. We will see that all cardiac
cells can, in principle, depolarize from their resting potentials and generate spontaneously an action
potential. However, the excitation process is highly organized and proceeds in a very precise
sequential fashion. Excitation begins in the SA (sinus-atrial) node cells. These cells (perhaps a few
thousand) are located in the enodcardial (interior) region of the right atrial wall very close to the
entrance of the superior vena cava into the atrium. SA node cells are the fastest 'pacemakers' of the
heart. They depolarize spontaneously during the diastolic period (between beats) and reach a
threshold that initiates an action potential faster than any other cells in the heart.

The SA node cells are specialized to be pacemaker cells, mostly because they LACK certain ion
channels, not because they have extra channels. Also, they do not build so rich a myofilament
network, as in the majority of atrial cells whose life-time job is to contract. Thus, we speak of
‘working’ atrial myocytes versus the small minority of ‘pacemaker’ atrial myocytes.

The action potential upstroke of SA node cells is slow.(see top action potential), and SA node cells
repolarize to only about -65 mV, instead of -80 or -85 mV achieved by most atrial myocytes.

30
Because the sinus node cells are coupled to neighboring atrial cells by gap junctions, excitation
spreads rapidly across the right atrium to the left atrium and downward toward the ventricles. The
‘working’ atrial cells have action potentials with a fast upstroke and a duration (from upstroke to
80% repolarization) of around 100 ms.

The excitation wave is transmitted from the atria to the ventricles ONLY at a specialized site
containing the so-called AV (atrio-ventricular) node cells. These cells, like the sinus node cells, also
show spontaneous depolarization and therefore are pacemaker cells. BUT they normally are not,
because the SA node cells pacemake faster and ‘lead’ the heart. If the SA node cells fail, or the
conduction to the node fails, the AV node cells will initiate ventricular contraction. This is called an
‘escape’ mechanism, and other myocytes can also initiate escape excitation, sometimes with
deleterious consequences.

The excitation is then transmitted through the so-called ‘bundle of His’ and the ‘Purkinje fibers’.
The Purkinje fibers run along the walls of the septum and then up the epicardial surface of the
ventricles. Like SA node cells, they also do not have a well developed contractile apparatus. The
Purkinje fibers can transmit excitation very rapidly because they have a very high density of Na
channels. Thus, the excitation moves very quickly through the Purkinje fibers to the apex of the
heart. At this time, the wave of excitation includes myocytes in the septum and it is transmitted
upwardly through the ventricles and across the ventricles from their endocardial side (i.e. the inside)
to the epicardial side. Importantly, the ventricular myocytes on the epicaridial side of the heart
repolarize more quickly than those on the endocardial side. As a result, repolarization occurs first on
the outside (epicardial side) of the heart and then on the inside (see Fig. 2). In other words, the
depolarization wave moves from inside to outside, and repolarization from outside to inside.

The ventricular action potentials are typically twice as long as the working atrial action potentials
(200 versus 80 ms from upstroke to 80% repolarization). Both remain depolarized in the positive
potential range for a prolonged time. This is called a ‘plateau’. By contrast, the Purkinje fibers
typically repolarize quickly to a potential close to 0 mV, and they maintain a plateau in this range for
100 to 150 ms, before repolarizing rapidly to around -85 mV.

As indicated on the Purkinje action potential, the cardiac electrical cycle is often divided into four
phases. Phase 0 is the upstroke phase. Phase 1 is a characteristic early repolarization phase, which
can also occur in some ventricular myocytes. Phase 2 is an extended plateau period. Phase 3 is the
major repolarization phase, and Phase 4 is the time from the end of repolarization to the next
upstroke.

31
The detailed coordination of the excitation wave requires that conduction velocities for the different
cardiac cell types are quite different, and we will see shortly that a major reason for these differences
is that the different myocyte types have different densities of Na channels:

SA node 0.05 m/s


inter-nodal tracts 1.0 m/s
atrial muscle 0.3 m/s
AV node 0.05 m/s
His-Purkinje cell 3.0-5.0 m/s
ventricular muscle 0.3 m/s

Conduction velocities of cardiac cells

II. The Electrocardiogram (ECG).


The lower record in Fig. 1 is the ECG, one of the most important non-invasive diagnostic signals in
medicine. ECG recording was developed by the Dutch Physiologist, Willem Einthoven, in 1913. In
short, Einthoven realized that fairly simple electrical recordings could be used to follow the
electrical activity of the heart in substantial detail on a millisecond and even submillisecond time
scale. Within this lecture, we can only just introduce the principles of how the ECG is generated and
recorded.

How is the ECG generated? Whenever one part of the heart is at one electrical potential and a
neighboring part of the heart is at another potential, a current flows between the two regions. This
current, between significant masses of heart muscle, gives rise to an electrical field that can be
measured on the surface of the body. Although the ECG is of small magnitude, on the order of 1 mV
or less, it can be measured very reliably, and conveys a great deal of diagnostic information.

The ECG shown in Fig. 1 is the typical waveform that is measured from two leads attached to the
right and left arms. The deflections observed are designated by letters, the major ones being called
the ‘P’, ‘Q’,’R’, ‘S’ and ‘T’ waves. To reiterate, these waves correspond to electrical fields that
develop instantly when a significant zone of myocardium is at a different potential from the rest of
the myocardium. The differences occur almost like ‘flashes’. The flashes can be viewed only from
certain angles (i.e. electrode positions), and the propagation of the electrical activity can be mapped
precisely by recording from different positions on the body.

