You are on page 1of 272

ADVANCED

EXERCISE
ENDOCRINOLOGY

Katarina T. Borer, PhD


University of Michigan

Human Kinetics
Library of Congress Cataloging-in-Publication Data
Borer, Katarina T.
Advanced exercise endocrinology / Katarina T. Borer.
p. ; cm. -- (Advanced exercise physiology series)
Includes bibliographical references and index.
I. Title. II. Series: Advanced exercise physiology series.
[DNLM: 1. Endocrine System--physiology. 2. Physical Exertion--physiology. WK 102]
612.4--dc23
2012031085
ISBN-10: 0-7360-7516-X (print)
ISBN-13: 978-0-7360-7516-9 (print)
Copyright © 2013 by Katarina T. Borer
All rights reserved. Except for use in a review, the reproduction or utilization of this work in any
form or by any electronic, mechanical, or other means, now known or hereafter invented, including
xerography, photocopying, and recording, and in any information storage and retrieval system, is
forbidden without the written permission of the publisher.
Acquisitions Editor: Amy N. Tocco; Developmental Editor: Amanda S. Ewing; Assistant Editors:
Kali Cox, Derek Campbell, Susan D. Huls, and Casey A. Gentis; Copyeditor: Amanda M. Eastin-
Allen; Proofreader: Red Inc.; Indexer: Andrea J. Hepner; Permissions Manager: Dalene Reeder;
Graphic Designers: Joe Buck and Fred Starbird; Graphic Artist: Tara Welsch; Cover Designer:
Robert Reuther; Photo Production Manager: Jason Allen; Art Manager: Kelly Hendren; Associate
Art Manager: Alan L. Wilborn; Illustrations: © Human Kinetics, unless otherwise noted; Printer:
Edwards Brothers Malloy
Printed in the United States of America 10 9 8 7 6 5 4 3 2 1
The paper in this book is certified under a sustainable forestry program.
Human Kinetics
Website: www.HumanKinetics.com
United States: Human Kinetics Australia: Human Kinetics
P.O. Box 5076 57A Price Avenue
Champaign, IL 61825-5076 Lower Mitcham, South Australia 5062
800-747-4457 08 8372 0999
e-mail: humank@hkusa.com e-mail: info@hkaustralia.com
Canada: Human Kinetics New Zealand: Human Kinetics
475 Devonshire Road Unit 100 P.O. Box 80
Windsor, ON N8Y 2L5 Torrens Park, South Australia 5062
800-465-7301 (in Canada only) 0800 222 062
e-mail: info@hkcanada.com e-mail: info@hknewzealand.com
Europe: Human Kinetics
107 Bradford Road
Stanningley
Leeds LS28 6AT, United Kingdom
+44 (0) 113 255 5665
e-mail: hk@hkeurope.com E4517
Contents
Preface v
Series Preface vii

Chapter 1 Unique Properties of Endocrine and Autonomic


Messengers . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Functions of the Endocrine System and Autonomic
Nervous System. . . . . . . . . . . . . . . . . . . . . 2
Classification of Hormones . . . . . . . . . . . . . . . . . 5
Chemical Structure of Hormones . . . . . . . . . . . . . . . 6
Hormone Release and Transport . . . . . . . . . . . . . . .15
Hormone Receptors and Hormone–Receptor Message
Transduction . . . . . . . . . . . . . . . . . . . . . .18
Properties of Hormone–Receptor Interactions. . . . . . . . . 26
Summary . . . . . . . . . . . . . . . . . . . . . . . 29

Chapter 2 Activation of Nonhormonal Signaling During Exercise . . . . . 31


Signal Transduction by Neuronal and Electrostatic Events
in Exercise . . . . . . . . . . . . . . . . . . . . . . .32
Intracellular Calcium Release as a Trigger of Signal
Transduction . . . . . . . . . . . . . . . . . . . . . 33
Signal Transduction of Mechanical Strain, Vibration,
and Fluid Shear . . . . . . . . . . . . . . . . . . . . 36
Signaling in Response to Sensing Energy Need . . . . . . . . 44
Free Radicals as Initiators of Message Transduction . . . . . . 46
Summary . . . . . . . . . . . . . . . . . . . . . . . 53

Chapter 3 Autonomic and Hormonal Control of the Cardiorespiratory


System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
Afferent Signals to the ANS . . . . . . . . . . . . . . . . 56
Integrative Brain Circuits and Efferent Components
of ANS Reflexes . . . . . . . . . . . . . . . . . . . . 58
Control of Cardiorespiratory Function by ANS and Hormones . . . 69
Summary . . . . . . . . . . . . . . . . . . . . . . . 77

Chapter 4 Body Fluid Balance . . . . . . . . . . . . . . . . . . . . . . . 79


Increases in Body Heat Load During Exercise . . . . . . . . . 80
Thermoregulatory Changes in Body Water During Exercise . . . . 80
Consequences of Fluid Loss Through Sweating . . . . . . . . 83
Cessation of Renal Reabsorptive Function During Exercise . . . . .87
Thirst and Na Hunger After Exercise. . . . . . . . . . . . . 90
Hyperhydration and Hyponatremia . . . . . . . . . . . . . .91
Strategies for Fluid Management in Exercise . . . . . . . . . 92
Summary . . . . . . . . . . . . . . . . . . . . . . . 95

iii
iv Contents

Chapter 5 Hormones and Fuel Use in Exercise . . . . . . . . . . . . . . . 97


Hormonal Mediation of Energy Balance . . . . . . . . . . . .97
Hormones in Fuel Mobilization and Utilization During
Aerobic Exercise . . . . . . . . . . . . . . . . . . . 109
Hormones in Fuel Mobilization and Utilization During
Anaerobic Exercise or RE . . . . . . . . . . . . . . . . 118
Summary . . . . . . . . . . . . . . . . . . . . . . . 124

Chapter 6 Hormonal Control of Energy Expenditure and Intake . . . . . 127


Energy-Regulating Mechanism . . . . . . . . . . . . . . . 128
Descriptors and Controls of Feeding Behavior . . . . . . . . . 134
Role of Physical Activity in the Regulation of Body Mass . . . . 137
Regulation of Body Mass Through Homeostatic Versus
Nonhomeostatic Controls . . . . . . . . . . . . . . . . 139
Body Mass Regulatory Mechanism Versus Obesity
and Weight Loss . . . . . . . . . . . . . . . . . . . 145
Summary . . . . . . . . . . . . . . . . . . . . . . . 148

Chapter 7 Exercise and Reproductive Hormones . . . . . . . . . . . . 149


Development of Phenotypic Sexual Dimorphism . . . . . . . . 149
Effects of Exercise on Sex Hormone Secretion. . . . . . . . . 161
Effects of Sex Hormones on Physical Performance . . . . . . . 170
Summary . . . . . . . . . . . . . . . . . . . . . . . 174

Chapter 8 Hormonal Mediation in Training Adaptations . . . . . . . . . 177


Systemic Hormones in Adaptations to Endurance Training . . . . 178
Role of Systemic Anabolic Hormones in Adaptations
to Resistance Training . . . . . . . . . . . . . . . . . 184
Nutritional Modulation of Hormonal Adaptations
to Exercise Training . . . . . . . . . . . . . . . . . . 192
Summary . . . . . . . . . . . . . . . . . . . . . . . 199

Chapter 9 Exercise and Endocrine Rhythms . . . . . . . . . . . . . . . 201


Basics of Biological Rhythm Physiology and Terminology . . . . 201
Synchronization of Circadian Rhythms by the
Suprachiasmatic Nucleus . . . . . . . . . . . . . . . . 205
Control of Biological Rhythms by the Food-Entrainable
Oscillator . . . . . . . . . . . . . . . . . . . . . . 209
Control of Ultradian Rhythms . . . . . . . . . . . . . . . 209
Control of Biological Rhythms by Exercise . . . . . . . . . . 211
Summary . . . . . . . . . . . . . . . . . . . . . . . 218

Chapter 10 Measuring Hormones . . . . . . . . . . . . . . . . . . . . . 219


Measuring Hormone Concentration . . . . . . . . . . . . . 219
Measuring Biological Actions of Hormones . . . . . . . . . . 224
Measuring Hormone Synthesis and Site of Production . . . . . 225
Measuring the Rate of Hormone Secretion . . . . . . . . . . 225
Summary . . . . . . . . . . . . . . . . . . . . . . . 228

References 229
Index 257
About the Author 264
Preface

I
ndividuals in developed societies are increasingly plagued by health problems that
are inextricably rooted in hormonal abnormalities. Excessive food intake and relative
inactivity lead to insulin resistance and type 2 diabetes (T2D), a condition associated
with metabolic and endocrine abnormalities that are collectively described as metabolic
syndrome. In metabolic syndrome, many tissues that are essential for metabolism become
resistant to the adipocyte hormone leptin as well as to insulin. The stress of contemporary
life also increases secretion of cortisol to levels that adversely affect mental well-being,
muscle and bone integrity, immune function, and metabolism. Amazingly, most of these
metabolic and endocrine abnormalities can be prevented and even resolved through increased
physical activity, indicating that hormone secretion and action are functionally linked to
exercise and are necessary for the maintenance of health.
The connection between exercise, hormones, and health has been studied and acknowl-
edged for several decades. However, the field of kinesiology to date has not sufficiently
engaged in exploring and exploiting this connection. Although medical science acknowledges
the importance of exercise in preventing morbidities and in resolving many health conditions,
it still relies more heavily on medications and surgery than on exercise prescription and
lacks sufficient understanding of exercise. Kinesiologists, experts on all matters regarding
physical activity, remain inadequately exposed to endocrinology despite the recent addition
of chapters on hormones to exercise physiology textbooks. Courses in exercise endocrinology
are not prominently offered or taught in most exercise science curricula. It is my firm belief
that this neglect or disinterest in exercise endocrinology is a temporary condition that is caus-
ing a loss of research opportunities, which a new generation of kinesiologists will correct.
My own involvement in exercise endocrinology started with an accidental discovery in
the mid-1970s of a striking endocrine effect of physical activity. This changed the course
of my research and professional career and turned me into an exercise endocrinologist. My
core interest in energy regulation led me to pit voluntary exercise against palatable food to
see how animals presented with these two options would maintain body weight. The unex-
pected discovery that mature hamsters provided with a running wheel showed a dramatic
increase in skeletal and somatic growth led me to measure their growth hormone (GH),
develop and validate radioimmunoassay (RIA) specific to GH in this species, and try to
understand why exercise appeared to reverse the slow growth in the mature stage of life to
the rapid growth that is characteristic of immature animals. I have pursued the connection
between physical activity, hormones, and human physiology ever since.
I soon realized that to tell the story about how hormones and physical activity interact one
has to go beyond presenting hormone actions organized by their source of origin, a convention
that is used in most endocrinology textbooks and in most chapters on exercise physiology.
All hormones have pleiotropic actions—anabolic, catabolic, metabolic, or immune—that
differ by the tissue in which they act and by type of physical activity. When I contemplated
presenting an integrative account of the interaction of hormones and exercise, few texts on
hormones and physical activity existed. Some of them were very old, and some were edited
collections of poorly integrated chapters. Generally missing was an accurate view of how

v
vi Preface

acute physical activity affects hormone secretion and how such endocrine action affects
adaptations to habitual exercise. My first attempt to present this story took 10 years and
resulted in Exercise Endocrinology, published by Human Kinetics in 2003.
Science has changed a great deal since 2003. Endocrinology and exercise physiology
have become more molecular, and a great deal more knowledge has accumulated about
hormones and exercise. Because a good teacher and writer interprets and organizes new
knowledge, I accepted the challenge of writing Advanced Exercise Endocrinology. Format
requirements in this series made Advanced Exercise Endocrinology more compact than
Exercise Endocrinology, but it remains highly integrative and richly illustrated. Interactions
between physical activity and hormones are presented from a more molecular perspective
rather than a phenomenological perspective. Therefore, students and other readers seeking
more descriptive information about hormone secretion and physiology can revisit Exercise
Endocrinology.
Chapter 1 of Advanced Exercise Endocrinology introduces readers to the essentials of
hormone and autonomic messenger chemistry, hormone–receptor interactions, and principal
transduction pathways. Chapter 2 focuses on elicitation of endocrine signaling and actions
by muscle contractions and physical activity without the need for hormonal involvement or
hormone–receptor binding. Endocrine and autonomic control of cardiorespiratory function
is presented in chapter 3, which connects the role of peripheral messaging discussed in chap-
ter 2 with the endocrine control of heart function by medullary and pontine brain centers.
Chapters 4 and 5 discuss endocrine involvement in fluid and fuel regulation, respectively,
during exercise. Chapter 6 deals with hormonal regulation of energy balance through brain
controls over intake and expenditure at rest and during exercise. The role of reproductive
hormones in acute exercise and training adaptations is discussed in chapter 7, and involve-
ment of other hormones in these adaptations is discussed in chapter 8. Because essential
aspects of life-supporting processes such as HR, respiration, hormone secretion, and cel-
lular growth are rhythmical and are affected by exercise, the role of biological rhythms in
hormone secretion and exercise physiology is the subject of chapter 9. I am hoping readers
of this text will become energized and interested in engaging in exercise endocrinology
research. To facilitate this, chapter 10 discusses principles of hormone measurements and
ways to avoid some common pitfalls.
One exciting aspect of exercise endocrinology is how it connects many sciences that
study both hormones and physical activity. An exercise endocrinologist could explore
ways that physical activity interacts with nutrition to enhance exercise performance and
training adaptations or could focus on the effectiveness of exercise in preventing pediatric
obesity, metabolic disease in middle age, or manifestations of aging. I hope that readers of
this book, including graduate students, advanced undergraduates,
and professionals in many health fields, will find research ideas
for the study of unresolved questions on how interactions between
physical activity and hormone action can help maintain health,
improve exercise performance, and prevent metabolic disabilities.
Series Preface

H
aving a detailed knowledge of the effects of exercise on specific physiological
systems and under various conditions is essential for advanced-level exercise physi-
ology students. For example, students should be able to answer questions such as
these: What are the chronic effects of a systematic program of resistance training on cardiac
structure and function, vascular structure and function, and hemostatic variables? How do
different environments influence the ability to exercise, and what can pushing the body to
its environmental limits tell us about how the body functions during exercise? When muscles
are inactive, what happens to their sensitivity to insulin, and what role do inactive muscles
play in the development of hyperinsulinemia and type 2 diabetes? These questions and many
others are answered in the books in Human Kinetics’ Advanced Exercise Physiology Series.
Beginning where most introductory exercise physiology textbooks end their discussions,
each book in this series describes in detail the effects of exercise on a specific physiological
system or the effects of external conditions on exercise. Armed with this information, students
will be better prepared both to conduct the high-quality research required for advancing
scientific knowledge and to make decisions in real-life scenarios such as the assessment
of health and fitness or the formulation of effective exercise guidelines and prescriptions.
Although many graduate programs and some undergraduate programs in exercise science
and kinesiology offer specific courses on advanced topics in exercise physiology, there are
few good options for textbooks to support those classes. Some instructors adopt general
advanced physiology textbooks, but such books focus almost entirely on physiology without
emphasizing exercise physiology.
Each book in the Advanced Exercise Physiology Series addresses the effects of exercise
on a certain physiological system (e.g., cardiovascular or neuromuscular) or in certain con-
texts (e.g., in various types of environments). These textbooks are intended primarily for
students, but researchers and practitioners will also benefit from the detailed presentation
of the most recent research regarding topics in exercise physiology.

vii
This page intentionally left blank.
H APTER

1
C

Unique Properties
of Endocrine
and Autonomic
Messengers
T
he endocrine system and the autonomic nervous system (ANS) are the two main
control systems that regulate a body’s internal environment and adjust physiological
responses to exercise and other stressors. The endocrine system consists of chemi-
cal messengers released into circulation or into extracellular and intracellular fluids by
endocrine glands (glands of internal secretion) and by endocrine cells dispersed through
many body tissues and organs. The endocrine system uses circulation as a principal mode
of dispersing its messengers, whereas the ANS communicates primarily through efferent
neurons that transmit messages of the central nervous system (CNS) to target organs by
way of neurotransmitters released into synaptic clefts. The ANS also uses the endocrine
route of communication because its S neurotransmitters, the catecholamines, spill over into
circulation and act as hormones. The adrenal catecholamine epinephrine (E) is a hormone
secreted into circulation upon activation by a sympathetic (S) nerve. Many neurons secrete
hormones into the blood and are therefore defined as neuroendocrine cells. For instance,
secretion of pituitary hormones is activated by neuroendocrine messengers released into
hypothalamo-hypophyseal portal circulation by hypothalamic neurons. The functions of
the endocrine system and the ANS are introduced together in this chapter because their
features regarding secretion and action are similar.

1
2 Advanced Exercise Endocrinology

FUNCTIONS OF THE ENDOCRINE SYSTEM


AND AUTONOMIC NERVOUS SYSTEM
Homeostasis, the regulation of the internal environment by the endocrine system and the
ANS, entails maintenance of a steady state of body fluids, circulation, blood pressure, and
a number of variables in the extracellular fluids such as ionic and metabolite concentra-
tions. The endocrine system and the ANS directly influence energy balance by mobilizing
metabolic fuels out of storage depots and controlling fuel utilization and energy expendi-
ture. They also control intake and uptake of nutrients into cells and synthesis of fuel stores.
Hormones and the ANS messengers also affect body growth and the body’s structural and
functional differentiation by orchestrating developmental changes and by facilitating adapta-
tions to metabolic and mechanical challenges of exercise. Hormones control reproductive
functions, which also can undergo changes in response to energy drain and the stress of
some types of exercise.
Hormones and the ANS also respond to environmental and bioenergetic challenges and
states of emergency (including exercise, food deprivation, exposure to cold, injury, and
infection) when additional metabolic, cardiorespiratory, and immune responses are called
for. At such times, homeostatic body functions can be overruled to accommodate operation
of emergency body functions. For instance, during exercise, aspects of fluid regulation and
digestive functions are suspended when blood flow is routed away from visceral organs
toward the muscles. Hormones that regulate extracellular fluid volume in a sedentary
steady state are applied toward supporting and enhancing circulatory function (see chapter
3). Suprahomeostatic concentrations of metabolic fuels may be mobilized at high exercise
intensities in anticipation of energy needs.
Both hormones and the ANS complement the function of the other, and they frequently
interact. Certain hormones can modulate autonomic function, as illustrated by induction of
S adrenergic receptors (AdR) by thyroxine (T4). Autonomic nerves, on the other hand, influ-
ence secretion of many hormones. During exercise, S nerves suppress secretion of insulin
and facilitate secretion of glucagon and adrenal E. S nerves also influence the timing of
many physiological and endocrine functions.
In both the endocrine system and the ANS, adjustments in body functions are auto-
matic and largely beyond conscious detection or control. We are not aware of endocrine or
autonomic adjustments in concentrations of metabolic fuels or in plasma volume (PV) and
osmolarity. The known exceptions to this generalization are rare but noteworthy. Release
of cholecystokinin (CCK) in response to eating a meal rich in protein and fat is associated
with sensations of satiation, which can be reproduced by intravenous injection of the hor-
mone itself (Little et al., 2005). Loss of plasma sodium (Na) and PV stimulates secretion
of angiotensin II. This hormone increases both thirst and preference for salty solutions and
food. Again, the effects can be reproduced by intravenous or intracerebroventricular injec-
tion of the hormone (Fitzsimons, 1998).
Although many hormones have access to all cells in the body through systemic circula-
tion, they achieve their specific effects by possessing distinctive chemical properties and
by binding to stereospecifically and chemically matched receptors located either on cell
membranes or inside the cells or their nuclei. Autonomic nerves secrete their neurotrans-
mitters from axon terminals into a synaptic cleft on target organs. There, the autonomic
neurotransmitters act on membrane receptors.
Unique Properties of Endocrine and Autonomic Messengers 3

Neurotransmitters and Hormones of the ANS


The most recognized part of the ANS is its efferent component consisting of the S and
parasympathetic (PS) divisions (figure 1.1). These include efferent nerves, ganglia, and
nerve networks (plexuses). Efferent nerves in both divisions consist of two neurons—a
preganglionic neuron and a postganglionic neuron. An enteric nervous system is usually not
included as part of the ANS, although it is under ANS control. It consists of neural plexuses
and glands within the walls of the GI tract.
Figure 1.1 shows the origins of PS and S nerves and their trajectories through, or synapses
in, the autonomic ganglia. PS nerves originate in the hindbrain, which they exit through
the 3rd, 7th, 9th, and 10th (vagus) cranial nerves, and in the sacral spinal cord, from where
they innervate lower alimentary and genital organs. The vagus innervates most body organs
except the skin. Long preganglionic fibers of the vagus and other PS nerves synapse with
postganglionic neurons in PS ganglia (figure 1.1, A-G) that are in proximity of the target
organs. Preganglionic fibers of splanchnic S nerves originate in the gray matter of the spinal

A
III

B
VII
IX
C X H

D
J
T1
2
3 Arm
E 4
5 Heart
F 6 Viscera
7
Heart 8 K
9
10
11
12
L1 Adrenal
2 medulla
(preganglionic
Bowel supply)

G 25
3 L

Leg
Terminal ganglion
(coccygeal) Sympathetic chain

Figure 1.1 Organization of the peripheral ANS.


Reprinted from Clinics in Endocrinology and Metabolism Vol. 6, M.S. Moskowitzby, “Diseases of the autonomic nervous system,” pgs. 745-
768, copyright 1977, with permission from Elsevier.
E4517/Borer/Fig 1.1/400179/alw/pulled-R2
4 Advanced Exercise Endocrinology

cord, which they exit through the thoracic and lumbar spinal nerves. They either synapse
with the postganglionic fibers in the paravertebral S chain ganglia (figure 1.1, H and J) or
traverse through this chain to synapse in the prevertebral (preaortic) ganglia (figure 1.1, K
and L). The postganglionic S neurons then innervate most of the same organs that receive
PS postganglionic fibers.

Sympathetic Division of the ANS


Preganglionic S nerves originate in the intermediolateral (IML) region of the spinal cord
gray matter (figure 1.2). The preganglionic fibers of cardiac accelerator nerves synapse in
three cervical and four to five thoracic S ganglia forming the dorsal end of the 24 paired
paravertebral S ganglia that lie adjacent to the spinal cord (figure 1.1, H and J). Others, like
the splanchnic nerves to the gut organs and the adrenal medulla (figure 1.1, K), traverse
the paravertebral ganglia to synapse in the prevertebral or preaortic ganglia lying near
the origin of arteries branching off the aorta. Postganglionic S nerves originate in the S
ganglia and terminate in target organs and tissues throughout the body. S nerves increase
metabolic and circulatory functions during exercise through axonal release of NE from
postganglionic neurons into synaptic clefts in target tissues where NE binds to AdRs on
target cell membranes. Besides this neurotransmitter action, NE also spills over into target
organ vascular systems from where it acts as a hormone reaching distant cell membrane
receptors through circulation.
A notable exception to this anatomical pattern of S nerves is cells in the adrenal medulla
(figure 1.2). The adrenal medulla is an S ganglion in which the postganglionic neurons
have developmentally been converted into hormone-secreting endocrine cells under the

Intermediolateral
column of the
spinal gray matter

Paravertebral sympathetic chain


T10

T11 Celiac sympathetic ganglion


T12
L1
Adrenal
Preganglionic gland
sympathetic
neuron

Kidney

Figure 1.2 S innervation of the catecholamine-producing cells in the medulla.


E4517/Borer/fig.1.2/400184/TB/R2-alw
Unique Properties of Endocrine and Autonomic Messengers 5

influence of cortisol, which reaches them through blood vessels originating in the adrenal
cortex. Adrenomedullary cells are activated by preganglionic S neurons arising from the
T10 through L1 segments of the spinal cord. Approximately 80% of adrenal medullary
endocrine output consists of E and only 20% consists of NE (see figure 1.2).

Parasympathetic Division of the ANS


Parasympathetic preganglionic neurons originate in the hindbrain nuclei. Long preganglionic
PS neurons originating in the cranial nuclei associated with the 3rd, 5th, 7th, 9th, and 10th
cranial nerves synapse in the ganglia located near the target organs (see figure 1.1). A short
postganglionic neuron in the first three PS nerves activates tear and salivary glands and facial
muscles (Porges, 1998). Some preganglionic neurons of the 10th (or vagus) nerve originate
in the DMV and synapse in the ganglia that innervate most visceral organs (see figure 3.1).
Preganglionic vagal neurons from the nucleus ambiguus (NA) synapse with postganglionic
neurons in ganglia close to the heart and respiratory organs. Thus, the vagus nerve inner-
vates the vast majority of organs and tissues that are also innervated by the S nerves. The
exception to this dual innervation is the skin and the blood vessels, which receive only the
S innervation. The vagus nerve usually antagonizes S actions.

Enteric Nervous System


The enteric nervous system is usually considered to be separate from the ANS, but it oper-
ates under the control of the ANS. Its component hormones and neural plexuses are engaged
in controlling digestion and absorption of nutrients and usually are activated by the vagus.
Because of its functional relationship to vagal control of digestion, the enteric nervous
system, along with the vagus, is suppressed during exercise that activates the S nervous
outflow. A number of gut peptides, such as glucose-dependent insulinotropic peptide (GIP,
previously called gastric inhibitory peptide) and glucagon-like peptide-1 (GLP-1), have
insulin-stimulating (incretin) functions. Due to endocrine actions of incretin hormones,
ingested nutrients trigger an earlier and more pronounced insulin response than do the same
nutrients infused intravenously. The biological function of the incretin effect is to amplify the
insulin response in advance of the circulatory arrival of absorbed glucose. The autonomic
nerves can activate enteric endocrine cells directly or through the diffuse neural plexuses
(networks that permeate different layers of the GI tract as well as other hollow organs).

CLASSIFICATION OF HORMONES
Hormones are usually classified according to their chemical structure. Originally, the
terms hormone (from the Greek verb hormein, which means “to excite”) and endocrine
were applied only to biologically active messengers synthesized within glands of internal
(Greek endos) secretion (Greek hrinein), which release their products into systemic circu-
lation (figure 1.3; table 1.1). Secretin, secreted by the duodenum, was the first hormone to
be discovered more than a century ago (Bayliss & Starling, 1902). Since then, the defini-
tion of hormones has been extended to routes of delivery other than systemic circulation.
Paracrine messengers are released into extracellular fluids in the proximity of cells that are
of different type than the cell of origin. For instance, in addition to its endocrine route of
dispersal, somatostatin or somatrotropin-release inhibiting hormone (SRIF) is released out
of D cells into extracellular space of pancreatic islets to suppress the release of glucagon
and insulin from the adjacent A and B cells, respectively. Autocrine messengers bind to
the cell membrane receptors of the cell that released them. Intracrine messengers act on
6 Advanced Exercise Endocrinology

Intracrine action

Endocrine action

Autocrine action Paracrine


action
H
Secretory
cells H

Target
Neurotransmission
cells

Capillary
Nerve
cells H

Neuroendocrine action

Figure 1.3 Hormones (H) are biologically active substances that are released into circulation by endocrine
E4517/Borer/fig.1.3/400188/TB/R3-kh
and neuroendocrine cells. Neurotransmitters act by releasing messengers into synaptic clefts. Biologically active
substances released into intracellular space act in paracrine fashion on adjacent cells and in autocrine fashion on
the cells of origin. Intracrine action occurs when biologically active substances bind to receptors in the cell of origin.

receptors in the cells that produced them (figure 1.3). Hormones also can be released from
neuroendocrine neurons, and neurotransmitter binding to receptors in synaptic clefts has
the same properties as hormones binding to their receptors.
Just as the definition of hormones has been extended to include routes other than cir-
culatory routes, the concept of the source of hormones has been extended beyond glands
of internal secretion (summarized in table 1.1) to include cells and tissues not confined to
endocrine glands (summarized in table 1.2). Hormones are released by tissues and organs
that have other distinctive functions (e.g., brain, muscle, bone, kidney, liver, and skin).

CHEMICAL STRUCTURE OF HORMONES


Hormones can be classified on the basis of their chemical structure as protein and peptide
hormones, hormones derived from amino acids, and steroid hormones and other lipid mes-
sengers.

Protein and Peptide Hormones


Some proteins and peptides, some steroids and other lipids, and some amino acid derivatives
(or amines) can act as hormones. Protein or peptide hormones and amine hormones are water
soluble and bind to cell surface receptors because they cannot penetrate the lipid-rich cell
membrane. Steroid hormones can cross cell membranes and bind to intracellular or nuclear
Table 1.1 Glandular Sources of Chemical Messengers
Organ Location Hormone
Adrenal gland: cortex Adjacent to the dorsal Aldosterone, androstenedione, cortisol,
surface of the kidney dehydroepiandrosterone (DHEA), estrone
Adrenal gland: medulla Core of the adrenal Enkephalin, E, NE
gland
Gonads: ovary Abdominal cavity Activin, androstenedione, inhibin, estradiol (E2),
estriol, estrone, FSH-releasing peptide, Müllerian
inhibiting hormone (MIH), progesterone (P4), relaxin
Gonads: testis Scrotum Activin, androstenedione, E2, MIH, testosterone (T)
Pancreatic islets of Pancreas (abdominal Gastrin, glucagon T, insulin, pancreatic polypeptide
Langerhans cavity) (PP), vasoactive intestinal peptide (VIP)
Parathyroid gland Thyroid gland Parathyroid hormone (PTH)
Pineal gland Dorsal midbrain Biogenic amines, melatonin, various peptides
Pituitary gland: anterior Ventral surface of the ACTH (or corticotropin), beta-lipotropin (β-LPH),
lobe (adenohypophysis) brain β-endorphin, follicle-stimulating hormone (FSH or
gonadotropin), growth hormone (GH), luteinizing
hormone (LH or gonadotropin), prolactin (PRL),
thyroid-stimulating hormone or thyrotropin (TSH),
pro-opiomelanocortin (POMC)
Pituitary gland: Ventral surface of the Beta-endorphin (β-endorphin), alpha melanocyte-
intermediate lobe brain stimulating hormone (α-MSH)
Pituitary gland: Attached to the ventral Antidiuretic hormone (ADH) or arginine vasopressin
posterior lobe surface of the brain by (AVP), oxytocin (OXY)
(neurohypophysis) infundibular stalk
Thyroid gland Anterior aspect of the Calcitonin, tri-iodothyronine (T3), reverse T3 (rT3),
neck thyroxine (T4)

Table 1.2 Nonglandular Sources of Chemical Messengers


Organ Location Hormone
Adipose tissue Subcutaneous, visceral, ASP, adiponectin, adipsin, angiotensinogen,
retroperitoneal, estrogen, leptin, plasminogen activator inhibitor-1
perirenal, epididymal, (PAI-1), resistin, tumor necrosis factor alpha
and other dispersed (TNFα)
locations
Bone cells (osteoblasts, Skeleton Bone morphogenetic proteins, fibroblast growth
osteoclasts) factors (FGFs) -1 and -2, osteocalcin, osteopontin,
platelet-derived growth factor (PDGF), ligand to the
receptor activator of nuclear factor-κB (RANKL),
transforming growth factor beta (TGF-β)
(continued)

7
Table 1.2 (continued)
Organ Location Hormone
Brain Skull Agouti-related peptide (AgRP), alpha-melanocyte-
stimulating hormone (α-MSH), AVP, β-endorphin,
brain natriuretic peptide (BNP), corticotropin-
releasing hormone (CRH) or factor (CRF),
dynorphin, gastrin, growth hormone-releasing
hormone (GHRH), gonadotropin-releasing hormone
(GnRH), insulin, leu-enkephalin, LHRH, melanin-
concentrating hormone (MCH), met-enkephalin,
neuropeptide Y (NPY), orexin (hypocretin), oxytocin
(OXY), pro-opiomelanocortin (POMC), somatostatin
(somatotropin-release inhibiting hormone, SRIF),
thyrotropin-releasing hormone (TRH), vasoactive
intestinal peptide (VIP)
GI: stomach Abdominal cavity Gastrin, gastrin-releasing peptide (GRP), ghrelin, T
(islets of Langerhans)
GI: duodenum Abdominal cavity Cholecystokinin (CCK), secretin, ghrelin,
gastric inhibitory peptide (glucose-dependent
insulinotropic peptide, GIP)
GI: pancreas T Abdominal cavity Gastrin, glucagon, insulin, pancreatic polypeptide
(PP), VIP
GI: small intestine Abdominal cavity Enteroglucagons (glicentin and oxyntomodulin),
glucagon-like peptide-1 and -2 (GLP-1, GLP-2),
GIP, peptide tyrosine tyrosine (PYY)
GI: large intestine Abdominal cavity PYY
Heart Atria Atrial natriuretic peptide (ANP, or auriculin or
atriopeptin)
Kidney Abdominal cavity 1,25’-dihydroxycholecalciferol (vitamin D3,
calcitriol), erythropoietin (EPO), renin
Liver Abdominal cavity Insulin-like growth factor-I (IGF-I)
Macrophages, Blood ACTH, cytokines: interleukins IL-1 through IL-17,
monocytes, POMC-derived peptides, transforming growth
lymphocytes factor beta (TGFβ), tumor necrosis factor alpha
(TNFα)
Muscle Muscles IGF-I, IGF-II, IL-6, irisin
Placenta Attached to uterus Estrogen, GH variant, human chorionic
gonadotropin (HCG), human placental lactogen
(HPL), progesterone (P4)
Platelets Blood Platelet-derived growth factor (PDGF), TGFβ
Skin Body surface Epidermal growth factor (EGF), transforming
growth factor alpha (TGFα)
Thymus Chest Thymosin (thymopoietin)
Various locations Various tissues, organs Growth factors (including IGF-I), eicosanoids
Vascular endothelium Blood vessels Angiotensin II, endothelin-1 (ET-1), endothelium-
derived hyperpolarizing factor (EDHF), vascular
endothelial growth factor (VEGF), nitric oxide
(NO), PDGF, prostaglandins (PG), PGH2, PGI2,
thromboxane A 2, TGFβ

8
Unique Properties of Endocrine and Autonomic Messengers 9

receptors. Chemical properties of hormones influence the rate of their degradation. The
half-life of amine hormones is less than 5 min, of polypeptides less than 40 min, of protein
and steroid hormones about 2 to 3 h, and of thyroid hormones from 1 to 7 d.
Protein and peptide hormones range in size from as few as three amino acids in TRH
to several hundred amino acids in GH and PRL. This prevalent type of hormone is synthe-
sized in the granular or rough endoplasmic reticulum as inactive prohormone that requires
posttranslational cleavage and removal of a signal peptide and, frequently, of other inactive
molecular fragments as a preprohormone. As an example, proinsulin contains three amino
acid chains: A, B, and C (figure 1.4). A cleavage of the C peptide chain that connects A and
B chains produces active insulin. Because it is not as rapidly degraded as insulin, changes
in C peptide concentration can serve as a measure of insulin secretion.