Figure 2. We imagine a large cell with an inside-negative


potential. Electrodes placed at the ends of the cell detect no
potential difference. If one side of the cell depolarizes (i.e.
becomes negative outside and positive inside), then the electrode
on that side detects a negative signal, while the electrode on the
opposite side detects a positive signal. When the entire cell is
depolarized there is again no electrical difference between the
two electrodes. Copyright, 2003, U.T. Southwestern Medical Center, Office
of Medical Education, Web Curriculum

32
For clarity, Fig. 2 illustrates the principles of ECG recording in terms of ‘dipoles.’ In summary,
each myocardial cell can be thought of as a dipole, as depolarization sweeps over it. Many dipoles
then add as vectors, which have both magnitude and direction. The sum of all dipoles at any
instant in time can be represented as a single cardiac vector (see arrows in Fig 3). An electrode
designated ‘positive’ in an ECG recording will show a positive deflection when a wave of
depolarization is moving toward that electrode. The electrode designated ‘positive’ will show a
negative signal if a wave of repolarization is moving toward it. Imagine now that the electrodes
were rotated by 90 degrees in Fig. 2, so that they measure the potential difference perpendicular to
the depolarization wave. What would they measure? Nothing! The electrical fields experienced by
the two electrodes would be changing in an identical time course. So, there would never be a
difference.

Cellular basis of the ECG waveform. In brief, the P-wave corresponds to the depolarization of the
left and right atria, as the excitation wave moves across and down the atria. Passage of the electrical
wave through the conduction system into the upper left region of the septum does NOT give rise to a
signal, as the number of cells involved is relatively small. The first substantial mass of ventricular
cells that depolarize is the upper left part of the septum, and this depolarization gives rise to the
small downward ‘Q’ wave, as the wave moves downward and across the septum toward the right
side. The large ‘R’wave corresponds to the depolarization of most of the ventricular myocardium.
Since the Purkinje fibers run along the endocardial surface of the septum and up the inside walls, the
excitation begins on the endocardiac side and moves to the epicaridial side. The ‘S’ wave
corresponds to a final portion of ventricular depolarization, as the wave moves upwardly into the
base of the left ventricle. From the end of the S to the T wave, the ventricles are rather uniformly
depolarized. Thus, there is no signal. The atria have already repolarized before the ST segment.
Their repolariztion is effectively hidden by much larger QR signals arising from ventricular
depolarization.

The last major wave is the ‘T’ wave. This corresponds to the repolarization of the ventricles.
Whereas depolarization occurs from the endocardial side to the epicardial side (outside) of the heart,
repolarization starts on the epicardial side and proceeds to the endocardial side. This sequence
reflects the fact that action potentials are longer on the endocardial side of the heart (see Fig. 1).
Why are the R- and T-waves both upward in sign in the Lead I recording? The repolarization wave
moves in the opposite direction as the depolarization wave, and the direction of the membrane
potential change is also opposite. REMEMBER: two negatives make a positive!

Two important ECG intervals are indicated in Fig. 1. The ‘PR’ interval is the time from the onset of
the P wave to the onset of the R wave. PR intervals of more than 200 ms are indicative of
conduction problems between the atria and the ventricles (i.e. in the AV-node & His bundle
regions). The ‘QT’ interval gives the time from onset of ventricular action potentials to the end of
ventricular repolarization.

33
Right Left
- +
P
Arm Arm

Q
T
-

+
ECG
-
R
S

+
-

+
Figure 3. Generation of the ECG. Arrows indicate the direction of the movement of the electrical wave.
Note that the sign of the arrow for the T-wave is opposite to the sign of the other arrows. P, Q, R and S all
reflect depolarization of cardiac tissue mass, while the T reflects repolarization of ventricular tissue mass.
Figure by Hilgemann, 2006.

Standard ECG Leads. Figure 4 shows the standard leads used in ECG recording. Nine recording
electrodes are normally employed, and one ground electrode. One recording electrodes is placed on
the right arm (RA), one on the left arm (LA) and one on the left leg (LL). Six electrodes are placed
on the chest (as shown in Fig. 4), and the grounding electrode (not shown) is placed on the right leg.
These three leads form the “Triangle of Einthoven “, and there are three possible ‘bipolar’
recordings from these leads. Lead I is the left arm-right arm difference, Lead II is the left leg –right
arm difference, and Lead III is the left leg-left arm difference. The P and T waves are usually
largest in Leads I and II. That is because these waves occur almost perpendicular to the axis of
recording in Lead III, and therefore are not seen well in the Lead III recording. The axis of Lead I is
defined as 0 degrees.

The three electrodes placed on the extremities are also used to make ‘unipolar’ recordings. The
reference voltage for each of these unipolar leads is combined potential of the two opposite bipolar
leads. These three leads are called the augmented limb leads, and they are designated aVR, aVL,
aVF. These unipolar limb leads are not very different from the bipolar limb leads just discussed.
They are designed to have larger deflections than the standard limb leads.

34
The precordial leads are called V1,
V2, V3, V4, V5, and V6 - These
unipolar leads examine the ECG in
the horizontal plane around the heart.
The reference voltage is taken as the
whole body potential or the sum of
the three limb leads. Note their
placement with respect to the
sternum. V1 on the right side of the
sternum, V2 on the sternum, and V6
is on the axillary line.

You should now be able to see why


the R-wave in each of the Leads
shown in Fig. 4 is of small or large
magnitude and has the polarity (up or
down) that it has. To do this,
remember that the main
depolarization wave of the left
ventricle takes place (normally) in a
downward direction and from right to
left. If two leads are precisely
parallel to the wave direction, they
will record the largest possible ECG
signal, and that occurs here in Lead
II. If the wave is moving
perpendicular to the axis defined by
two leads, there will be no signal.
Accordingly, the R-wave in Lead III
is typically small. If the wave of
depolarization is moving away from
an electrode (defined as ‘positive’),
the ECG deflection will be negative.
Since the wave is moving from right
to left in the chest, the aVR
(augmented right lead) typically
shows a negative wave during
ventricular depolarization, as does
the V1 precordial lead.