NH2 NH2
C-peptide
Gly Phe
Ile Val
Val Asn
Glu Gln
Gln His
Cys Leu
Cys Cys
Thr Gly
NH2 Ser Ser
Ile His
Cys Leu
Ser Val
Leu Glu
Tyr Ala
Gln Leu
A chain
Leu Tyr
Glu Leu
Asn Val
Tyr Cys
Cys Gly
COOH
Asn Glu
Arg
COOH Gly
A chain Phe
Phe
Tyr
Thr
Pro
Lys
Thr

B chain COOH
B chain
Proinsulin Insulin

Figure 1.4 Structure of proinsulin. Proinsulin is converted to the biologically active insulin after cleavage of the
C peptide that connects A and B chains in the prohormone.
Springer and Chapman Hall, Hormones: E4517/Borer/Fig 1.4/400189/alw/pulled-R2
From molecules to disease, 1990, page 26, Hormones: A complex communication network, E.E.
Baulieu, edited by E.E. Baulieu and P.A. Kelly, figure 1.22, with kind permission from Springer Science+Business Media B.V.
10 Advanced Exercise Endocrinology

Further processing of a preprohormone may yield more than one hormone. Preprohormone
POMC (figure 1.5) is cleaved into ACTH along with β-endorphin and β-LPH or γ-LPH.
After synthesis and posttranslational processing, protein and peptide hormones are
released from the Golgi apparatus and stored in secretory vesicles that protect them from
enzymatic degradation and allow for regulated release. Some protein hormones are coupled
with carbohydrate. Examples of such glycoproteins are the gonadotropic hormones LH and
FSH as well as TSH.

Hormones Derived From Amino Acids


Amino acid hormones are derivatives of aromatic amino acids. Catecholamines NE and
E are synthesized through a series of enzymatic modifications of the amino acid tyrosine
(figure 1.6).

Signal

N-terminal
peptide
N-terminal
fragment -MSH

Connecting
peptide

-MSH
(1-13)
ACTH
ACTH
(1-39) CLIP
(18-39)

-LPH
(1-56)

-LPH -MSH
(1-89) (84-101)
Metenkephalin
-endorphin (104-108)
(59-89)

Figure 1.5 POMC and its derivative hormones. After cleavage and removal of the signal peptide and N-terminal
fragment, the POMC preprohormone rise in the anterior pituitary gland to ACTH and β-LPH. Subsequent
gives1.5/400190/alw/pulled-R2
E4517/Borer/Fig
cleavage of β-LPH yields β-endorphin and γ-LPH.
Springer and Chapman Hall, Hormones: From molecules to disease, 1990, page 230, ACTH and related peptides, A.S. Liotta and D. Krieger,
edited by E.E. Baulieu and P.A. Kelly, figure IV.1, with kind permission from Springer Science+Business Media B.V.
Unique Properties of Endocrine and Autonomic Messengers 11

OH OH OH OH OH
HO HO HO HO
TH AADC DBH PNMT

HCH HCH HCH HC–OH HC–OH


HC–COOH HC–COOH HCH HCH HCH
HNH HNH HNH HNH HN–CH3
L-tyrosine L-DOPA Dopamine L-norepinephrine L-epinephrine

Figure 1.6 Catecholamine biosynthesis. Catecholamines are synthesized from the essential aromatic amino
acid tyrosine. Enzymes tyrosine hydroxylase (TH), aromatic L-amino acid decarboxylase (AADC), and dopamine
β-hydroxylase (DBH) mediate formation of NE from tyrosine. Conversion of NE to E in the adrenal medulla and
some brain neurons is catalyzed byE4517/Borer/Fig 1.6/400191/alw/pulled-R1
the enzyme phenylethanolamine-N-methyltransferase (PNMT).
Reprinted from Metabolic Control and Disease, 8th ed., edited by P.K. Bondy and L.E. Rosenberg, Catecholamines and the adrenal medulla, L.
Landsberg and J.B. Young, pg. 1626, copyright 1980, with permission from Elsevier.

Serotonin and melatonin are derivatives of the amino acid tryptophan. The thyroid
hormones T4, T3, and rT3 are synthesized from two molecules of tyrosine and three or four
atoms of iodine (figure 1.7). Amine hormones also are stored in secretory vesicles.

Steroid Hormones and Other Lipid Messengers


Steroid hormones are synthesized from cholesterol in the smooth endoplasmic reticulum
in the adrenal cortex and in male and female gonads. Following a side-chain cleavage, the
residual portion of the cholesterol molecule undergoes a series of enzymatic transformations
in the adrenal cortex (figure 1.8) to produce the mineralocorticoid aldosterone in the exter-
nal zone (zona glomerulosa), glucocorticoid cortisol in the middle zone (zona fasciculata),

I 3' I 3
Thyroxine NH2
(T4) HO O CH2CH
COOH
I 5' I 5

I I
3,3',5- NH2
Triiodothyronine HO O CH2CH
(T3) COOH
I

I I
3,3',5'- NH2
Triiodothyronine O CH2CH
(reverse T3) COOH
I

Figure 1.7 Processing of thyroid hormones. In the thyroid gland, tyrosine is iodinated by tyrosine peroxidase to
form T4 that is stored bound to the 660Borer/E4517/Fig.
kDa thyroglobulin.1.7/400192/TimB/R2-alw
After hydrolysis by lysosomal proteases, T4 is released
into circulation. Eighty percent of T4 is converted to T3 or rT3 in the liver and the kidney by 5’-deiodinase and
5-deiodinase, respectively.
Adapted, by permission, from J.E. Griffin, 1982, Manual of clinical endocrinology and metabolism (New York: McGraw-Hill), 50. © McGraw-
Hill Companies.
12 Advanced Exercise Endocrinology

21 22

20

19 12 18 17
11 16

1 9
C D
10
2 14 15
A B
HO 3 5 7
4 6
Cholesterol

D 17H L
Pregnenolone 17-Hydroxypregnenolone DHEA DHEA-S
3β HSD
Progesterone 17-Hydroxyprogesterone Androstenedione
21 H
Deoxycorticosterone 11-Deoxycortisol
11 H
Corticosterone Cortisol O
18 H
18-Hydroxycorticosterone
CH2OH
18 O
Aldosterone C=0
HO
HO
CH2OH O
O
C=0 Androgen
HCl
HO pathway
O
Glucocorticoid
pathway
O
Mineralocorticoid
pathway

Figure 1.8 Biosynthetic pathways of adrenal corticosteroid


Borer/E4517/Fig. hormones. Adrenal corticosteroid biosynthesis
1.8/400193/TimB/R3-kh
from cholesterol is stimulated by corticotropin or ACTH. ACTH stimulates the activity of 20,22 desmolase (D), the
rate-limiting side-chain cleavage enzyme in the P450 cytochrome. A series of hydroxylases (H), hydroxysteroid
dehydrogenases (HSD), or oxidases (O) and chain-cleaving Ds [17,20 D, or lyase (L)] channel the corticosteroid
derivatives down the mineralocorticoid, the glucocorticoid, or the androgen biosynthetic pathways.
Adapted from Endocrinology and Metabolism Clinics of North America, Vol. 23(2), P.C. White and P.W. Speiser, “Steroid 11ß-hydroxylase
deficiency and related disorders,” pgs. 325-329, copyright 1994, with permission from Elsevier.

and androgen androstenedione in the innermost zone (zona reticularis) of the cortex (see
“21-Hydroxylase Deficiency and Na Hunger” at the end of this chapter).
Gonadal steroids are synthesized in specialized cells in the ovary and the testis (figure 1.9)
in response to stimulation by the anterior pituitary gonadotropins LH and FSH. LH stimulates
production and secretion of T from the interstitial Leydig cells in the testis and synthesis
of androgens from the interstitial theca cells of the ovary. FSH then converts androgens
into estrogen in the inner granulosa cell layer that forms the follicle surrounding the egg.
Vitamin D is another hormone derived from cholesterol. However, conversion into active
vitamin D3 (also called calcitriol) requires rupture of the second benzene ring and two
sequential hydroxylations in the liver and the kidney (figure 1.10).
Adrenal
Ovary and testis
21 22 24 26

20 23 25
19 18 17
12 16 27
11 13
1 C D
2 9 14 15
10 8
A B
HO 3 5 7
4 6
Cholesterol
O O
1
2

HO O
Pregnenolone Progesterone

3 O 3 O O
OH OH
2

HO

HO O HO
17α-hydroxypregnenolone 17α-hydroxyprogesterone Catechol estrogen

4 O 4 O O
2 5

HO O HO
Dehydroepiandrosterone Androstenedione Estrone

6 OH 6 OH
5

O HO
Testosterone Estradiol
5α-reductase OH

Peripheral
tissues
O
H
Dihydrostestosterone

Figure 1.9 Biosynthesis of sex steroid hormones. Arrows identify the range of sex steroid hormones synthesized
in the testis, ovary, adrenal glands, and peripheral 1.9/400194/Tim
Borer/E4517/Fig. tissues. B/R2-alw
Adapted from TEXTBOOK OF ENDOCRINE PHYSIOLOGY, SECOND AND THIRD EDITIONS, edited by James E. Griffin and Sergio R. Ojeda,
copyright © 1990 and 1996 by Oxford University Press, Inc. Used by permission of Oxford University Press, Inc.

13
14 Advanced Exercise Endocrinology

24
25
H

Sunlight H
1

HO

7-Dehydrocholesterol

OH

CH2

HO
Liver
Cholecalciferol
Hydroxylation

CH2

HO

25-Hydroxycholecalciferol
Hydroxylation
OH

OH OH

Kidney

CH2 CH2

HO OH HO

1,25-Dihydroxycholecalciferol 24,25-Dihydrocholecalciferol
(vitamin D3)

Figure 1.10 Biosynthesis of vitamin D3.


Reprinted from TEXTBOOK OF ENDOCRINE PHYSIOLOGY, THIRD EDITION, edited by James E. Griffin and Sergio R. Ojeda, copyright © 1996
by Oxford University Press, Inc. Used by permission of Oxford University Press, Inc.

E4517/Borer/Fig
Ultraviolet (UV) light breaks the bonds1.10/400195/alw/pulled-R1
between carbons 9 and 10 in the cholesterol deriva-
tive 7-dehydrocholesterol in the mammalian skin to form the prohormone cholecalciferol.
Provitamin D3 is converted in the liver to 25-hydroxycholecalciferol and in the kidney to
the biologically active vitamin D3 (or calcitriol). D3 is inactivated by atom 24 hydroxylation.
Invertebrates, marine phytoplankton, and fungi produce vitamin D2, or ergosterol, from
exposure to UV light. Ergosterol differs from vitamin D3 in that it has a double bond between
Unique Properties of Endocrine and Autonomic Messengers 15

atoms 22 and 23 and a methyl group on atom 24. Mammals can convert it to vitamin D3
after eating food containing ergosterol.
Unlike amine and protein hormones, steroid hormones are not stored in the cells of
origin. Instead, they are released directly into circulation immediately after their synthesis
from the Golgi apparatus via transport vesicles. Eicosanoids, prostaglandins, prostacyclins,
thromboxanes, and leukotrienes also are lipid messengers. They are derived from the essential
20-carbon arachidonic fatty acid in a variety of cells in response to mechanical stimulus of
exercise, growth factors, cytokines, and diets rich in ω-6 fatty acids. These stimuli activate
phospholipase enzymes to hydrolyze phospholipids and diacylglycerol (DAG) in cellular
membranes. The liberation of arachidonic acid by these actions initiates production of four
families of eicosanoids: prostaglandins, prostacyclins, thromboxanes, and leukotrienes.
In general, these autocrine or paracrine messengers facilitate inflammation and influence
immune responses and formation of free radicals, and their action is limited to seconds
or minutes. Consumption of dietary sources of ω-3 fatty acids results in biosynthesis of
molecules that act in an anti-inflammatory fashion.

HORMONE RELEASE AND TRANSPORT


Hormones are released under varied and sometimes complex control. Steroid hormones are
constitutively controlled in that they are released as soon as they are synthesized in response
to a given stimulus. All other hormones are stored in intracellular vesicles, and their release
is regulated by selective responses to specific stimuli. Most water-soluble forms of hormones
such as peptide, protein, and amino-acid derived hormones are transported in circulation in
their native form, whereas the transport of lipid hormone molecules requires the assistance
of water-soluble binding proteins (BPs). GH and some growth factors present an exception
to this rule, most likely to provide an added layer of control over their actions.

Control of Hormone Release


Hormone secretion can be elicited by a variety of changes in the internal or external envi-
ronments that require either an immediate corrective action or adaptation when changes are
of longer duration. Among the immediate hormonal triggers in the service of homeostasis
are changes in the internal environment such as variations in the following: plasma glucose
concentration influencing compensatory release of either insulin or glucagon; plasma Ca
concentration influencing compensatory release of either calcitonin or PTH; plasma Na
concentration stimulating the release of angiotensin II; hydration leading to release of ADH,
aldosterone, or ANP; blood pressure triggering renin–angiotensin–aldosterone reflex; low
temperature triggering the release of NE to brown adipose tissue (BAT) to stimulate ther-
mogenesis; and illumination influencing the release of melatonin. External stimuli also can
trigger immediate hormonal responses. For example, physical or psychological stress will elicit
S activation and the release of E, NE, and, depending on the severity or duration of the stress,
cortisol; exposure to low oxygen levels will elicit the release of EPO; mechanical loading of
bones will stimulate the release of osteocalcin; and touching of the nipples or even thinking
about nursing of an infant will elicit OXY and the milk-letdown reflex in a lactating female.
Prolonged exposure to changes in internal and external environments—the usual sequelae
of exercise training or dietary manipulations—also results in hormonal adaptations. Exam-
ples of hormonal changes to sustained changes in the internal environment include reduced
basal secretion of leptin and insulin in response to loss of body fat, reduced secretion of
growth factors such as IGF-1, reduced secretion of gonadotropic hormones LH and FSH,
16 Advanced Exercise Endocrinology

reduced secretion of thyroid hormones in response to negative energy balance generated


by either exercise energy expenditure or dieting, and increased PTH secretion in response
to a chronic deficit in plasma Ca (secondary hyperparathyroidism). Hormonal changes to
prolonged changes in the external environment include changes in the pattern of secretion
of some hormones, such as oversecretion of thyroid hormones to prolonged cold exposure,
oversecretion of cortisol to chronic stress, and seasonal changes in secretion of melatonin
and sex hormones in seasonal breeders in response to changes in the photoperiod length.
One can discern at least three mechanisms of hormonal responses to these various stimuli.
Among the immediate triggers of hormone release are neurotransmitters from neurons acting
on endocrine cells, as illustrated by the autonomic nerves; hormones released into circulation
by neuroendocrine cells; and hormones released by endocrine cells. These means of hormone
elicitation could be initiated by changes in nutrient or ion concentration in plasma, error signals
triggering negative feedback, mechanical stress acting on mechanotransductive cells, and
photic stimuli acting on photosensitive cells. The chemical nature and the method of storage of
hormones exerts additional influence over hormone release. Thus, peptide and amine hormones
are released from storage vesicles in response to a neural, hormonal, or metabolic stimulus.
By contrast, constitutive secretion of steroid hormone automatically follows its synthesis.
Secretion of hormones may involve a direct stimulus action on neurons or endocrine cells
or a complex interaction between the two cell types. The simplest situation is when neurons
are directly stimulated and release their neuroendocrine product into systemic circulation.
An example is magnocellular neurons in the lateral part of the PVN and in the SON that pro-
duce AVP (also known as ADH) in response to becoming dehydrated (see figure 4.7). These
hormones are stored in secretory vesicles in the posterior pituitary in the axon terminals of
PVN and SON neurons (although they also may release some hormone into hypothalamo-
hypophyseal portal circulation). Upon dehydration of the PVN and SON neurons, AVP is
released into systemic circulation. Likewise, upon stimulation of the nipples of a lactating
woman, OXY is released into circulation. Similarly, nutrients in the gastrointestinal tract
usually directly stimulate secretion of the gastrointestinal hormones gastrin, secretin, CCK,
GIP, GLP-1, PYY, and others, and changes in plasma glucose concentration can elicit the
release of insulin or glucagon from pancreatic β and α cells, respectively.
A more common situation is for brain neurons to control hormone secretion indirectly
through the mediation of ANS neurons or anterior pituitary hormones. Neurons in dorsal and
ventral parts of the PVN control S nerve outflow and thus stimulate secretion of a number
of hormones such as adrenomedullary E (see figures 1.2 and 3.18), glucagon, and renin and
inhibit secretion of insulin. Neurons from several brain regions influence vagal preganglionic
neurons in the dorsal vagal nucleus (see figure 1.4) and facilitate the release of insulin and
digestive gut peptides. Brain regions that control HR exert an influence over the NA in the
medulla oblongata (MO) to elicit vagal release of the neurotransmitter acetylcholine (Ach),
which reduces depolarization of the cardiac pacemaker.
Neurons in the medial part of the PVN nucleus secrete the hormone CRF into hypo-
thalamo-hypophyseal portal circulation (see figure 3.18). In the anterior pituitary, CRF
stimulates pituitary corticotrophs to secrete and release ACTH into systemic circulation.
ACTH promotes secretion of cortisol from the adrenal cortex. Cortisol secreted in the
adrenal cortex reaches the adrenal medulla through centripetal adrenal blood vessels and
stimulates E synthesis (see figure 1.6) in the adrenal medulla. E, in turn, stimulates secre-
tion of ACTH from the anterior pituitary (see figure 3.18).
Similar to the control of adrenal cortisol is the chain of neural and hormonal events in
the control of thyroid and gonadal hormones and of GH secretion. The brain neurons that
Unique Properties of Endocrine and Autonomic Messengers 17

start the chain reaction belong to the hypophysiotropic area in the basal hypothalamus
and secrete releasing hormones (such as CRF in the previous example) or release inhibit-
ing factors (such as somatostatin in the control of GH secretion) into the hypophyseal
portal circulation to influence the release of anterior pituitary hormones (see figure 4.6).
Capillaries in hypophyseal portal circulation pick up TRF, GnRH, GHRH, and inhibiting
hormone SRIF and transport them in portal veins down the hypothalamo-hypophyseal
stalk to a capillary bed in the anterior pituitary gland. There, these hormones initiate the
release of anterior pituitary hormones thyrotropin or TSH, gonadotropins LH and FSH,
and GH, respectively. These anterior pituitary hormones initiate the release of additional
hormones from peripheral endocrine glands: T4 and T3 from the thyroid, estrogens and
androgens from the gonads, and IGF-I from the liver. This illustrates the sequential
endocrine mediation of hormone secretion.
When hormone concentration is maintained within the regulated range, hormone secretion
is under negative feedback control. Thus, ACTH stimulates the secretion of cortisol. When
cortisol concentration exceeds the regulated range, it exerts inhibitory negative feedback
over CRF secretion by PVN neurons and ACTH secretion by pituitary corticotrophs. When
cortisol concentration falls below the target range, reduced negative feedback allows more
cortisol secretion. Similar negative feedback operates in the control of thyroid hormones T3
or T4. These hormones facilitate secretion of TSH when their concentration decreases and
inhibit it when their concentration increases. A diagnostic test for clinical depression takes
advantage of the principle of negative feedback. A low (1 to 2 mg) dose of the synthetic
glucocorticoid analog dexamethasone (DEX) is administered. DEX binds to GR in excess
of endogenous cortisol and initiates negative feedback that signals excess glucocorticoids
in circulation. In healthy individuals, DEX suppresses ACTH and cortisol secretion, but in
clinically depressed individuals the negative feedback is deficient and cortisol concentration
remains elevated. Similar negative feedback is exerted by IGF-I over GH secretion and by
sex steroids over gonadotropin secretion.
Regulation of variables in the internal environment requires two opposing negative
feedbacks. Plasma glucose is regulated within the range of 73 to 110 mg/dl. A decline
below this range elicits the secretion of glucagon (from alpha cells in pancreatic islets)
and other counterregulatory hormones. Glucagon and some other counterregulatory hor-
mones increase plasma glucose concentration by stimulating hepatic glucose release to
restore plasma glucose concentration to the regulated range. Increased insulin response to
postprandial hyperglycemia returns plasma glucose to the regulated range by facilitating
muscle glucose uptake from plasma and increased glucose metabolism. Similarly, PTH and
calcitonin regulate concentration of plasma Ca within a range of 4 to 6 mEq/l by increasing
and decreasing Ca release from the bone, respectively, and through other compensatory
actions on Ca metabolism.
Hormones can also exert positive feedback over secretion of other hormones. In this
case, secretion of one hormone amplifies secretion of the second hormone. An example
is reciprocal augmentation of secretion of the stress hormones E and cortisol (see figure
3.18). During intense exercise or other stress, cortisol is released from the adrenal cortex.
It reaches adrenal medullary cells through centripetal adrenal cortical blood vessels and
activates the PNMT enzyme to produce E (see figure 1.6). E, in turn, reaches the anterior
pituitary gland through systemic circulation to promote ACTH release and more cortisol
secretion. This positive-feedback loop is interrupted when the emergency subsides, and the
release of cortisol declines. Another positive feedback is established between the secretion
of E2 and LH, producing an LH surge that precipitates ovulation.
18 Advanced Exercise Endocrinology

Hormone Transport
Hormone transport in circulation is facilitated in some instances by BPs. BPs extend the
period of protein hormone availability and action by slowing rapid degradation of the hor-
mone by proteolytic enzymes. In some instances BPs facilitate the binding of hormones to
their receptors. For instance, the structure of GH BP and GHR is identical except for the
shorter segment in the GH BP that corresponds to the intracellular segment in GH receptor.
GH BP attaches to the GH molecule already bound to a GHR to form a dimer and initiate
hormone action. The hormones, such as IGF-I, that are bound to carrier proteins persist in
circulation for several hours, whereas most water-soluble peptide and protein hormones are
degraded within minutes to 2 h.
All of the lipophilic steroid hormones are bound to specific carrier proteins when they
circulate in plasma. Examples are sex hormone-binding globulin (SHBG), which has high
affinity for androgens, and thyroglobulin, which binds to thyroid hormones. Nonselective
binding of hormones and FFA occurs with albumin and prealbumin.

HORMONE RECEPTORS AND HORMONE–RECEPTOR


MESSAGE TRANSDUCTION
Hormones produce specific effects by binding only to a stereospecifically and chemically
compatible receptor. Specificity in the binding of hormones to their receptors is sometimes
not absolute because some hormones can cross-react with receptors of a closely related hor-
mone. For instance, insulin and IGF-I can both bind to the same receptors, although each
binds with greater affinity to homologous receptors than to heterologous receptors. Recep-
tors are proteins that can be classified in several ways. Based on their location, receptors
can be membrane, intracellular, or nuclear receptors. Based on their structure, receptors
can be classified by the presence or absence of tyrosine kinase enzyme as a component
of intracellular domain and by the number of their transmembrane domains (figure 1.11).
Based on method of signal transduction, receptors can be sorted by families such as GPCR,
cytokine, and nuclear receptor families.

Membrane Receptors and Their Mechanisms of Action


A lipophilic portion of the amino acid chain of membrane receptors is embedded in the
lipid membrane bilayer. These sequences of a receptor amino acid chain that traverse the
membrane constitute the transmembrane domain of the receptor. As an example, cholinergic
muscarinic receptor (MR) and AdR have seven transmembrane domains (see figure 1.11c).

Insulin Receptor
A tyrosine kinase can be attached to the intracellular domain of receptors with a single
transmembrane domain (see figure 1.11b), as illustrated by dimeric insulin receptor (IR) and
some growth factor receptors (see figure 5.1). Activation of this receptor is not essential for
glucose uptake during exercise but is responsible for postprandial removal of plasma glucose
by muscle and other insulin-dependent tissues. Insulin binding to IR initiates autophos-
phorylation of the phosphotyrosines on the IR and its conformational change. This allows
the phosphotyrosines on IR to recruit effector proteins containing the Src homology 2 and
3 domains. A major recruited protein is insulin response substrate (IRS), which also under-
goes tyrosine phosphorylation. This leads to hydrolysis of membrane phosphatidylinositol
biphosphate (PIP2) by the enzyme phosphatidylinositol-3-kinase (PI3K) to a membrane phos-
pholipid (PIP3; see figure 5.1). PIP3 facilitates recruitment and binding of effector proteins
to the cell membrane. An important recruited protein in this signaling pathway is protein
Unique Properties of Endocrine and Autonomic Messengers 19

NH2 NH2

Membrane

Membrane

TK HOOC

a COOH b d NH2

NH2

Membrane
HVD

C DBD
G

HOOC
c
HBD1

HBD2

e COOH

Figure 1.11 Receptor (R) structure. Four types of membrane receptors for water-soluble hormones (a-d) are
distinguished by the number of their transmembrane domains. (a, b) R with a single transmembrane region may
have a tyrosine kinase (TK) in theirE4517/Borer/Fig
intracellular domain. (c) Rs with seven transmembrane domains are associated
1.11/400196/alw/pulled-R2
with G proteins and catalytic enzyme subunits, and (d) those with five transmembrane regions can form an ion
channel. (e) Intracellular or nuclear R for steroid and thyroid hormones consist of a hypervariable hormone-specific
NOTE: SIZE 2/3-2
domain (HVD), a DNA-binding domain (DBD), and one or two hormone-binding domains (HBD).
TO FIT IMAGE
Figures a-d from Springer and Chapman Hall, Hormones: From molecules to disease, 1990, page 76, Hormones: A complex communication
network, E.E. Baulieu, edited by E.E. Baulieu and P.A. Kelly, figure 1.77, with kind permission from Springer Science+Business Media B.V.
Figure e adapted from TEXTBOOK OF ENDOCRINE PHYSIOLOGY, THIRD EDITION, edited by James E. Griffin and Sergio R. Ojeda, copyright
© 1996 by Oxford University Press, Inc. Used by permission of Oxford University Press, Inc.

kinase B (PKB or Akt). Akt initiates several signaling pathways that are responsible for
cellular uptake of glucose, inhibition of lipolysis, and synthesis of glycogen, fat, and protein.
The second insulin signaling pathway entails activation of MAPK. Effectors upstream of
MAPK include the Grb-2–SOS complex, which binds to IRS, and the membrane-bound
Ras G-protein, which, when phosphorylated, acts as a signaling switch. Ras then activates
Raf and the mitogen-activated protein kinase kinase (MEK). MEK phosphorylates and acti-
vates mitogen-activated protein kinase (MAPK), which consists of ERK 1 and 2. The same
pathway may be activated by GH. Downstream signaling of Akt is discussed in chapters 2
and 5, and downstream signaling of MAPK is discussed in chapters 2 and 8.
20 Advanced Exercise Endocrinology

Cytokine Family of Receptors


Some receptors with a single transmembrane domain that do not possess tyrosine kinase
(see figure 1.11a) include the large cytokine, integrin, low-density lipoprotein, nerve growth
factor, and transferrin families of receptors. The cytokine receptor family includes recep-
tors for GH, EPO, leptin, PRL, cardiotropin, IL, and IFN α, β, and γ. After binding to their
respective receptors, GH and leptin initiate receptor dimerization. Currently, evidence
suggests that GHRs are localized in the membrane as dimers.
Tyrosine kinase is not a component of a GH receptor (GHR) molecule. Instead, upon
binding with GH, GHR recruits a Janus activated kinase-2 (JAK-2). GH binding to GHR
triggers partial GHR rotation, which aligns recruited JAK-2 to propagate GH signal (Lanning
& Carter-Su, 2006). Upon the tyrosine phosphorylation of JAK-2 and the phosphorylation of
GHR, a pair of JAK-2 initiates tyrosine phosphorylation on several sites on effector proteins.
This is a starting point for several GH signaling pathways (see figure 8.1). JAK-2 recruit and
phosphorylate tyrosines on STAT factors (signal transducers and activators of transcription)
that detach from the receptor, dimerize, enter the nucleus, and initiate gene transcription.
Cytokine receptor signal transduction is terminated when STATs induce expression of SOCs
(suppression of cytokine signaling proteins) that competitively block phosphotyrosine acti-
vation. Activated GHR also recruits other substrates containing SH2 domains, such as IRS
and Src kinase. This allows GH to exert insulin-like actions through IRS–PI3K signaling
and to affect protein synthesis through the Shc–Ras–MEK–ERK1/2 signaling cascade. The
downstream effects of these signaling pathways are discussed in chapter 8.

GPCR Family of Receptors


Many receptors have multiple transmembrane domains. For instance, the nicotinic receptor
on the postganglionic cells in S and PS nerve ganglia and on skeletal muscle membranes
contains five transmembrane domains arranged in a circular fashion to form a channel for
transport of Na ion (see figure 1.11d). Seven transmembrane domains characterize cholin-
ergic MR and AdR (see figure 1.11c) that belong to the guanine-protein coupled receptors
(GPCR) family of receptors coupled with G regulatory protein and a catalytic subunit. G
regulatory proteins are heterotrimers consisting of alpha, beta, and gamma subunits (shown
in figures 1.13, 1.14, 1.15, and 1.17) and are coupled to the intracellular loops of AdR and
MR formed by the fifth and sixth transmembrane molecule crossings (shown in figure 1.12).
Hormones Acting Through the Gsα Subunit of GPCR
Autonomic nerves increase heart function through activation of β1 AdR and, along with
angiotensin II, contribute to increases in blood pressure during exercise. Stimulation of β2
and β3 AdR leads to mobilization of FFA and glucose from their storage sites. Fuel mobi-
lization is also facilitated by counterregulatory hormones E, GH, and glucagon and the
steroid hormone cortisol. All of these actions, except for signaling by GH and cortisol, are
mediated by a category of GPCRs that have a stimulatory alpha G protein Gsα (figure 1.13).
The alpha subunit binds a guanosine nucleotide diphosphate (GDP) that is exchanged for a
triphosphate (GTP) upon activation of the receptor by the hormone or S neurotransmitter.
GTP-bound Gsα dissociates from beta and gamma subunits and binds to and activates a
catalytic enzyme, adenylyl cyclase (AC). AC catalyzes conversion of ATP to the second mes-
senger, cyclic adenosine monophosphate (cAMP). cAMP then phosphorylates and activates
cAMP-dependent protein kinase A (PKA) that, in turn, phosphorylates different effector
proteins in several tissues. In the heart and arterioles, PKA phosphorylates a protein that
opens a voltage-sensitive Ca channel on plasma membrane. Increased plasma Ca2+ concen-
Unique Properties of Endocrine and Autonomic Messengers 21

NH2 Extracellular space

Membrane

C
G

HOOC
Cytoplasm

Figure 1.12 A schematic representation of G regulatory protein-coupled receptors that include AdR and MR
with their seven transmembrane domains. Dark blocks indicate intracellular sites where the receptor binds to G
E4517/Borer/Fig
protein. C represents the enzymatic catalytic unit. 1.12/400202/alw/pulled-R2
Springer and Chapman Hall, Hormones: From molecules to disease, 1990, page 59, Hormones: A complex communication network, E.E.
Baulieu, edited by E.E. Baulieu and P.A. Kelly, figure 1.56A, 1.57A, with kind permission from Springer Science+Business Media B.V.

Isoproterenol

Na+
Ca2+ β

γ
βGα
s
AC
K+
GTP
ATP

cAMP

Phosphatase inhibitor 1
Troponin 1
PKA Myofibrillar C protein
Phosphorylase kinase
Phospholamban

Figure 1.13 Signaling pathways of beta AdR. The alpha subunit of G protein dissociates and stimulates AC
catalytic subunit to form the second messenger, cAMP. cAMP activates PKA to open a Ca channel and induce
other enzymatic changes through serial phosphorylations.

E4517/Borer/fig.1.13/400203/TB/R2-alw
tration causes cardiac muscle contractions and arteriolar constriction. In the liver and the
skeletal muscle, PKA phosphorylates and activates phosphorylase kinase, which in turn
phosphorylates and activates the glycogenolytic enzyme glycogen phosphorylase. In the
adipose tissue, PKA phosphorylates and activates the lipolytic enzyme HSL. Phosphorylation
by PKA activates catabolic hormones that break down storage fuels and inactivates enzymes
that synthesize storage fuels, such as glycogen synthetase and fatty acid synthetase (FAS).
22 Advanced Exercise Endocrinology

Catalytic subunits of PKA also enter the nucleus, where they activate a transcription factor
(TF), cAMP response element-binding protein (CREBP), to bind to the cAMP response
element (CRE) near the promoter area of genes (see figure 2.10). This increases the tran-
scription of gluconeogenic proteins such as phosphoenolpyruvate carboxykinase (PEPCK)
and some other enzymes. An intrinsic GTP phosphatase and β-arrestins terminate GPCR
signaling. Cholera toxin inhibits GTP phosphatase and causes a persistent activation of AC.
Hormones Acting Through the Giα and Gqα Subunits of GPCR
A number of hormones oppose the action of hormones acting through Gs regulatory protein
by activating Gqα (figure 1.14) and Giα subunits of GPCR (figure 1.15). These include the

Methoxamine

α1 Ca2+
γ γ
PLC
PIP2 IP3 + DAG β Gqα β Gsα AC

GTP

ATP
PKC

PDE cAMP

PKC-catalyzed
protein phosphorylations 5'AMP

Figure 1.14 Stimulation of α1 receptors (by, for instance, methoxamine) results in PKC-catalyzed protein
phosphorylations and activation of the enzyme phosphodiesterase (PDE) by the alpha subunit of Gq protein. PDE
deactivates cAMP by converting it E4517/Borer/fig.1.14/400204/TB/R2
to 5’ AMP.