Figure 4 Standard ECG recordings. Southwestern Web Curriculum

35
III. The electrophysiology of cardiac cells.

We are now ready to consider electrophysiology of the different myocyte types. The SA node cells
and the AV node cells are fundamentally different from the other cardiac cells, because their
membrane potential never reaches potentials more negative than about -65 mV, and the upstroke of
their action potentials is very slow compare to other myocytes. As we will see, the major reason is
that the SA and AV node cells lack two types of ion channels, voltage-activated Na channels and
background (inward rectifier) K channels. For all cardiac myocytes, besides SA and AV node cells,
we can generalize that they rely on voltage-activated Na channels for their upstroke, and that they
rely on K channels for repolarization. All cardiac myocytes contain voltage-activated Ca channels
that generate Ca influx upon depolarization. Differences in the shapes of repolarization and in the
length of the action potentials are mostly due to differences in the types of K channels that are
expressed. Given the range of different action potential shapes in the heart, it should not come as a
surprise that the cardiac myocytes contain many different potassium channels.

How do we know the basic roles of Na, Ca and K channels in myocytes? Experimentally, Na
channels can be blocked with the toxin, tetrodotoxin, and an action potential can still be evoked by a
strong electrical shock (see Fig. 5). The depolarization rate is slow, and these action potentials are
called 'slow response' action potentials. The upstrokes of the ‘slow responses’ are generated by
inward Ca current, which is less than 1/10th as large as the normal Na current. As you might expect,
the upstroke velocity increases when extracellular Ca is increased in experiments. We will see later
that the depolarization of S/A and A/V node cells, which is normally very slow, represents 'slow
responses' mediated by Ca channels.

Figure 5. Effect of blocking Na channels in cardiac Figure 6. Effect of blocking Ca channels in cardiac
ventricular myocytes ventricular myocytes

From the previous lecture, you know that the major Ca channels in heart are blocked by the
dihydropyridines, which belong to the general class of drugs called calcium channel blockers. When
the Ca channels are blocked, the action potential upstroke in ventricle is unchanged (see Fig. 6).
However, the plateau phase of the action potential is reduced by a several millivolts, and
repolarization begins at an earlier time. These effects define the role of Ca channels in the

36
ventricular action potential. Ca channels generate an inward current that supports the long cardiac
action potential plateau and resists repolarization. However, the cardiac action potential remains
very long after the Ca channels are blocked. The major reason that the cardiac action potential is
long compared to nerve action potentials, is that K-channels are less dense in cardiac membrane,
and the cardiac K-channels open much more slowly than in nerve.

The Cardiac Ventricular Action Potential


Ca2+ influx

0mV
K+ efflux

200ms
Na+ influx Ca2+ efflux

-85mV

Open Na/Ca exchangers


Na channels Ca efflux takes place via Na/Ca
Closed exchangers, which thereby generate an
inward current (i.e. a depolarizing
Open influence). The exchangers are simply
Ca channels responding to changes of Ca in the cell
Closed and the membrane potential.

Open
Delayed
K-channels
Closed

Open
Inward rectifier
K-channels Closed

Figure 7

Figure 7 illustrates the function of the most important current-generating systems in ventricle.

The Cardiac Ion Channels.

1. Sodium (Na) channels. All cardiac cells with a fast action potential upstroke use voltage-gated
Na channels to make the upstroke, just as in nerves. The cardiac Na channels inactivate within 1-3 ms
of upstroke. Thus, they are closed (inactive) during the action potential plateau, and they recover
from inactivation just after repolarization.

2. Potassium (K) channels. Like neurons, cardiac myocytes contain multiple K channel types. We
can explain reasonably the ventricular action potential with just two K channel types, ‘delayed’ (or
voltage-gated) and ‘inward rectifier’ K channels (or ‘background K channels).

a. ‘Delayed’ K channels are the voltage-gated K channels that bring about cardiac repolarization.
The major difference to K channels in axons is that cardiac channels respond to voltage changes

37
more slowly. They open slowly during prolonged depolarizations (100 to 300 ms) and close slowly
during and after repolarization.

b. ‘THE’ inward rectifier = ‘background K channel’. These cardiac K channels belong to the
general class of "inward rectifiers" that we have introduced in previous lectures. They are NOT
gated by typical voltage sensors. Unfortunately, the same word (‘inward rectifier’) designates these
specific channels AND designates the entire class of K channels that are not gated by voltage-
sensors. ‘The’ cardiac inward rectifier channels are open at negative potentials, they become
blocked (closed) almost instantly during depolarization, and they reopen during repolarization at
about -60 mV. Because they are open only at negative potentials, they stabilize membrane potential
at a negative potential close to the Nernst potential for potassium. When small perturbations occur
(e.g. by opening of a few Na channels), ‘the’ inward rectifier K channels bring potential back to a
more negative value and prevent the firing of an action potential. These channels account for the
high "resting K conductance" of most cardiac cells. They prevent MOST cardiac cells from
depolarizing spontaneously in the diastolic period.

Subsequently, we will have to make this picture of cardiac K channels a little more sophisticated.

3. Calcium channels. Cardiac Ca channels open within one to three milliseconds upon
depolarization. They allow Ca to enter the myocytes. The cell membrane must be depolarized to
around -50 mV for the Ca channels to open, whereas Na channels open already at around -70 mV.
The peak Ca inward current is more than 10-times less than the peak Na current. The Ca channels
inactivate much more slowly than Na channels, and some Ca channels remain open throughout the
action potential plateau. This is why the major cardiac Ca channels are called ‘L-type’ (L=Long),
and sometimes they are called ‘'slow calcium channels’. Because some Ca channels are open
throughout the plateau, total Ca influx at each beat depends on the duration of the action potential.

Although Ca channels typically open at more positive potentials than Na channels, there are some
Ca channels in myocytes that open around the same potentials as Na channels, and they inactivate
very quickly (more like Na channels). These are called ‘T-type’ (Transient) Ca channels.