Clonidine

Na+ α2
γ γ
β Gsα AC β G iα

GTP K+
ATP

cAMP

PKA-catalyzed
phosphorylations

Figure 1.15 Stimulation of α2 receptors (by, for instance, clonidine) leads to translocation of the beta–gamma
complex of the Gi protein and its interference with the alpha subunit of the Gs protein. This results in suppression
of cAMP synthesis.

E4517/Borer/fig.1.15/400205/TB/R2-alw
Unique Properties of Endocrine and Autonomic Messengers 23

catecholamine actions on α1 and α2 AdR and the actions of Ach on MR, opioids on mu
receptors, cannabinoids on cannabinoid 1 receptors (the last two activate Giα and Goα GPCR
subunits), and endothelin (ET) on ET receptors (activates Gqα GPCR subunit).
An important early role of catecholamines during exercise is redistributing blood from
general systemic circulation to the muscle. This is accomplished in part by vasoconstrictive
action of NE acting on α1 and α2 AdR, by vasopressin acting on V1 receptors on peripheral
vasculature, and in part by vasodilatory effects of NO in the muscle. These circulatory
adjustments and some metabolic effects, such as activation of phosphorylase kinase and
glycogenolysis in muscle by α1 adrenergic stimulation (see figure 2.2), are accomplished
through a signaling pathway involving the Gqα subunit of G regulatory protein (see figure
1.14). One of its actions is activation of the enzyme PDE that converts ATP to inactive 5'
AMP. The α2 receptors activate this pathway as well (see figure 1.15). Dissociation of Gqα
from beta–gamma subunits leads to activation of one of several phospholipases C (PLC-β,
PLC-γ, or PLC-δ) (see figure 1.14). PLC hydrolyses membrane phospholipid phosphatidylino-
sitol biphosphate (PIP2) to a cytoplasmic second messenger, inositol triphosphate (IP3), and
to a membrane messenger, diacylglycerol (DAG). IP3 binds to receptors on the endoplasmic
reticulum to produce a tenfold increase in cytosolic Ca2+ and facilitates influx of extracel-
lular Ca2+(see figures 1.17 and 2.3). DAG, along with phosphatidylserine, activates one of
several types of membrane-associated PKCs, enzymes whose action is enhanced in the
presence of Ca2+. Phorbol esters mimic the DAG-mediated activation of PKC and prolong
enzyme activation. PKCs propagate hormone signals through phosphorylation of serine and
threonine on a variety of effector proteins, including those in the insulin signaling pathway.
Substitution of serine or threonine for tyrosine phosphorylations impairs insulin signaling
and sensitivity. PKC also can activate phospholipase A2, which facilitates production of
proinflammatory eicosanoid molecules from arachidonic fatty acid. The Ca2+ ion influences
muscle contraction, hormone transduction, and other contraction-linked biological events
by binding to proteins such as troponin C, calmodulin (CaM), and CaM-dependent protein
kinase II (CaMK II) (see figure 2.2). One of these effects is stimulation of synthesis of NO
from the amino acid L-arginine (discussed in chapter 2). PKC also has genomic effects
through activation of a TF activator protein-1 (AP-1) that consists of proto-oncogenes c-jun
and c-fos. AP-1 binding to DNA can activate both kinases and phosphatases.
The principal function of α2 AdRs is reducing the magnitude of NE neurotransmission
across the synapse by virtue of neurotransmitter reuptake by receptors localized on presyn-
aptic membranes of S neurons (figure 1.16). Currently, it is recognized that α2 AdRs are also
found on postsynaptic membranes, where they inhibit cAMP formation that results from
stimulation of beta ARs. Activation of α2 AdRs and of Giα and Goα signaling is respon-
sible for the inhibition by catecholamines of insulin secretion (Lang et al., 1995). α2 AdR
inhibit cAMP formation through actions of the αi subunit of the Gi regulatory protein. The
αi subunit dissociates from its G protein and blocks cAMP formation by binding to the beta
and gamma subunits of the Gs regulatory proteins (see figure 1.15).
MRs of the PS nerves share both the seven-transmembrane protein configuration and
association with a G protein (see figure 1.11c) with beta AdRs. MRs can be classified into
five categories (M1-M5). They trigger intracellular signaling by several different cascades
(figure 1.17).
As is the case with α2 AdR, M2 and M4 receptors are coupled to Gi regulatory proteins
and inhibit the formation of cAMP (right side of figure 1.17) in the heart muscle, where M2
receptors predominate, and in other tissues. M1, M3, and M5 receptors activate a Gq regulatory
protein. As is the case with α1 AdR, they increase intracellular Ca concentration through
the action of the second messengers IP3 and DAG (left side of figure 1.14). In this way they
24 Advanced Exercise Endocrinology


Sympathetic NE
β
M
α2 E
Effector
α1 cell

Ach M
Vagal

Figure 1.16 Location of α2 AdR on presynaptic membranes, where they predominantly mediate catecholamine
reuptake from the synapse.
Reprinted, by permission from P.L. Wisler, F.L. Green,E4517/Borer/fig.1.18/400207/TB/R1
and A.M. Watanabe, 1992, Cardiovascular adrenergic and muscarinic cholinergic
receptors. In Cardiology, Vol. 1, edited by W.W. Parmley and K. Chatterjee (Philadelphia: PA: JB Lippincott), 9.

Acetylcholine

Ca2+ mAch Na+


γ γ
PLC GC
DAG + IP3 PIP2 β Gqα β Giα AC


GTP GTP K+
PKC ATP

PKC-catalyzed cGMP
protein

phosphorylations –
SR PDE cAMP –

Ca2+
ATP Ca2+ 5´- AMP
Ca2+
R-R

Ca2+

Figure 1.17 Signaling pathways of MR.

precipitate bronchoconstriction, and the metacholine (an Ach agonist) administration is


used to identify exercise-induced and other forms of asthma. MRs can also inhibit phos-
E4517/Borer/fig.1.17/400208/TB/R2-alw
phorylation of the protein on voltage-sensitive Ca channels in the heart plasma membrane
and thus block activation of the heart by β1 AdR stimulation. Additional actions in the heart
muscle involve activation by MR of cGMP and the enzyme PDE that inactivates cAMP into
5' AMP (figure 1.17). Finally, MR can hyperpolarize the atrial muscle through their action
on K channel by way of another type of Gi regulatory protein.
Unique Properties of Endocrine and Autonomic Messengers 25

Intracellular and Nuclear Receptors and Their Mechanisms


of Action
Intracellular receptors (see figure 1.11e) are proteins that belong to the nuclear receptor
superfamily. They are localized either in the cytoplasm (steroid hormones) or in the nucleus
(thyroid hormone receptor, TR, and vitamin D3 receptor, VDR). This family also includes
peroxisome proliferator-activated receptors (PPARs) and some other metabolic receptors.
Intracellular and nuclear receptors contain three domains: an HVD or N-terminal domain
(NTD) at the amino terminus of the protein, a DBD in the center, and usually two HBDs
toward the carboxy terminus of the receptor. The HVD differs in the amino acid length
and composition to convey specificity in hormone recognition of the receptors. This is
illustrated in figure 1.18 for the steroid receptor superfamily that includes receptors for
glucocorticoids (GR), mineralocorticoid aldosterone (MR), androgens (AR), estrogen (ERα
and ERβ), progesterone (PR), and is also true for retinoic acid (RAR), thyroid hormones,
and vitamin D3 receptors. Amino acid numbers and composition in the other two domains
are similar and highly conserved among different steroid receptors (figure 1.18). The DBD
contains two short amino acid loops (zinc fingers) involved in binding with DNA. Part of
one zinc finger conveys specificity to hormone–DNA binding. One HBD is the hydrophobic
binding pocket for the lipid hormone and the other is involved in transcriptional activation.
Intracellular and nuclear hormone receptors predominantly exert genomic action dis-
cussed in this chapter, but have more recently been found to also act through nongenomic
membrane receptors (discussed in chapter 7).
The most common transduction mechanisms of intracellular and nuclear hormone recep-
tors are acting as nuclear TFs after binding to the hormone responsive element (HRE) of
the particular hormone-responsive genes or activating several rapid, nonnuclear genomic
signaling pathways that involve plasma membrane-associated receptors. HRE is a region of
DNA that is able to bind with the zinc-finger portion of the DBD of intracellular hormone
receptors or with TF derived from membrane receptors. HREs are located close to the
promoter region of a specific gene that is responsive to a given hormone. Binding of TF to
HRE initiates or suppresses gene transcription. This process includes an interaction of TF

NH2 HBD1
HVD DBD HBD2
COOH

AR 100 100 100 919


PR < 15 80 53 935
GR < 15 77 50 777
MR < 15 77 52 984
ERα < 15 51 20 595
ERβ < 15 50 22 530

Figure 1.18 Structural homology (percentage) of the HVD or NTD, DBD, and HBD of several steroid nuclear
receptors relative to the AR.
Adapted from TEXTBOOK OF ENDOCRINE PHYSIOLOGY, THIRD EDITION, edited by James E. Griffin and Sergio R. Ojeda, copyright © 1996
by Oxford University Press, Inc. Used by permission of Oxford University Press, Inc. Data at the bottom of the figure from Li and Al-Azzawi
2009.
E4517/Borer/fig.1.18/400209/TB/R2-alw
26 Advanced Exercise Endocrinology

with coactivator or corepressor molecules, also called coregulators or comodulators. Peroxi-


some proliferator-activated receptor coactivator-1α (PGC-1α) plays an important modulating
role in gene expression. Coactivators have histone acetyltransferase activity, which relaxes
chromatin coiling and facilitates gene transcription. Corepressors block coactivator action.
Coactivator and corepressor action in selective modulation of gene expression by acetylation
of histones is characterized as epigenetic.
Genomic action involves steroid hormone binding to its cytoplasmic receptor, which
is, in the absence of the hormone, inactivated by heat-shock proteins (HSPs). In the case
of sex steroid hormones, Hsp90 and associated cochaperone molecules Hsp40 and Hsp70
keep the cytoplasmic hormone receptor inactive (Li & Al-Azzawi, 2009). Upon dissocia-
tion of Hsp90, the steroid hormone–receptor complex dimerizes and is translocated to the
nucleus. There it binds to the steroid response element (SRE) on the chromatin under the
influence of coregulators or comodulators that facilitate or inhibit gene transcription either
through interaction with RNA polymerase or through influencing histone and chromatic
remodeling (figure 1.19).
Nuclear receptors such as unbound TR, receptors for lipid-soluble proteins, and PPAR
are located bound to DNA in a repressed form. After binding with hormones, these recep-
tors cause relaxation of histone coiling and facilitate gene transcription.

PROPERTIES OF HORMONE–RECEPTOR
INTERACTIONS
Hormone–receptor reactions conform to some biochemical laws that control the affinity
with which the hormones bind to their receptors, the specificity of recognition of a particular
hormone receptor, binding properties of the hormone–receptor complex as a function of
hormone and hormone receptor abundance, and the rate of hormone degradation.

Androgen Heat-shock protein

Androgen receptor

P P
HR binding HRs dimerize

Co-activators

Nucleus P P
Gene
transcription
DNA

Hormone responsive element

Figure 1.19 Binding of testosterone (T) with AR, translocation to the nucleus, dimerization, phosphorylation,
E4517/Borer/fig.1.19/400210/TB/R2-alw
and binding to the androgen responsive element (ARE) on the chromatin of the T-sensitive gene.
Unique Properties of Endocrine and Autonomic Messengers 27

Affinity and Sensitivity of Hormone–Receptor Interactions


Hormone action depends on the affinity of each hormone for its receptor, specificity of the
dynamic hormone and receptor interaction, and hormone concentration. To assess specific
binding (Bs) of a hormone (H) to a receptor (R), it is necessary to subtract hormone binding
to nonreceptor substrates (BNS) (figure 1.20b) from total hormone–receptor (HR) binding
(BT). Affinity (derived from the hormone–receptor dissociation constant Kd) is the concen-
tration of hormone at which 50% of the available receptors become bound to the receptor.
Therefore, hormone affinity is inversely proportional to the free hormone concentration
at which 50% of binding occurs (figure 1.20a). This underscores the inverse relationship
between hormone concentration and the sensitivity of hormone–receptor interaction.
Hormone–receptor affinity is often assessed with a Scatchard plot, named after American
chemist George Scatchard (figure 1.20c).
The full biological effect of a hormone sometimes requires binding all receptor sites
in a cell with the hormone. This usually occurs under the conditions of reduced receptor
number. More commonly, the full biological effect is produced when only a fraction of avail-
able receptors are activated by a hormone. Insulin action in the liver and in adipose tissue
requires activation of fewer than 5% of available receptors. The remaining spare receptors
contribute to the sensitivity of hormone–receptor interaction (figure 1.20d).
Hormones circulate at very low concentrations (10 -7 to 10 -12 M). As outlined above, high
hormone concentrations affect H–R binding, sensitivity of such binding, and the biological
effect of the hormone. When receptor availability is reduced due to metabolic or pathologi-
cal events or interference with hormone signaling, higher concentrations of hormone are
needed to activate the requisite minimum number of receptors required for the full biological
response. A partial biological response is elicited when the number of receptors decreases
below the minimum number required for a full biological effect (figure 1.20d).
High hormone concentrations result in receptor desensitization or downregulation.
Desensitization is in part the result of endocytosis and subsequent degradation of clustered
unbound, along with hormone-bound, receptors in the clathrin-coated pits. Desensitization
may also result from recruitment of proteins β-arrestins by some activated hormone receptors
to single out the receptors for degradation. A decline in the number of receptors reduces
the sensitivity of H–R binding because a greater proportion of total available receptors
ends up being bound by the hormone (figure 1.20e). Maximal biological effect and tissue
sensitivity to hormone action depends on the number of activated receptors (which can be
altered by the physiological state of the organism) as well as the ease of transmission of the
hormone signal. For instance, impaired IR signaling is found in obese individuals and those
with T2D while prevailing high plasma insulin concentrations lead to IR downregulation.

Specificity of Hormone–Receptor Interaction


The specificity of hormone–receptor interaction depends in part on the stereospecific
match in their molecular structure. Specificity of hormone action can also be conveyed by
different effector mechanisms that respond to the same hormone. A single hormone may
activate different enzymatic reactions in the tissues in which its receptors are expressed.
Thus, beta AdRs cause contraction of the heart muscle but relaxation of muscles in blood
vessels, lipolysis in the adipose tissue, and glycogenolysis in the liver. On the other hand,
a given hormone may activate more than one type of receptor and elicit multiple biological
effects that are specific to this hormone. For instance, during exercise, activation of S nerves
5

BT
4

100 BNS
N = Bmax 3
Saturation
B
Binding (%)

2 BS
50 Specific binding

Kd (affinity)
0 0
a Free H concentration 0 1 2 3 4 5
b F

1
100% N
Percentage of available

Spare receptors
– Ka (–
E4517/Borer/Fig affinity)
1.22a/400211/alw/pulled-R1
receptors

E4517/Borer/FigRequired
1.20b/400212/alw/pulled-R2
occupancy 75% N
0.5
B/F

50% N
N
25% N
0
c B Kd
d log H

Biological response
100
Biological response (%)

E4517/Borer/Fig 1.20d/400214/alw/pulled-R2
E4517/Borer/Fig 1.22c/400213/alw/pulled-R1
100% N
75% N
50% N
50
25% N

0
e log H

Figure 1.20 Characteristics of hormone–receptor binding. (a) H–R saturation plot displaying affinity of H–R interac-
tion as concentration of free hormone at which H binding is half maximal. (b) Specific binding is total H–R binding less
nonspecific binding of the free hormone (marked F) to non-R substrates. (c) Scatchard plot. (d) Binding curves with variable
receptor availability and a requirement for 50% of total available R occupancy for maximal biological effect. Spare receptors
in excess of 50% of total available E4517/Borer/Fig
R occupancy increase 1.20e/400215/alw/pulled-R2
the sensitivity of H–R interaction. (e) As the number of spare
R decreases from 100% to 50% of total receptors (N), higher concentrations of H (marked F) are needed to achieve the
biological effect underscoring a decline in the affinity of H–R interaction. In a Scatchard plot (c), the affinity or association
constant (Ka) is obtained from a regression of the ratio of bound H over free H against bound H. It is calculated from the
formula Ka = [R − H]/[R][H], with the brackets denoting concentrations.
Figures a-d from Springer and Chapman Hall, Hormones: From molecules to disease, 1990, page 59, Hormones: A complex communication network,
E.E. Baulieu, edited by E.E. Baulieu and P.A. Kelly, figure 1.57, with kind permission from Springer Science+Business Media B.V.

28
Unique Properties of Endocrine and Autonomic Messengers 29

precedes stimulation of adrenal E release. Because the S neurotransmitter NE predominantly


stimulates alpha AdR, the early manifestation specific for S activation during exercise is
vasoconstriction, leading to redistribution of blood from viscera to the muscle, sweating, and
pupillary dilatation. Higher-intensity and prolonged exercise facilitates adrenal E release and
its predominant stimulation of beta AdRs. This specifically activates the heart, mobilizes
metabolic fuels, and stimulates metabolism.