Not all Ca channels open at each depolarization...some of them are effectively ‘sleeping.’ We will
see in the next lecture that Ca channels can be "up-regulated" (awoken from sleep!) when cAMP
increases in response to catecholamines. Besides staying open for longer time, the L-type Ca
channels are regulated differently from the Ca channels neuronal (N-type) channels. For example,
the N(euronal) Ca channels are NOT activated by an increase of cAMP. More on that in the next
lecture…

4. Na/Ca exchange. BEWARE! See Fig. 8. In both atrium and ventricle, the extrusion of Ca
generates inward current because three Na ions are transported into the cardiac cell for each Ca
which is extruded. This inward current is normally not very important, but it can become important
in pathological settings. When cardiac cells are metabolically stressed, they tend to get 'overloaded'
with Ca. Sometimes, in this setting, Ca is released by Ca-induced Ca release WITHOUT a
depolarization.

38
The Cardiac Na/Ca Exchanger Can Generate Arrhythmias

Membrane potential
Figure 8. Spontaneous
Ca release from the SR
can cause ‘extra’ action
[Ca2+] i
potentials by activating
Regenerative activation
3 of Na-channels
Na/Ca exchangers to
extrude Ca in exchange
for extracellular Na.
2
1 Na/Ca exchanger generates
Spontaneous 'after-depolarization'
SR Ca-release

Typically, this happens just after repolarization, which is a dangerous time in the cardiac cycle. It is
dangerous because Na and Ca channels are just becoming available to make another depolarization
(see Fig. 7). When there is a spontaneous release of Ca, the exchanger extrudes Ca and generates an
inward current. This causes a small depolarization, called an 'after-depolarization, and another action
potential may be fired if the threshold for Na channels is reached!

Now, we need to get slightly more sophisticated about K-channels…

Overview of cardiac potassium channels.


Voltage-gated K-channels.

1. The early (‘transient’) outward current. As described in Fig. 1, the Purkinje cells have a very
fast upstroke followed by a rapid initial repolarization called ‘phase 1’. Ventricular myocytes also
can display a phase 1 repolarization phase, as is the case in Fig. 7. This initial repolarization is
brought about by potassium channels that open very quickly on depolarization and then inactivate.
2. The delayed outward current. The ‘delayed’ potassium current is actually generated by two
types of voltage-gated potassium currents. As reiterated below, this is important because some of
the common genetic diseases that predispose the heart to developing fatal arrhythmias are caused by
mutations to one of those channels. The mutated channels are called ‘HERG’ channels.
Inward rectifier K-channels.
1. ‘The’ inward rectifier channels. As described above, these channels stabilize the resting
membrane potential in all cardiac cell types, EXCEPT the SA node and the AV node cells.
2. K-ATP channels. These are channels from the inward rectifier class that sense the metabolism
of cells and open when ATP levels fall and ADP levels rise. Cardiac muscle has a high density of
these channels, and they can drastically shorten the cardiac action potential when the heart becomes
ischemic.
3. Acetylcholine-activated K channels. These are inward rectifier channels found only in atrial
muscle. NOT in ventricle. They are activated by acetylcholine, released physiologically by the
vagus nerve. Their activity determines the heart rate, as will be discussed in the next lecture.

39
Potassium channel case #1: The delayed K current in heart is generated by two types of K
channels. One opens more quickly than the other. The second potassium channel type, called
‘HERG’, plays a critical role as the main repolarization phase gets going. HERG mutations can
cause the cardiac action potential to be very long. These mutations (like those in CFTR) cause the
channels to be degraded before being inserted into the surface membrane. The long action potentials
predispose the heart to generate arrhythmias that can lead to sudden death. You can see in Fig. 9
why this disease is called ‘long QT syndrome’. Long action potentials increase the QT period.

Long QT Syndrome
R R
0 mV

late G K P T P T
or Normal
Q S QS
late G N a
0.3 s
R R
T
0.3 s Long QT P
syndrome P
Loss of ‘Herg’ K channels or impaired Q S Q S
Na channel inactivation both lead to long
ventricular action potentials .

Figure 9. Action potentials and ECG with prolonged action potentials. Figure by Hilgemann,
2006.

In addition to mutations in HERG potassium channels, long QT syndrome can be caused by


mutations in Na channels that lead to incomplete Na channel inactivation. When Na channel
inactivation is delayed, the Na channels generate inward current for a prolonged time, thereby
lengthening the action potential, similar to the effect of HERG loss.

Potassium channel case #2: Surprisingly, the most numerous K channel in cardiac myocytes is
one that has no influence in normal physiological conditions. These are the ‘K-ATP’ channels,
which were introduced in a previous lecture. They are a member of the inward rectifier family.
These channels bind both ATP and ADP. ATP binding causes them to close, and ADP binding
promotes opening. When the metabolism of the heart is compromised, as occurs in ischemia, these
channels open and can cause massive action potential shortening. When blood flow is limited and
metabolism is compromised, extracellular K will tend to accumulate in the extracellular space. The
Nernst potential for K then becomes more positive, and resting potential increases. Thus, the
opening of these channels can ultimately promote depolarization as extracellular K rises. There will
be a significant ‘injury current’ flowing, whenever one part of the heart is at a potential that is
different from the rest of the heart. As a result, the ST-segment of the ECG can be changed
dramatically in multiple ways. Elevation of the ST segment is an outcome that is of great diagnostic
importance in myocardial infarction (see Fig. 10).

40
Figure 10. Left. Typical changes of cardiac action potentials during ischemia. Right. ECG changes associated with
myocardial infarction. Figures by Hilgemann, 2006

Potassium channel case #3. We have now seen how changes of the activity of several ion channel
types affect the ventricular cardiac action potential. Figure 11 shows one final example, namely the
effect of blocking the inward rectifier K channels. In this experiment, ventricular myocytes are at
first stimulated with electrical shocks to evoke action potentials. If the myocyte is not shocked,
there is no action potential.