SUMMARY
This chapter outlines common features of autonomic and hormonal chemical messengers.
It provides some detail about the structure and function of the ANS and the sources and
chemical characteristics of hormones and related messenger systems. Control of hormone
and ANS messenger release and message transduction pathways is discussed according
to characteristics of messenger receptors. Biochemical properties of H–R binding, trans-
port, and degradation are also outlined. These features, and the features of the ANS and
endocrine messengers, are presented in the context of the roles they play in the activation
of functions that support physical activity or functions that facilitate vegetative functions
during recovery from exercise.
This page intentionally left blank.
H APTER

2
C

Activation of
Nonhormonal
Signaling During
Exercise
T
he essential feature of actions of the endocrine and autonomic systems is message
transduction as a result of the binding of their messengers with the receptors on
cell membranes or inside the cells and nuclei. Most of the endocrine and auto-
nomic consequences of exercise discussed in this text represent endocrine and autonomic
actions resulting from messenger–receptor binding. However, it should be emphasized that
exercise-associated signal transduction that is usually initiated by endocrine and autonomic
messengers does not require that the messengers bind with the receptors. Most tissues and
organs possess the ability to initiate enzymatic signal transduction cascades and to activate
gene expression in response to local neural, ionic, mechanical, oxidative, and bioenergetic
challenges of exercise that are alternative and redundant to endocrine and autonomic mes-
saging. Usually, signaling induced by events in a tissue activates the transduction pathway
distal to receptors available for hormone–receptor interactions. This allows for a greater
richness of biological outcomes by adding to the endocrine and autonomic modulation
of contraction-associated events. In some cases, separate transduction pathways may be
activated by tissue-initiated events and by hormonal action, as is the case with the stimula-
tion of muscle glucose uptake that is insulin independent during exercise but is otherwise
insulin dependent (Cartee & Wojtaszewski, 2007; Richter et al., 1985, 1989). Although not
all exercise-associated endogenous signaling has been examined under the conditions that
definitively exclude endocrine mediation, existence of nonhormonal signaling is corrobo-
rated in animals with blocked hormone action, in in-vitro cellular and tissue models, and
in humans performing single-limb exercise.

31
32 Advanced Exercise Endocrinology

Muscle contraction represents a confluence of electrical, ionic, mechanical, metabolic, and


oxidative events, each of which can influence enzyme activity and gene expression in muscle
and bone and other organs affected by exercise-associated stimuli. In addition, specific
parameters of exercise such as intensity, mechanical loading, and duration recruit different
types of motor units and affect the relative contribution of electrical, ionic, mechanical, and
metabolic changes. These stimuli sometimes act independently and at other times act in
concert, or influence each other’s actions. In addition, these stimuli may act independently
of, or in conjunction with, hormonal stimulation.
The role of nonhormonal stimuli in signal transduction and gene expression is described
in the context of critical events triggered by exercise: motoneuron discharges and electrostatic
fields associated with muscle contraction, intracellular Ca release in the muscle induced by
muscle contraction, mechanical strain and electrostatic fields generated by muscle contrac-
tion and fluid shear in the bone, exercise-induced decrease in availability of cellular energy,
and oxidative production of free radicals.

SIGNAL TRANSDUCTION BY NEURONAL


AND ELECTROSTATIC EVENTS IN EXERCISE
Muscle contraction entails coupling of motoneuron depolarization of sarcolemma and
of the T-tubule system through the release of Ach, with contraction of actin and myosin.
Coupling is achieved through the release of Ca from the SR into the cytoplasm and the
entry of additional Ca from outside of the muscle cell through Ca channels. It is not easy to
dissociate motoneuronal release of a chemical messenger from the ionic events producing
excitation–contraction coupling, but sufficient evidence implicates their separate actions
in initiation of signal transduction and gene transcription. For instance, cross-innervation
experiments and delivery of tonic low-frequency electrical stimulation that mimics the firing
pattern of slow oxidative motoneurons upregulate slow oxidative muscle fiber genes and
genes essential for utilization of skeletal muscle lipid, such as Lpl (lipoprotein lipase), Cd36
(fatty acid transporter CD36), beta oxidation enzyme Had (enzyme 3-hydroxyacyl-CoA
dehydrogenase), and Cs (citrate synthase) (Schiaffino & Reggiani, 1996). Low-frequency
electrical stimulation also increases expression of genes that are necessary for glucose
uptake and storage, such as Glut4 (for glucose transporter GLUT-4) and Hk2 (for glucose
phosphorylating enzyme hexokinase), and modulates conversion of fast-twitch Type II
muscle fibers to the oxidative Type I fibers (Long & Zierath, 2008). Motoneuron activa-
tion, electrical stimulation of the muscle, and exercise all activate AMPK (AMP-activated
protein kinase) through phosphorylation of threonine on the alpha subunit of the enzyme
(Jorgensen & Rose, 2008).
Motoneuron discharges generate high-frequency electrical pulses. Bone deformation is
also associated with generation of electrostatic fields (Bassett et al., 1962; Fukada & Yasuda,
1957). Electrostatic fields are generated by deformation of highly structured hydroxyapa-
tite crystals in bone that have piezoelectric properties. Thus, compression of fluid in bone
canaliculi surrounding the network of mechanotransductive osteocytes during mechanical
loading induces streaming electrical potentials (Pienkowski & Pollack, 1983). Pulsing electric
currents have been found to facilitate healing of bone fractures. Even pulsing electromag-
netic fields (21 of 0.4 mT pulses in 260 μs with a 67 ms pause, performed 10 times in a day)
induce phosphorylation of the anabolic mTOR enzyme in preosteoblasts and fibroblasts in
the absence of growth factors in the cell medium. These pulsed electromagnetic fields also
activate the PI3K–mTOR signaling pathway distal to where this pathway is activated by
growth factors (Patterson et al., 2006; figure 2.1).
Activation of Nonhormonal Signaling During Exercise 33

Growth factor

p85 p110 PI3K


Electromagnetic field

PDK-1

TSC1
GSK3 Akt
TSC2

mTOR Rheb

eIF2B p70S6K 4E-BP-1


eEF2K PHAS-1

eIF2 eIF4B rpS6 eEF2 eIF4E

Protein translation

Figure 2.1 Activation of PI3K–mTOR signaling by electromagnetic fields distal to insulin or growth factor recep-
tors. The mTOR signal transduction pathway links insulin, mitogenic growth factor, or electromagnetic stimulation
and nutritional cellular energy level to eukaryotic translation initiation factors (eIFs) necessary for protein synthesis,
E4517/Borer/fig.2.1/401465/TB/R3-kh
cell cycle progression, and cell proliferation. This anabolic pathway includes an enzymatic cascade consisting
of PI3K (composed of catalytic p85 and p110 regulatory components), 3-phosphoinositide-dependent protein
kinase-1 (PDK-1), Akt, and mammalian target of rapamycin (mTOR). mTOR activation also involves inhibition by
TSC1/2 (tuberous sclerosis complex 1,2) of Rheb (Ras homolog enriched in brain). Akt promotes glycogen synthesis
through inhibition of GSK3 (glycogen synthase kinase 3) and translation of protein mRNA on the ribosomes through
activation of eukaryotic translation initiation and elongation factors (eIFs and eEFs, respectively).
Adapted from Patterson et al. 2006.

INTRACELLULAR CALCIUM RELEASE AS A TRIGGER


OF SIGNAL TRANSDUCTION
The chief event that couples motoneuronal electrical excitation with muscle contraction is
the release of calcium (Ca2+) out of the SR and entry of Ca from extracellular space through
Ca channels. Increases in cytosolic Ca2+ concentration during muscle contraction activate
several signaling proteins. These include CaM, the kinases CaMK and CaMK II and IV,
and CaMKK (CaM kinase kinase). CaMKK activates AMPK in response to Ca release in
general and in association with caffeine ingestion in particular.

Calmodulin
Calmodulin (CaM) serves as an intracellular receptor for Ca2+ ions. CaM is a delta compo-
nent of a hexadecameric (αβγδ)4 enzyme phosphorylase kinase, that activates phosphorylase
and breakdown of muscle glycogen. CaM binds four Ca2+ ions and then undergoes a con-
formational change that greatly increases its affinity for various target proteins involved in
control of biochemical pathways, including protein kinases and phosphatases such as Cn,
adenyl cyclase (C), cytoskeletal proteins, ion channels, and Ca2+ pumps. Ca activates both
the CaM component of the kinase and troponin I, with which it shares high similarity of
34 Advanced Exercise Endocrinology

amino acid sequence (Jeyasingham Glycogenolysis


et al., 2007). Thus, Ca2+ binding with
Phosphorylase
both proteins simultaneously triggers α
kinase
both muscle contraction and glycoge- β γ
nolysis (figure 2.2). δ = CaM
Another role of CaM is to facili-
Sarcolemma Ca2+
tate dilatation of blood vessels in the depolarization
exercising muscle. An increase in Ca
concentration activates this enzyme’s Troponin + CaM
association with eNOS, a constitutive
P450 cyclo-oxygenase consisting of a Muscle contraction
reductase and an oxygenase domain
in the blood vessel endothelia (figure Figure 2.2 Simultaneous binding of Ca2+ to the troponin
2.3). Activation of CaM initiates nitric and to the delta subunit (CaM) of phosphorylase kinase
oxide (NO) synthesis by eNOS. The triggers contraction and glycogenolysis and thus couples
E4517/Borer/fig.2.2/401466/TB/R1
eNOS reductase contains FAD (flavine excitation–contraction to energy production in the muscle.
adenine dinucleotide), which reduces Adapted from Journal of Molecular Biology, Vol. 377, M.D. Jeyasingham et
al., “Structural evidence for co-evolution of the regulation of contraction
NADPH + (nicotinamide adenine and energy production in skeletal muscle,” pgs. 623-629, copyright 2007,
dinucleotide phosphate) and passes its with permission from Elsevier.

α1
Ca2+ γ
R PIP2 DAG
β G α PLC
q
PKC
Ca2+ β γ

IP3 Gcα
Ca2+ IP3R

GC
CaM Ca2+
L-arginine
R-R
eNOS BH4 O2
R
mACH

CADPr ER/SR cGMP


NADPH+ FAD FMN H H2O
L-citrulline H– NAD
+ C PDE Ca2+
PKG

NO GC
cGMP

Figure 2.3 Mediation by cytosolic Ca of NO synthesis. NO (bottom left) is a product of the activation of endothelial
NO synthetase (eNOS) by CaM. The enzyme has a heme group (H) and a BH4. Its reductase (flavine mononucleotide,
FMN) component reduces E4517/Borer/fig.2.3/401467/TB/R2-alw
NADPH and converts L-arginine to citrulline; NO and water are byproducts. NO can
activate cytoplasmic GC to produce cGMP that, in turn, stimulates Ca release from the SR through a bifunctional
H–C enzyme that can act as either a C or hydrolase (H-). NO also can be produced by hormonal activation α1 AdR,
Gqα, and PLC that increases intracellular Ca via a message-transduction pathway that includes IP3 and DAG (see
also figure 1.14). IP3 increases Ca release from the SR after binding to IP3R.
Activation of Nonhormonal Signaling During Exercise 35

electrons on to oxidase. The oxidase (O) has a heme group and is associated with a BH4 (H4
biopterin cofactor). It oxidizes L-arginine to L-citrulline, reduces molecular oxygen to water,
and releases NO. NO in turn produces vasodilation in smooth muscles by closure of L-type
Ca2+ channels via cGMP and protein kinase G (PKG). It also stimulates cytosolic Ca release
from the endoplasmic reticulum by activating cytosolic guanylyl cyclase (GC) to produce
cGMP for stimulation of a bifunctional enzyme that can act as either C or hydrolase (H)
(figure 2.3). In the absence of L-arginine and the cofactor, NOS generates the free radicals
superoxide and hydrogen peroxide.

Calcineurin
Calcineurin (Cn) is a cytoplasmic heterodimeric serine-threonine protein phosphatase that
is usually activated by CaM. It acts as a Ca sensor that couples neuronal signals and Ca
release to activation of lipid oxidation, insulin sensitivity, insulin-induced glucose uptake,
upregulation of metabolic genes, mitochondrial biogenesis, and slow-fiber genetic transcrip-
tional programs (Long & Zierath, 2008; Sakuma & Yamaguchi, 2010).
The low, sustained pattern of Ca release by type I motoneurons that activate slow-twitch
oxidative fibers, but not the transient burst of cytosolic Ca2+ triggered by fast-twitch fibers,
selectively activates Cn-mediated signaling that is necessary for conversion of muscle fiber
from fast glycolytic to slow oxidative type and for activation of oxidative metabolism genes.
Upon binding to Ca2+, Cn dephosphorylates the Nfatc TF and mediates its translocation to the
nucleus. There, Nfat (nuclear factor of activated T cells), along with the myogenic regulatory
factors (MRFs) myogenin, MEF2 (myocyte enhancer factor 2), MyoD, and TF GATA-2 BP,
stimulates transcription of oxidative and mitochondrial genes (Long & Zierath, 2008) and
satellite cell differentiation during recovery from muscle damage (Sakuma & Yamaguchi,
2010). Cn activates the lipid metabolic genes Lpl, Cd36, Cpt (for the rate-limiting enzyme
carnitine palmitoyl transferase I that transports long-chain fatty acids into the mitochondrion)
and enzymes for beta oxidation of fatty acids in the mitochondria such as Had. Cn also
increases protein content of transcription regulators such as PGC-1α, PPARα, and PPARδ.
However, Cn is not required for the expression of PGC-1α in weight-bearing skeletal muscles
(Benzet et al., 2012). Overexpression of Cn in fast-twitch muscles increases IR number,
activity of Akt, and glucose uptake by skeletal muscle through GLUT-4 translocation to the
cell membrane. This effect appears to be mediated by changes in gene expression because it
is not dependent on the PI3K pathway. Cn also activates immune response by upregulating
the expression of the cytokines IL-4 and IL-2 that, in turn, activate T-helper cell immune
function (Yamashita et al., 2000). This activity is controlled by immunosuppressant drugs
cyclosporine A and tacrolimus. Cn also is a trigger for osteoclast activation in the bone in
a pathway that is initiated by the binding of RANK ligand (receptor activator of nuclear
factor κ B ligand or RANKL) to RANK (Hwang & Putney, 2011).

PGC-1α
Activation of oxidative enzymes and conversion of Type II glycolytic to Type I oxidative
fibers by Cn requires facilitation by PGC-1α, a transduction coactivator that is expressed in
weight-bearing oxidative muscles. PGC-1α coactivates transcription of genes for slow Type
I muscle proteins troponin I, slow myosin, and myoglobin by coactivating the gene Mef2
for the TF MEF2. It also induces mitochondrial biogenesis and expression of cytochrome
genes in the skeletal and cardiac muscle and gluconeogenesis in the liver by coactivating
the gene Nrf1 for the TF NRF1 (Lin et al., 2002; figure 2.4). PGC-1α also induces adaptive
36 Advanced Exercise Endocrinology

Myocyte

Ca2+ Calmodulin

Exercise Calcineurin ATP


Hypoxia CaMK depletion
Motoneuron
activation
White adipocyte Nfatc

Nucleus PGC-1α Nfat

UCP1 Hormone
irisin
fndc5 Nrf Mef2 AMPK

“Beige” adipocyte
thermogenesis
Mitochondrial Oxidative fiber
biogenesis biogenesis
Oxidative
phosphorylation

Figure 2.4 Stimulation by Ca-binding enzymes and PGC-1α of transcription of genes for slow-twitch fibers,
mitochondrial biogenesis, oxidative enzymes, and adaptive thermogenesis. Nfat and Mef2 (right) are the genes for
the TF for slow-fiber genetic transcription, Nrf1 is a gene for the TF inducing mitochondrial biogenesis and oxidative
metabolism, and fndc5 is the gene for the E4517/Borer/fig.2.4/401468/TB/R5-alw
expression of the hormone irisin that converts the white adipocytes to
more thermogenic beige adipocytes expressing UCP1.
Adapted from Nature Vol. 418, J. Lin et al., “Transcriptional co-activator PGC-1alpha drives the formation of slow-twitch muscle fibres,” pgs.
797-801, copyright 2002, with permission from Elsevier.

thermogenesis through coactivation of the fndc5 gene that produces the hormone irisin. Irisin
acts on white adipose tissue (WAT) to upregulate uncoupling protein 1 (UCP1) expression
and convert WAT to more thermogenic beige fat (Boström et al., 2012).

SIGNAL TRANSDUCTION OF MECHANICAL STRAIN,


VIBRATION, AND FLUID SHEAR
Exercise produces mechanical strain, compression, and torsion, and these produce deforma-
tion of cells or their molecular components that can activate enzyme activity and genetic
transcriptional machinery. In tissues habitually subjected to mechanical forces, proteins
in the extracellular matrix are physically linked to the actin cytoskeleton through proteins
called integrins and other linker proteins that adhere to cell membrane. Forces generated
either by intracellular actin (as is the case with concentric and eccentric muscle contractions)
or by external compression or torsion (as in mechanical loading of bones) are transmitted
through these protein links and activate a wide range of signaling pathways (Schwartz, 2010).

Signal Transduction in Response to Muscle Mechanical Loading


Important transductive stimuli for hypertrophic muscle growth are stretch and tension in
skeletal muscle cells and stretch in cardiomyocytes. Four decades ago Alfred Goldberg
induced the same relative magnitude of hypertrophic growth in overloaded muscles of
Activation of Nonhormonal Signaling During Exercise 37

hypophysectomized rodents as in pituitary-intact animals by ablating synergist leg muscles,


which demonstrated that transcription and translation of overloaded contractile muscle
proteins did not depend on GH (Goldberg, 1967). About a quarter of a century later, nonhor-
monal induction of muscle hypertrophy was again demonstrated when similarly overloaded
muscles of hypophysectomized rats increased the expression of genes for IGF-I and IGF-II
(DeVol et al., 1990; figure 2.5).
Mechanically overloaded muscles therefore can initiate transcription of genes for locally
acting growth factors that mediate protein synthesis and hypertrophic growth without a direct
influence of circulating anabolic hormones. Similarly, high-resistance exercise in humans
and passive stretch or tension of animal muscles isolated from hormonal influences increase
expression of genes for myogenic growth factors that stimulate protein translation largely
through activation of mTOR and ribosomal p70s6 kinase (Baar et al., 1999; Hornberger &
Esser, 2004; Hornberger et al., 2004). The multiaxial stretch of isolated myotubes activates
MAPK signaling that depends on the actin cytoskeleton for transduction of the mechanical
stimulus (Hornberger et al., 2005). Differential patterns of stretch and tension in eccentric
and concentric exercise elicit expression of different genes as early as 3 h after exercise
(Kostek et al., 2007).
How mechanical overload activates the molecular mechanisms for the intramuscular
transcription of genes for growth factors in the absence of circulating GH and other anabolic

IGF-I
15

10

5
Relative hybridization

0
Days: 2 4 8 2 4 8 2 4 8 2 4 8

IGF-II
8

0
Days: 2 4 8 2 4 8 2 4 8 2 4 8
Control Hypertrophy Control Hypertrophy
Soleus Plantaris

Figure 2.5 Expression of mRNA for the growth factors IGF-I (top) and IGF-II (bottom) in the soleus and plan-
taris muscles of hypophysectomized rats after ablation of synergist muscles. Sham-operated animals show no
hypertrophy or increase in mRNA compared with rats with ablated synergist muscles.
E4517/Borer/Fig2.5/401469,470/alw/R2
Reprinted, by permission, D.L. DeVol et al., 1990, “Activation of insulin-like growth factor gene expression during work-induced skeletal
muscle growth,” American Journal Physiology 259: E89-E95.
Note:
Image size
increased to 1/2-2
to fit image
38 Advanced Exercise Endocrinology

hormones is emerging from other lines of research (Favier et al., 2008). Synthesis of new
contractile proteins in response to mechanical loading requires incorporation of satellite
cell DNA content into the existing myofibers. IGF-I and myostatin are the key MRFs that
regulate satellite cell number and activity by initiating or repressing muscle hypertrophy,
respectively (figure 2.6). Muscle overload upregulates IGF-IEa, a hepatic-like isoform of
the hormone, and IGF-IEc, a human muscle isoform that is also called mechano growth
factor (MGFs). Muscle overloading also upregulates genes for additional muscle-specific
MRFs such as MyoD, myogenin, and MRF4. IGF-I activates the PI3K and Akt enzymes,
and they in turn stimulate protein synthesis (figure 2.1) and inhibit expression of degradation
enzymes and proteolysis mediated by the Forkhead box TF FOXO. Myostatin represses the
expression of MRFs and activates FOXO (figure 2.6). The PGC-1α gene coactivator also
inhibits FOXO and is in turn suppressed by myostatin.
IGF-I causes hypertrophy by activating PI3K–Akt and by affecting two of its downstream
pathways. It activates the anabolic enzyme mTOR pathway and inhibits by phosphoryla-
tion the antianabolic GSK3 (Matheny et al., 2010; Rommel et al., 2001; figure 2.7). PI3K
phosphorylates and activates Akt on two sites (Ser473 and Thr308) to turn on mTOR,
although this effect can be caused by mechanical stimuli independently of PI3K activation.
Akt signaling is necessary for IGF-I-induced hypertrophy because a blockade of PI3 kinase
by its blocker LY294002 abolishes it. Hypertrophy in response to IGF-I is diminished but
not completely abolished by a blockade of the mTOR pathway by rapamycin. Inhibitory
regulators of translation are the Akt–TSC2–Rheb–mTOR and GSK3 pathways (figure 2.7).
Activation of TSC by AMPK (figure 2.8) may explain why endurance training does not lead
to substantial muscle hypertrophy. GSK3, in turn, inhibits the eIF2B translation initiation
factor. Upregulation of a dominant negative form of GSK3β leads to significant myotube
hypertrophy. The hypertropic effects of IGF-I, therefore, appear to depend on activation by
Akt and by mechanical loading of mTOR and on inhibition of TSC and GSK3β pathways.

IGF-I Myostatin

R R

p85 p110 PI3K

PDK-1

Akt

mTOR FOXO-P

Nucleus
FOXO
Protein Atrogenes
synthesis proteolysis

Figure 2.6 Regulation of protein synthesis in the muscle in response to stretch. Muscle stretch upregulates
IGF-I, which in turn promotes protein synthesis through the PI3K–Akt–mTOR cascade. Akt blocks the activation
of FOXO by myostatin. Myostatin inhibits muscle hypertrophy.
E4517/Borer/fig.2.6/401471/TB/R3-kh
Activation of Nonhormonal Signaling During Exercise 39

Ca2+
Growth factor

p85 p110 PI3K Ca2+

PDK-1

TSC1 Calcineurin
GSK3 Akt CAMK
TSC2

mTOR Rheb

Nfatc-P Nfat
eIF2B p70S6K 4E-BP-1 Nucleus
eEF2K PHAS-1
Nfat
eIF2 eIF4B rpS6 eEF2 eIF4E

Gene transcription
Type I myogenesis
Protein translation

Figure 2.7 IGF-I signaling involved in muscle hypertrophy requires activation of PI3K and Akt. Akt, in turn,
activates mTOR and inhibits GSK3β through signaling pathways shown also in figure 2.1. The endpoint is stimulation
of eIF2, eIF4B, eIF4E, andE4517/Borer/fig.2.7/401472/TB/R4-kh
eEF2. mTOR facilitates dephosphorylation of Nfat-P, its translocation to the nucleus,
and transcription of slow-twitch oxidative muscle fibers. Cn signaling, on the other hand, does not facilitate muscle
hypertrophy.

Hypertrophic response to mechanical loading requires translation of contractile protein


genes. Translation entails replication of a genetic blueprint through synthesis of mRNA and
protein synthesis on the ribosomes. Ribosomal translation requires three steps: initiation of
the protein chain, its elongation, and its release into the sarcoplasm. Each step is facilitated
by, respectively, the eukaryotic initiation, elongation, and release factors (eIFs, eEFs, and
eRFs) that associate with the ribosomes to regulate mRNA translation. mTOR increases
phosphorylation of p70S6 kinase and releases eIF4E from its repressor BP-1 (4E-BP-1; see
figures 2.1, 2.7, and 2.8). Activation of p70S6 kinase also phosphorylates the ribosomal protein
S6, which, in turn, enhances the translation of ribosomal proteins and elongation factors.
Downstream of 4E-BP-1, eIF4E is a rate-limiting facilitatory initiation factor because it
binds to the cap structure at the 5' end of mRNA transcripts (Urso, 2009). IGF-I also inhibits
expression of degradation enzymes and proteolysis mediated by FOXO. Myostatin, on the
other hand, represses the expression of MRFs and activates FOXO (figure 2.6). PGC-1α
also inhibits FOXO and is in turn suppressed by myostatin.
Protein status in the cell depends on a balance between protein synthesis and degrada-
tion. Various factors may influence endogenous control of protein degradation and produce
muscle atrophy as a result of disuse, sarcopenia, or apoptosis during aging (Favier et al.,
2008). The roles of FOXO TF, the proteolytic muscle-specific ubiquitin ligases, atrogin 1
and MuRF, and myostatin in muscle protein degradation in response to different forms of
unloading are not fully understood (Jespersen et al., 2011; Urso, 2009). Protein degradation
is the result of the suppression of the PI3K–Akt–mTOR pathway and decreased phosphoryla-
tion of FOXO. This leads to the translocation of FOXO from the cytoplasm to the nucleus,
40 Advanced Exercise Endocrinology

Growth factor

p85 p110 PI3K

PDK-1
ERK1/2

TSC1
GSK3 Akt FAK
TSC2
AMPK
mTOR Rheb

eIF2B p70S6K 4E-BP-1


eEF2K PHAS-1

eIF2 eIF4B rpS6 eEF2 eIF4E

Protein translation

Figure 2.8 Mechanical overload increases the activity of PI3K and phosphorylation by PDK of the serine–threonine
kinase Akt. Akt phosphorylates and inactivates TSC1/2 and activates mTOR. mTOR stimulates ribosome biogenesis
and protein translation by activating initiation and elongation factors through phosphorylation of the ribosomal protein
E4517/Borer/fig.2.8/401473/TB/R3-kh
p70S6 kinase. It also prevents binding of 4E-BP-1 to eIF and inhibits GSK3β. TSC1/2 is activated by GSK3β and
AMPK and inhibited by Akt, MAPK (consisting of ERK1/2), and FAK (focal adhesion kinase).

activating genes involved in proteolysis and cell death. Under conditions of reduced muscle
activity and catabolic states, the dephosphorylation of FOXO TF increases atrogin 1 and
MuRF1 gene expression, and inhibits anabolic mTOR signaling and MRF gene expression
(Urso, 2009). Disuse proteolysis also may be activated by the TF of the NF-κB (nuclear
factor kappa light-chain-enhancer of activated B cells) family and by upregulation of myo-
statin (Jespersen et al., 2011).
The key mediator of muscle wasting is the ubiquitin proteasome system under the con-
trol of the FOXO family of TFs (see figure 2.6). The proteasome is a stack of four rings
encompassing a multisubunit enzyme complex in the cytoplasm that plays a central role
in the control of cell-cycle progression and apoptosis. The ubiquitin molecule and three
enzymes target specific proteins for degradation and generate a polyubiquitin chain. The
26S proteasome binds the polyubiquitin chain and denatures the attached protein as it is fed
into the proteolytic ring core of the proteasome that consists of trypsin-like, caspase-like,
or chymotrypsin-like enzymes. On the other hand, increased nutrition along with insulin-
stimulated activation of Akt, gene coactivator PGC-1α, and strength training counteract
protein degradation and stimulate protein synthesis by blocking the translocation of the
FOXO family of TFs to the nucleus and thus by reducing their abundance there.

Signal Transduction in Response to Bone Mechanical Loading


Bone cells also have the capacity to respond to mechanical stimulation in the absence of
anabolic or catabolic influence of circulating hormones. The putative mechanosensors
are the osteocytes constituting about 90% to 95% of cells in the human skeleton and are
embedded in the mineralized bone matrix. They form an interconnected three-dimensional
network through processes that have gap junctions, and they have the capacity to change
Activation of Nonhormonal Signaling During Exercise 41

the shape and connectivity of their processes (Bonewald & Johnson, 2008). Osteocytes are
immersed in interstitial fluid in a network of tiny channels, the canaliculi. Fluid shear in
the interstitial fluid in the canaliculi is caused by the pressure differential of the circulatory
system and by fluid flow caused by compression or distension of bone.
Intermittent but not continuous shear stimulates intracellular communication between
osteocytes through gap junctions and induces expression and release of PGE2. The signaling
initiated by this event regulates cytoskeletal reorganization or adaptive bone remodeling
(Bonewald & Johnson, 2008; Liu et al., 2012; Rubin et al., 2006, 2007). PGE2 is, in turn,
responsible for activation of the Wnt–β-catenin osteogenic pathway (figure 2.9; Bonewald &
Johnson, 2008). The first step in this cascade is upregulation of the gap junction α1 protein
(connexin) gene Cx43. This allows for an increase in intracellular Ca. PGE2 also acts in an
autocrine fashion to activate its EP12/4 receptor. The beta–gamma subunits of the associated
G protein activate PI3K, which in turn activates Akt. Akt phosphorylates GSK3β thereby
blocking its β-catenin inactivation by phosphorylation. β-catenin is a membrane protein
that is associated with scaffolding proteins and is involved in controlling cell adhesion and
anchoring actin cytoskeleton. The alpha subunit of the EP2 G protein prevents β-catenin

EP2/4 Lrp5
Fz
Int Wnt
eg Akt
rin
s
GSK3β
βcat
Cs43 HC
βcat
PGE2
ERα
Nucleus

βcat Wnt
Wnt

Figure 2.9 Activation of osteocyte signaling


E4517/Borer/fig.2.9/401474/TB/R2-alw
for adaptive bone remodeling. Shear stress induces PGE 2 expression
and release. PGE2 upregulates gap junction proteins (Cs43 HC) that act as hemichannels that admit Ca, activates
PI3K and Akt after binding to its receptor EP2/4, and facilitates β-catenin signaling through the trimeric components
of its receptor G protein. ERα facilitates β-catenin translocation to the nucleus. There, β-catenin upregulates the
Wnt gene, and the Wnt protein binds to Lrp5 and Fz to initiate bone remodeling.
Adapted from Bone, Vol. 42, L.F. Bonewald and M.L. Johnson, “Osteocytes, mechanosensing and Wnt signaling,” pgs. 605-615, copyright
2008, with permission from Elsevier.
42 Advanced Exercise Endocrinology

degradation and thus facilitates its signaling. Estrogen receptor ERα may play a role in
shuttling β-catenin into the nucleus in response to mechanical strain in osteoblasts. In the
nucleus, β-catenin upregulates the Wnt gene. Wnts are secreted glycoproteins that are post-
translationally modified by the addition of the palmitic fatty acid. The interaction of Wnt
with Lrp5 (lipoprotein receptor-related protein) and Fz (frizzled transmembrane spanning
protein) is involved in adaptive bone remodeling by Wnt. In this way, the β-catenin–Wnt
signaling pathway stimulates both osteoblast differentiation and bone formation in response
to shear stress and loading. PGE2 also facilitates the translocation of the CREBP to the
nucleus and transcription of prosurvival genes, including Bcl-2 (figure 2.10).
Bone-forming osteoblasts derived from mesenchymal stem cells (MSCs) in bone marrow
and bone-resorbing multinuclear osteoclasts derived from uninuclear hematopoietic cells
are functionally interrelated because they are responsible for bone turnover that occurs
spontaneously as well as in response to changes in tension, compression, torsion, and vibra-
tion. Studies on MSC, the osteoblast progenitor cells, grown on artificial scaffoldings and
subjected to forces that produce deformation and acceleration have demonstrated that these
cells can directly respond to such forces in the absence of circulating hormones.
Tension, compression, and interstitial shear cause deformation and displacement of cyto-
skeletal proteins (e.g., actin), membrane-spanning proteins (e.g., integrins and connexins) that
couple the cell to its environment, and cell-matrix adhesion proteins that anchor the cell in
place. Collagens I and II and fibronectin bind to integrin receptors. Exposure of osteoblasts
to compression results in rapid induction of bone morphogenetic protein (BMP), a stimulus
that is necessary for differentiation of MSC into osteoblasts (Rath et al., 2008). In addition,
compression induces expression in osteoblasts of runt-related TF 2 (Runx2) that is necessary
for expression of antiapoptotic genes such as Bcl-2 (figure 2.10). Finally, compression leads
to expression of genes and proteins that are required for extracellular matrix production,
such as alkaline phosphatase, collagen type
I, osteocalcin, osteonectin, and osteopontin Proteasomal PGE2
(Rath et al., 2008). degradation
Compared with ordinary physical activi- cAMP
ties, sport activities expose portions of the Bad
skeleton to greater strain magnitude, distribu- PKA
tion, and gradients. Therefore, an examination SMURF1 CREB 14
Bad P 3
of the effects of sports that exert high-impact, Runx2
3
odd-impact, and repetitive, low-impact load-
ings can reveal how these forces affect bone P CREB Survival Bcl-2 Bcl-2
adaptation. Bone cross-sectional geometry genes
responds to mechanical loading by periosteal
expansion. A small increase in bone diam- p21
eter and cortical thickness will increase the
resistance of bone diaphysis to bending loads.
An examination of the effects of long-term Antiapoptosis
involvement in sports revealed some com-
monalities in the adaptations based on the Figure 2.10 Shear stress induces PGE2, which
E4517/Borer/fig.2.10/401475/TB/R2-alw
mechanical properties of exercise stimuli such activates cAMP and PKA. This, in turn, facilitates
as high-magnitude vertical impacts; moder- translocation of CREBP into the nucleus and tran-
scription of the survival gene Bcl-2.
ate-magnitude impacts from varying, unusual
Adapted from Bone, Vol. 40, R.L. Jilka, “Molecular and cellular
directions; high-magnitude muscle forces; a mechanisms of the anabolic effect of intermittent PTH,” pgs.
great number of consecutive impacts of low to 1434-1446, copyright 2007, with permission from Elsevier.
Activation of Nonhormonal Signaling During Exercise 43

moderate magnitude; and a great number of consecutive contractions of non-weight-bearing


muscles on the cross-sectional geometry of the weight-bearing tibia (Nikander et al., 2010).
The following sports were associated with a thicker cortex at the distal tibia: high-impact
sports such as volleyball, hurdling, triple jump, and high jump; odd-impact sports such
as soccer and racket games (squash, tennis, and badminton); and repetitive, low-impact
exercise loading such as endurance running. At the tibial shaft, impact loading was associ-
ated with a thicker cortex as well as a larger cross-sectional area (CSA). High-magnitude
exercise loading such as powerlifting did not show such associations at either site but was
comparable with repetitive nonimpact loading and reference data for nonathletes. Thus, the
relevant determinants of these osteogenic adaptations in the weight-bearing tibia are high
strain rate and moderate to high strain magnitude.
Two additional signaling pathways are activated in response to mechanical loading of
the bone. Tension promotes differentiation of MSC to osteoblasts through activation of
MAPK signals and expression of collagen I and additional osteogenic genes (sialoprotein
2, osteocalcin, osterix; Ward et al., 2007). Stressed bone downregulates RANKL and
upregulates expression of eNOS in osteocytes through the MAPK signaling pathway (Rubin
et al., 2003). Reduced expression of RANKL by osteocytes diminishes osteoclastogenesis
because RANKL stimulates osteoclast differentiation by binding to the RANK receptor.
Mechanotransduction in bone has specific temporal requirements. Bone responds to
dynamic but not to static loading. Intermittent shear stimulates accumulation of PGE2
in osteoblasts, whereas continuous flow for 30 min stimulates ERK but has no effect on
expression of collagen I and osteopontin. However, when continuous flow is followed
by 13 d of no stimulation, gene expression of the last two matrix proteins as well as of
sialoprotein and osteocalcin increases (Kreke et al., 2008). This underscores the impor-
tance of alternating between exercise stress and restorative rest in promoting the anabolic
effects of exercise.