Control 0.5 s

0mV

Stimulation

+ 1 mM barium to block inward rectifier K-channels

0mV

Spontaneous electrical activity

Figure 11. Effect of blocking inward rectifier (background)


K channels. Fig by Hilgemann, 2006

The inward rectifier channels are known to be blocked specifically by the divalent cation, barium.
You do not need to remember this fact, but the effect of blocking inward rectifier potassium
channels is very revealing. Block of the inward rectifier channels turns the great majority of cardiac
myocytes into pacemakers. The spontaneous activity occurring in this kind of experiment is very
much like the repetitive action potential firing discussed for neurons. Between action potentials, the
‘background’ cation conductance of the membrane causes depolarization until the threshold for Na
channels is reached. Then, an upstroke occurs, K channels respond to depolarization by opening
slowly. K channels repolarize the myocyte, and they then close again between beats. As they close,
they allow the cation conductance to gain influence and depolarize the cell again.

41
The take-home message from this example is that all cardiac cells can potentially initiate electrical
activity and that ‘the’ inward rectifiers are critical to stabilize cardiac electrical activity and suppress
arrythmogenic activities.

Cardiac SA node pacemaker activity.

Cardiac Sinus Node Pacemaking


K - channels open
and repolarize
3
Ca - c
channels
hannels 2
Ca -channels
open and 0 3 close
Sinus node
depolarize
Pacemaker
cycle
Non - specific 4 4 K - channels
channels close
depolarize

2
Membrane
Potential 0
3
4 4

Channels Ca
K
Open
Non - specific
Closed

Figure 12. Sinus Node Pacemaking. Action potential phases


are numbered as in Fig. 1. There is no phase ‘1’, which is the early
rapid repolarization in Purkinje cells and some ventricular
myocytes. Sinus node pacemaking is an interplay of delayed K-
channels, Ca channels, and nonselective cation channels.
Fig. by Hilgemann, 2006

SA and AV node cells lack both inward rectifier channels and Na channels. Cardiac pacemaking in
SA and AV node cells can be explained largely by the interplay of delayed K channels, Ca channels,
and nonspecific 'background' cation channels. The basic cycle of depolarization, upstroke and
repolarization is shown in Fig. 12. Take a close look.

Since K conductance is very small at negative potentials, small (nonspecific) inward currents,
sometimes called 'leak' currents, can depolarize the cells easily. The depolarizing influence becomes

42
very large when voltage-activated K channels turn off. There are no inward rectifiers to keep resting
potential stable. The non-specific channels depolarize cells until around -50 mV, when the voltage-
gated Ca channels begin to open in a positive-feedback fashion; more depolarization leads to
activation of more Ca channels; Ca current depolarizes and activates more Ca channels in a positive
feedback loop. Thus, SA and AV node cells depolarize by Ca-dependent 'slow responses.' After
depolarization, "delayed" K channels open and repolarize, and then they begin to close. Please note:
That is after repolarization. Depolarization again begins as the K channels close (deactivate), and
inward currents carried by nonspecific channels come to dominate the outward currents. This
process repeats itself again and again, for a lifetime.

Non-specific current called "F" current (IF). Nonselective cation channels have been identified
in S/A node cells, as well as neurons, which may provide the background cation current that
promotes pacemaking. They are called "F" channels. They are also called 'H' channels because they
are activated by hyperpolarization (i.e. by repolarization). They close slowly upon depolarization.
When these channels are open during diastole, they contribute to the pacemaker depolarization. We
will see in the next lecture that these channels are activated by sympathetic stimulation, thereby
generating a larger background inward current and speeding up the cardiac cycle.

As we have seen, all myocytes are potentially pacemakers. We make three distinctions.

A true pacemaker is a cell that in normal hearts generates the action potential leading to a heart
beat. Normally, SA node cells comprise the true pacemaker.

A latent pacemaker is any cardiac cell with the capability of automaticity that is located outside the
SA node. In the normal heart, latent pacemakers are not allowed to spontaneously generate action
potentials because the cells are preemptively excited by propagating AP’s that are generated in the
SA node. The most important latent pacemakers are the AV node cells and the Purkinje cells. These
pacemakers lead the heart into an EC-coupling cycle when activity from the sinus node is disrupted,
as occurs when conduction from the atria is blocked (or when the atria are in fibrillation).

An ectopic pacemaker is an automatic cell outside the SA node that spontaneously generates an AP
leading to one or more beats, as described in connection with the Na/Ca exchanger. A range of
physiological and pathological processes can activate latent pacemaking. We will see in the next
lecture why catecholamines strongly promote ectopic pacemaking.

43
Figure 13. Sinus Node Pacemaking. Figure by Hilgemann, 2006

A Graphic Summary. Fig. 13 shows all of the ionic currents that occur in cardiac pacemaker
cells during spontaneous activity. Remember that inward current is plotted downward, outward
current upward. We will use this summary for discussion in lecture as time permits. And I will
prepare a detailed explanation as requested. For now, give a try to explain why the current wave
forms appear as they do! It’s fun.

44
CHAPTER 4. AUTONOMIC CONTROL OF THE HEART

This chapter gives an overview of the control of cardiac function by the autonomic nervous
system.

Key Words & Concepts


1. Innervation patterns of the heart by the parasympathetic and sympathetic systems
2. ‘inotropic’, ‘dromotropic’, and ‘chronotropic’.
3. Chrontropic, inotropic and dromotropic effects of catecholamines on the heart
4. Chronotropic, iontropic, and dromotropic effects of acetylcholine on the heart
5. The targets and effects of cAMP-activated protein kinase in heart.
6. The major effects of catecholamines and acetylcholine on ion channels in heart and the
resulting changes of ventricular action potentials and sinus node pacemaking.

The Vocabulary of Cardiac Physiology

Three terms describe the most important effects of neurotransmitters and drugs on the heart.

• Inotropic refers to contraction strength. An increase of contraction strength is a positive


inotropic effect.
• Chronotropic refers to heart rate. An increase of heart rate is a positive chronotropic effect.
• Dromotropic refers to conduction rate. An increase of conduction velocity is a positive
dromotropic effect.

Most cardiovascular physiologists refer to the ‘contractile state’ or ‘contractility’ of the heart.
‘Contractility’ is the ability of the heart to contract, independent of preload and afterload. We start
therefore with a review of cardiac excitation-contraction coupling.

45
Figure 1. Cardiac excitation-contraction coupling Figure by SouthwesternWedCurriculum

I. Review of cardiac excitation-contraction coupling.

● Cardiac action potentials propagate into t-tubules that run along the Z-lines of each sarcomer.
● Voltage-gated Ca channels open upon depolarization (by Na channels) and let significant amounts
of Ca into myocytes. Also, the Na/Ca exchanger can generate Ca influx during the initial rise of
cytoplasmic Ca.
● A local (‘subsarcolemmal’) increase of cytoplasmic Ca causes ‘Ryanodine (RyR) Receptors’ (=
SR Ca release channels) to open and release Ca.
● Cytoplasmic Ca does NOT maximally activate myofillaments in heart. Thus, changes of the
magnitudes of Ca transient can have a large effect on the degree of activation.
● Ca is taken up by the SR and extruded by the Na/Ca exchanger to bring about relaxation. A small
amount of Ca can be extruded by surface membrane Ca pumps (as in all cells).
●NEW TERRITORY: Phospholamban (PLB) is a regulatory protein that interacts with the SR Ca
pumps. Phosholamban inhibits the Ca pumps when it is NOT phosphorylated. And phosphorylation
of phospholamban by protein kinases relieves the inhibition (i.e. activates SR Ca pumping). By this
mechanism, the SR Ca pump is strongly activated by cAMP-dependent protein kinase.
●The SR Ca release channels (RyR) remain inactivated after each release cycle, and recover their
ability to release Ca between beats. This mechanism prevents tetanus contraction in heart.
●NEW TERRITORY: The cytoplasmic Ca transient indirectly controls metabolism of the
mitochondria. At each beat, a few percent of the cytoplasmic Ca is taken up into mitochondria by a
selective Ca transporter, a uniporter. In the mitochondria, oxidative phosphorylation (i.e. generation
of ATP) is stimulated by the rise of Ca. Between beats Ca is extruded back into the cytoplasm by a
mitochondrial Na/Ca exchanger that cytoplasmic Na for mitochondrial Ca.

46
II. Regulation of cardiac excitation-contraction coupling by autonomic neurotransmitters.

The major extrinsic (external) control of the heart comes from the autonomic nervous system. The
sympathetic system innervates the entire heart richly. Within the heart, norepinephrine is the major
catecholamine released by sympathetic nerve endings; epinephrine is the major catecholamine that
circulates in the blood, after being released from the adrenal medulla. The parasympathetic system
(i.e. fibers from the vagus nerve) innervates the atria richly, but the ventricles only weakly. The
vagus releases acetylcholine.

Figure 2. Autonomic control of the heart. Figure by SouthwesternWedCurriculum

Sympathetic system: norepinephrine (=noradrenaline) and epinephrine (=adrenaline).


Catecholamines can act strongly on the heart in just 1 to a few seconds. These fast effects are
mediated mostly via the cAMP system. These neurotransmitters bind to β receptors of the heart.
The receptors then activate a G-protein (Gs) that in turn activates the adenylate cyclase. The
adenylate cyclase produces cyclic AMP from ATP, and cAMP activates protein kinases called 'A-
kinases.' The A-kinases phosphorylate multiple specific targets in heart, and this phosphorylation
changes their function. Those of you who have been reading pharmacology will know that
norepinephrine, in general, activates α adrenergic receptors better than β receptors. However, the
β receptors of heart are a special sub-type, called β1 receptors, and they are activated effectively by
norepinephrine. Adrenergic receptors of the ‘alpha’ family also exist in heart. They become
important in the long-term regulation of the heart, in particular ini4 the control of cardiac cell
growth as occurs in cardiac hypertrophy.

47
Figure 3

What happens to the heart in the presence of catecholamines? (1) There is a strong positive
chrotropic effect (i.e. increase of heart rate), and we will discuss the mechanism in a moment. (2)
There is a strong positive inotropic effect (thus, the heart contracts more forcefully). (3) The heart
relaxes more quickly. If the heart didn't relax more quickly, it would not be possible for the heart to
beat at high frequencies in the range of 180 beats per minute. (4) There is a positive dromotropic
effect (i.e. increase of conduction rate through the pacemaker system). (5) The plateau of the action
potential is increased by a few millivolts. And (6), the duration of the action potential is decreased.
Thus, the action potential in the presence of catecholamines crosses over the action potential without
catecholamines, when they are plotted on top of each other (see Figure 4).

How do these effects come about? 1) Ca channels are phosphorylated by A-kinase, and as a result
more channels open with each depolarization. The increased inward Ca current makes the plateau of
the action potential higher (more depolarized). 2) As already mentioned, the SR protein,
phospholamban, is phosphorylated, and the SR Ca pump is stimulated to take up Ca more quickly
and accelerate relaxation. 3) The increased pump activity tends to keep the SR loaded with Ca, and
therewith increase the amount of Ca released. 3) Troponin (on the thin filaments) is also
phosphorylated. This results in a faster dissociation of Ca from troponin which supports the faster
relaxation. 4) In ventricles and in ‘working’ atria, some of the delayed K-channels are
phosphorylated by A-kinase. As a result, their open probability is increased, and they cause a faster
repolarization and therewith action potential shortening.

48
Figure 4

High frequencies of contraction, and high rates of contraction and relaxation, cost a lot of ATP.
Thus, the overall effects of catecholamines are very energy-demanding. It will not be surprising,
then, that patients with compromised coronary circulation often experience problems when they
release catecholamines. Also, it will not be surprising why drugs that block the beta receptors ('beta
blockers') are a mainstay of therapy for such patients.

Figure 5

49
Compare in Fig.5 the effect of catecholamines and the effect of the heart glycoside, ouabain
(=digitalis) on the contraction of an isolated cardiac muscle. Heart glycosides are used (sometimes)
to treat cardiac insufficiency. These drugs increase the cardiac contraction almost symmetrically, i.e.
with only a small increase of the relaxation rate. This is energetically much less demanding for the
heart than to increase dramatically the rate of contraction and relaxation, as with norepinephrine.
Also, ouabain does not increase the frequency of the heart when it is given to patients. In fact, the
frequency decreases, mostly due to a vagal reflex in response to increased cardiac output.

Practice exam question: How do heart glycosides act on the heart? The textbook explanation is
as follows: A small fraction of the Na/K pumps are inhibited, cytoplasmic Na rises a few
millimolar, and the ability of the Na/Ca exchanger to remove Ca from the heart cells is
decreased. Thus, Ca is retained in the SR, and more Ca is released at each beat.

Parasympathetic system: acetylcholine.

Acetylcholine acts on the heart by two mechanisms: (1) Specific K channels are opened in the atria,
and the result is a decrease of heart rate (=negative chronotropic effect) and a decrease of atrial
contraction strength (=negative inotropic effect). These channels do NOT exist in ventricle. (2)
Acetylcholine can antagonize the effects of catecholamine everywhere in the heart.

As indicated in Fig. 6, acetylcholine binds to muscarinic receptors throughout the heart, although
vagal innervation of the ventricles is very weak. The muscarinic receptors are coupled to adenylate
cyclase by so-called 'Gi' G-proteins. These G-proteins, when activated, block the activation of
adenylate cyclase by catecholamines through the 'Gs' G-proteins. Thus, acetylcholine is a
'functional antagonist' of catecholamines everywhere in the heart.

Ach-activated K channels, which exist throughout the atria, belong to the class of 'inward rectifier' K
channels. When acetylcholine binds to muscarinic receptors, a G-protein is activated and
subsequently acts to open the K channels. This G-protein, like other G-proteins, has three subunits.
When it is activated, it dissociates into alpha (a) and beta/gamma (b/g) subunits. The beta/gamma
subunits bind to and activate the K channels. These channels are called GIRK because they are G-
protein-activated Inward Rectifier K-channels. We consider the electrical details in the next lecture.

50
Figure 6

Local metabolic factors

In addition to the autonomic nervous system, the metabolic state of heart cells strongly influences
their function. In general, cardiac contraction strength decreases when the metabolic state of the
heart is compromised. There are several reasons. The sensitivity of the contractile proteins to Ca
can decrease when pH falls, K currents are in general increased (to shorten the action potential), and
Ca channels can be down- regulated. Acidosis and local changes of ATP concentrations have big
effects on many of the E/C coupling elements. Adenosine is released by heart cells to the
extracellular space during ischemia, and adenosine decreases heart rate. Also, extracellular
adenosine induces a strong vasodilation of coronary arteries. You will hear more about the
regulation of cardiac circulation in later lectures.

51
III. Neurotransmitter Effects on Pacemaking

Now, it's easy to understand how neurotransmitters change the frequency of pacemaking, as shown
in Fig. 7.

Modulation of Cardiac Pacemaking by Neurotransmitters


Acetylcholine activates K-channels that cause hyperpolarization
and thereby lengthen the time to reach threshold.
Control +Acetylcholine
SA-node
membrane
potential

400ms

Norepinephrine increases the open probability of Ca-channels that


cause depolarization. More inward current means a shorter time to
reach threshold, as well as an increase of the depolarization velocity.
+Norepinephrine
Control

SA-node
membrane
potential

400ms
Figure 7

(1) Acetylcholine activates a specific set of inward rectifier K channels in atria, as already
discussed. These channels cause the membrane potential to become more negative (hyperpolarized)
at the end of each action potential. It then takes longer for the nonspecific cation channels to
depolarize the membrane to the threshold for another action potential. Thus, the diastolic interval
becomes longer....and we have explained the "negative chronotropic effect" of acetylcholine. When
acetylcholine release is massive, as can occur in boxers after a punch to the stomach, the heart can
stop completely for many seconds...until the acetylcholine is hydrolyzed and the hyperpolarization
by the Ach-activated channels is reversed.

(2) Catecholamines (i.e. norepinephrine and epinephrine). Regulation of heart rate must be
robust and reliable for a life-time. Therefore, two mechanisms are used to increase heart rate. In the
sinus node cells, as in all other cardiac cells, an increase of cAMP up-regulates Ca current. The Ca
channels open more easily with depolarization. This leads to a faster and earlier positive feedback to
activate more Ca channels and start the upstroke. The result is a faster heart rate ("positive
chronotropic effect). Second, an increase of cAMP in the sinus atrial node cells directly activates
nonselective cation channels to generate more ‘background’ inward current, thereby also tending to

52
increase the rate of diastolic depolarization. These are called ‘F’ or sometimes ‘H’ channels. There
is even a third mechanism to increase the heart rate. Since more Ca enters the node cells at each beat,
more Ca must also be extruded by Na/Ca exchange. Ca extrusion by 3Na/1Ca exchange makes
inward current that promotes depolarization between beats. In the S/A node, as elsewhere,
acetylcholine can reverse all the actions of catecholamines, mediated by an increase of cAMP, by
effectively blocking the activation of adenylate cyclase.

IV. Other Effects of Neurotransmitters in the Heart

Acetylcholine Effects in ‘Working Atrium’

+40 Without
Without
With With
acetylcholine
acetylcholine
mV

100ms

-80

In 'working' atrium, acetylcholi ne activates specific K-channels that


cause a faster repolarization. Shortening of the action potential
'deactivates' Ca channels, and contraction force decre ases because
Ca influx is decreased.

Figure 8

Acetylcholine effects in 'working atrium.' Acetylcholine activates K channels throughout the atria.
As shown in Fig. 8, this leads to hyper-polarization of the atrial cells by a few millivolts. Also, it
leads to shorter action potentials in the atrium, and the shorter action potentials lead to a shortened
period of Ca influx. That is because Ca channels close immediately when the atrial cells repolarize.
The reduced Ca influx leads to a negative inotropic effect of acetylcholine in atrium.

What is carbachol?

Carbachol is a drug that mimics the effects of Ach at muscarinic receptors, but it is not broken
down (i.e. hydrolyzed) by esterases. Carbachol was used for many years to stop dangerous atrial
arrhythmias because of its strong hyperpolarizing effect on the atrium. Today, better drugs are
available. When administered I.V. it causes urination and sweating. Why?

53
Effects in the Ventricle

+40 With and without +4 0 Without acetylcholine


acetylcholine With ac etylcholine

mV mV

-80 -80
Without acety lcholine
With acetylcholine

Without norepinephrine With norepinephrine

In the ventricle, acetylcholine has no effect in the absence of catecholamines,


but it can reverse (partially) the catecholamine effects.

Figure 9

You know already that acetylcholine has almost no effect in the ventricle in the absence of
stimulation by catecholamines. However, acetylcholine can reverse for the most part the effects of
catecholamines in ventricle. This includes, as illustrated in Fig. 9, all effects of catecholamines on
the ventricular action potential and contraction.

54
V. Summary and Review
As a review, you should be able to describe all major effects of catecholamines in heart.

1. Catecholamines.
a. Increase in the heart rate
b. Increase in contractility
c. Increased relaxation rate
d. Accelerated AV conduction
e. Faster repolarization

2. Summary of direct acetylcholine actions.


a. Decrease in heart rate.
b. Decrease in atrial contraction
c. Slower AV conduction rate.
Hutter/Trautwien

Autonomic NS controls cardiac pacemaking by


modulating many ion channels via G-protein signals
Sympathetic nerves (noradrenaline, beta receptors)

-20

-40

-60
0 5s

Vagus nerve (acetylcholine, muscarinic receptors)

0 3s
Hutter & Trautwein
Figure 10. Original recordings from 1956 of the effects of activating the sympathetic and
parasympathetic inputs to the heart on sinus pacemaking. Note that acetylcholine acts faster and its
effect reverses faster than catecholamines. Try to example the details of these recordings!

55
You should now be able to describe the fundamental differences between cardiac and skeletal
muscle.

Similarities of Cardiac Muscle to Skeletal Muscle

1. Cardiac muscle is organized in sarcomer units and is thus striated.


2. Cardiac muscle has transverse tubules ("T-tubes").
3. Cardiac muscle has sarcoplasmic reticulum ("SR").
4. Contraction is preceded by an action potential and a cytoplasmic Ca transient.
5. The heart contracts more forcefully when stretched.
6. The heart contracts more rapidly when unloaded (force-velocity relation).

Differences between Cardiac and Skeletal Muscle

1. In heart, action potentials propagate from cell to cell via gap junctions.
2. Cardiac action potentials are very long (100 to 300 ms).
3. In heart, action potentials originate spontaneously.
4. The heart requires extracellular Ca to contract.
5. In heart, contractions cannot summate to a tetanus.
6. The heart is regulated by the autonomic nervous system.
7. In heart, contraction is regulated by changing the amount of Ca released.
8. In heart, Ca release from SR is activated by Ca influx.
9. Cardiac cells cannot be stretched beyond overlap of thick and thin filaments.

Regulation of Cardiac Contraction Force and Frequency

1. The 'Frank-Starling' mechanism up-regulates contraction force when the filling volume of the
heart increases.
2. The sympathetic system up-regulates force, speed, and frequency of contraction.
3. The parasympathetic system down-regulates frequency of contraction and can reduce all effects of
the sympathetic system by preventing activation of adenylate cyclase.

56
VI. AFTERWORD

Who was Otto Loewi?

Otto Loewi (June 3, 1873 – December 25, 1961) was a


German-American pharmacologist. His discovery of
acetylcholine helped in enhancing medical therapy and
personally earned for him the 1936 Nobel Prize in
Physiology or Medicine which he shared with Sir Henry
Dale. The idea that chemicals are released by neurons to
communicate to other cells was validated first in the
heart.

Figure 10. How the ‘Vagusstoff’ was discovered. Figures from the Wilkepedia.

In his most famous experiment, Loewi took fluid from one frog heart and applied it to another,
slowing the second heart and showing that synaptic signaling used chemical messengers. Before
Loewi's experiments, it was unclear whether signaling across the synapse was bioelectrical or
chemical. Loewi's famous experiment, published in 1921, largely answered this question. According
to Loewi, the idea for his key experiment came to him in his sleep. He dissected out of frogs two
beating hearts: one with the vagus nerve which controls heart rate attached, the other heart on its
own. Both hearts were bathed in a saline solution (i.e. Ringer's solution). By electrically stimulating
the vagus nerve, Loewi made the first heart beat slower. Then, Loewi took some of the liquid
bathing the first heart and applied it to the second heart. The application of the liquid made the

57
second heart also beat slower, proving that some soluble chemical released by the vagus nerve was
controlling the heart rate. He called the unknown chemical Vagusstoff. It was later found that this
chemical corresponded to acetylcholine.

Loewi's investigations “On an augmentation of adrenaline release by cocaine” and “On the
connection between digitalis and the action of calcium” were profound concepts and were studied
relentlessly by others decades later.

He also clarified two mechanisms of eminent therapeutic importance: the blockade and the
augmentation of nerve action by certain drugs.

58

You might also like