Effects of Vibration on Gene Expression


Vibration is the extreme mechanical oscillatory stimulus that can be made to approximate
frequencies of 20 to 50 Hz that are characteristic of type II motoneuron discharges during
muscle contraction. MSC from bone marrow have the capacity to differentiate into adipo-
cytes, osteocytes, and myoblasts. Fifteen minutes of vibration at 90 Hz during 15 weeks
suppressed MSC differentiation into adipocytes and increased differentiation into osteoblasts
(Rubin et al., 2007). At 12 wk, vibration produced a 28% reduction of fat to lean mass in
the torso of mice.
This vibration-induced shift toward osteogenesis at the expense of adipogenesis was
related to increased expression of Runx2 mRNA and decreased expression of PPARγ mRNA
in bone marrow cells (Luu et al., 2009). Osteoblastic differentiation from marrow pluripotent
stem cells is regulated by Runx2 (figure 2.10), followed by a factor known as osterix. Runx2
promotes expression of bone matrix protein genes such as type I collagen, osteocalcin, and
osteopontin. Adipogenic differentiation is driven by members of the C/EBP (the CCAAT/
enhancer-binding protein) family, including C/EBP α, β, and δ and PPARγ2. C/EBPβ and
C/EBPδ induce the expression of C/EBPα and PPARγ2. C/EBPα and PPARγ2 regulate
adipocyte-specific gene expression and play a role in the formation of mature lipid-filled
adipose cells from stem cells.
Mechanical loading appears to also increase osteogenic TF expression and decrease
adipogenic TF expression. For that reason, the relationship between vibration and mechani-
cal loading of exercise was examined in an animal model where young rats were trained
44 Advanced Exercise Endocrinology

over several weeks to jump from progressively greater elevations. Like whole-body
vibration, high-impact jump training induced an increase in bone volume and a decrease in
fat volume in the proximal metaphyses of tibiae. Like vibration, it did so by upregulating the
expression of BMP-4 and Runx2 in osteoblast lineage bone marrow stem cells. Although
no changes were found in adipogenic bone marrow factors, adipogenesis was most prob-
ably suppressed by loading-induced factors that promote osteogenesis, such as the cytokine
TGFβ, phosphorylation of PPARγ2, and upregulation of cells expressing Runx2. It appears
that the mechanism of vibration-induced osteogenesis and reduced adipogenesis strongly
resembles such changes elicited by the physical activity of jumping.

SIGNALING IN RESPONSE TO SENSING ENERGY NEED


The core process associated with physical activity is utilization of free energy in the form of
ATP. The inevitable consequence of exercise, therefore, is a momentary or more protracted
decrease in ATP availability. This explains the existence of AMPK, a widely distributed
sensor of changes in cellular energy depletion. AMPK is a heterotrimer consisting of a
catalytic alpha subunit and regulatory beta and gamma subunits (figure 2.11).
Each AMPK subunit has two or more isoforms. The α1 subunit is widely expressed,
including in the brain, whereas the α2 is the primary isoform in skeletal muscle, the heart,
and the liver (Winder & Thompson, 2007; Violett et al., 2007). The gamma subunit has two
Bateman domains for allosteric activation by cooperative binding of two AMP molecules
(Witzak et al., 2008); this activation is dose-dependent on exercise intensity (Jorgensen &
Rose, 2008). AMPK is inhibited by ATP and by high muscle glycogen concentration through
its beta subunit. Finally, hydrogen peroxide (H2O2), one of the reactive oxygen species (ROS)
that is a byproduct of oxidative phosphorylation (OXPHOS), also can increase the activity
of AMPK. H2O2 appears to activate AMPK through CaMKK signaling.
AMPK capacity, to become activated by intracellular increases in AMP:ATP ratio,
shuts down energy-consuming biosynthetic and anabolic pathways in service of energy
conservation, including those that mediate glycogen, fat, and ribosomal protein synthesis
(Long & Zierath, 2008). Instead, AMPK activates pathways responsible for glucose uptake,
increased insulin sensitivity, and increased lipid oxidation and mitochondrial biogenesis
that amplify the capacity for metabolic ATP production (Koh et al., 2008; Long & Zier-
ath, 2008). AMPK increases production of carnitine palmitoyltransferase I (CPT I), the
rate-limiting enzyme for mitochondrial long-chain fatty acid uptake, by phosphorylating
and inhibiting acetyl coenzyme A carboxylase (ACC), which produces CPT I inhibitor
malonyl-CoA. Endurance-exercise training also upregulates genes for two TF necessary
for mitochondrial gene transcription, Nrf-1 and PGC-1α, both of which require stimulation
by AMPK (Jorgensen et al., 2007; Long & Zierath, 2008).

Ca2+ LKB1
CaM TAK1
CaMKK

α
γ β
AMP

Figure 2.11 AMP binding to the γ-AMPK increases the affinity of LKB1 to phosphorylate the α subunit of the
enzyme. Phosphorylation of the α subunit also is facilitated by calmodulin and CaMKK as well as TAK1. Stimulation
of AMPK by TAK1 is initiated by TGFβ and some proinflammatory cytokines.
Based on Jorgensen and Rose 2008.

E4517/Borer/fig.2.11/401476/TB/R3-kh
Activation of Nonhormonal Signaling During Exercise 45

Covalent AMPK activation can be more potent than allosteric AMPK activation by AMP.
Covalent activation occurs through phosphorylation of threonine T172 on the catalytic alpha
subunit by any of several kinases: LKB1 (a serine–threonine kinase and a tumor suppressor
gene product), TAK1 (a ubiquitin-dependent kinase of MKK and IKK), and CaMKK; the
last two are members of the MAP-kinase-kinase family (see figure 2.11). LKB1 is more
highly expressed in slow oxidative skeletal muscles than in fast-twitch skeletal muscles
and is constitutively active during metabolic stress. Endurance-exercise training appears to
induce mitochondrial biogenesis and aerobic adaptations in skeletal muscle in part through
an increase in LKB1. LKB1 is involved in AMPK activation in resting muscle in response
to AICAR (5-aminoimidazole-4-carboxamide-1-rosiglitazone) and intensive electrical
stimulation. AICAR, along with the antidiabetic drugs metformin (N,N-dimethylimidodi-
carbonimidic diamide) and rosiglitazone, represent a pharmacological means of AMPK
activation. These TZD (thiazolidinedione) drugs are PPARγ agonists that appear to facilitate
muscle glucose uptake through activation of AMPK. Activation by AICAR of AMPK is
restricted to fast glycolytic muscles. It depends on the γ3 regulatory subunit of the enzyme
and is abolished by muscle fiber conversion to oxidative Type I (Long & Zierath, 2008).
AMPK exerts these metabolic effects largely by regulating the transcription of metabolic
genes. AMPK-induced increase in Glut4 and Hk2 gene expression and GLUT-4 translocation
is one of several insulin-independent mechanisms that increase glucose transport during
muscle contraction.
Fed muscle accounts for up to 75% of glucose disposal. An insulin-sensitive muscle has
the capacity to switch from carbohydrate to lipid utilization during exercise, depending
on exercise intensity and fuel availability. This metabolic flexibility is reduced in obese
individuals due to the resistance of peripheral tissue to insulin action. It is therefore of great
interest that muscle contraction facilitates insulin-independent glucose uptake and can
increase glucose uptake in insulin-resistant individuals who exercise (Richter et al., 1985;
see also “The Underappreciated Power of Exercise and Training to Prevent Diabetes”). As
previously stated, activation of AMPK and Cn during exercise increases muscle glucose
uptake, insulin sensitivity, and glycogen storage capacity (Long & Zierath, 2008). The current
understanding of contraction-induced increase in glucose uptake is that a number of Ca2+-
sensitive proteins (including CaM, Cn, CaMKII, CaMKK, AMPK, and PKC) participate
in translocation of GLUT-4 to the muscle cell membrane (Cartee & Wojtaszewski, 2007;
Jessen & Goodyear, 2005; Röckl et al., 2008; figure 2.12). Protein AS160, the Akt substrate

Insulin Glucose
IR uptake Muscle contraction

+
Ca2 AMP

CaM
P13K GLUT-4

CaMKII LKB1

CaMKK
PKB/Akt AS160 AMPK

Figure 2.12 Insulin-independent, contraction-stimulated glucose uptake is mediated by multiple signaling


E4517/Borer/fig.2.12/401477/TB/R3-kh
pathways, including aPKC, Cn, CaMKII, CaMKK, LKB1, and AMPK. Insulin-dependent signaling is shown on the left.
Adapted, by permission, from K.S.C. Röckl, C.A. Witczak, and L.J. Goodyear, 2008, “Signaling mechanisms in skeletal muscle: Acute
responses and chronic adaptations to exercise,” IUBMB Life 60: 145-153.
46 Advanced Exercise Endocrinology

of 160 kDa, and the protein TBC1D1 regulate GLUT-4 translocation to the cell membrane
(figure 5.2). Both are activated by PKB or Akt, and TBC1D1 also has an AMPK binding
site (Cartee & Funai, 2009; Cartee & Wojtaszewski, 2007; Röckl et al., 2008).

FREE RADICALS AS INITIATORS OF MESSAGE


TRANSDUCTION
Strenuous exercise results in oxidative stress to the muscle and other organs through genera-
tion of ROS during oxidations in the mitochondria. ROS include such oxygen derivatives as
singlet oxygen, hydroxyl radical, and hydrogen peroxide, all of which have a highly reactive
unpaired electron in their outer valence shell. A moderate pro-oxidant state has the capacity
to damage DNA and lipid membranes and produce protein aggregation. It has also been
shown to increase skeletal muscle contractile force through increases in intracellular Ca2+
concentration and sensitivity to Ca2+ by myofilaments (Powers & Jackson, 2008). In addi-
tion, discovery of genes and metabolic pathways affected by ROS has revealed their role
as subcellular messengers in gene regulatory and signal transduction pathways (Allen &
Tresini, 2000; Niess & Simon, 2007).
Cellular components that are sensitive to redox changes are NF-κB, AP-1 (activator pro-
tein-1), MAPK, PKC, HFS1 (heat shock transcriptional factor-1), AMPK, insulin receptor
(IR) kinase, and MSK-1 (stress-activated protein kinase). MAPK and NF-κB account for
about half of ROS effects on cell function.

MAPK
The mitogen-activated protein kinases (MAPKs) comprise a family of enzymes that modulate a
variety of processes associated with cell growth or apoptosis. These processes include transcrip-
tion and translation of proteins, cell proliferation and differentiation, and muscle hypertrophy.
The MAPK family consists of four signaling cascades: stress-activated protein kinase
p38, JNK1/2 (c-jun terminal kinases 1 and 2), ERK (extracellular signal-regulated kinases)
1 and 2, and (MAPK-5). Their extracellular stimulators are growth factors and proinflam-
matory cytokines (TNFα, IL-1, IL-6). In addition, both exercise and ROS activate MAPKs
(Hawley & Zierath, 2004). Acute concentric exercise activates ERK1/2, whereas very intense
or eccentric exercise activates both ERK1/2 and p38. MAPK signaling cascades follow a
common pattern (figure 2.13). A multiprotein complex, Shc–Grb-2–SOS (see figure 5.1),
converts an inactive G protein, Ras-GDP, to active Ras-GTP. ROS directly activate both
the Shc–Grb-2–SOS complex and Ras (Allen & Tresini, 2000).
In addition, Ras can generate additional ROS by stimulating another small G protein,
Rac. Rac activates a membrane-bound NADPH oxidase, which generates more ROS. Ras
also regulates the activity of MAP3K, and they control serial activations of MAP2K and
MAPK. Members of each of the three MAPK cascades phosphorylate nuclear proteins and
the early-response genes c-jun and c-fos that initiate gene transcription. c-jun is activated by
JNK and c-fos is activated by ERK1/2. The two form a heterodimeric TF, AP-1, that also
regulates transcription of a number of genes in a redox-sensitive manner.

Nuclear Factor κB
NF-κB is a TF that mediates immune and inflammatory responses as well as cellular growth
and apoptosis. It exists in the cytoplasm as an inactive heterodimer consisting of proteins
p50 and p65 and an inhibitory (IκB) unit (figure 2.14). The ROS hydrogen peroxide (H2O2)
activates NF-κB by initiating serial phosphorylations of NIK (NF-κB-activating kinase) and
Activation of Nonhormonal Signaling During Exercise 47

Growth factors
Stress

Proinflammatory
Ras cytokines
Rac1
PAKK Rho1
Cell membrane

Raf MEKK1 MEKK

MEKK1/2 MKK4 MKK3/6

ERK1/2 JNK1/2 P38MAPK

Nuclear membrane
EIk

SRF

myc

Max
ATF-2

Ets/SRE CRE/TRE E-Box

Figure 2.13 MAPK signaling stress pathways consist of ERK1/2, JNK1/2, and p38 MAPK. They can be activated
by growth factors, inflammatory cytokines, and ROS at the level of the Shc–Grb-2–SOS complex and Ras G protein.
E4517/Borer/fig.2.13/401479/TB/R3-kh
Adapted from Free Radical Biology and Medicine, Vol. 28, R.G. Allen and M. Tresini, “Oxidative stress and gene regulation,” pgs. 463-499,
copyright 2000, with permission from Elsevier.

IκK (IκB kinase) that mark the IκB for proteolytic degradation in the proteasome. The p50/
p65 dimer then translocates to the nucleus, where it activates transcription of the genes for
the antioxidant enzyme MnSOD, cell adhesion molecules, iNOS (inducible NO synthetase),
acute phase proteins, cytokines, and hematopoietic growth factors.
A pro-oxidant state also upregulates Hsp gene transcription and translation. HSPs pro-
vide antioxidant protection during, for example, cardiac reperfusion after ischemia. HO1
(heme oxidase 1), an antioxidant stress protein, is strongly induced by ROS and hypoxia
that also can be generated by high-intensity exercise. HO1 gene expression is regulated by
NF-κB, AP-1, MAPK, and HIF-1 (hypoxia inducible factor-1). Thus, ROS, the byproducts
of OXPHOS, have the capacity to affect muscle gene expression during acute exercise and
contribute to adaptations to exercise training.

Effect of ROS on Apoptosis


ROS also can initiate apoptosis or cell death. This is often a consequence of ROS-incurred
DNA damage that is signaled through the release into cytoplasm of p53 protein as well
as damage to the lipids in the mitochondrial membrane (Jackson et al., 2002). Increase
in expression of p53 arrests cell growth cycling in the G1 interphase stage to allow for
DNA repair or may precipitate transition to the G0 stage and apoptosis (figure 2.15). Cells
transition between the four phases of the cell-division cycle: the G1 interphase phase of
biosynthetic increase in cell size; the S, or synthesis, phase when DNA replication occurs;
the G2 interphase; and the M, or mitosis, phase. In the G0 phase, cells are reversibly or
48 Advanced Exercise Endocrinology

NIK IκK Ub
Ub
Ub
IκB P Ub
P P Ub
P Proteasome
NLS NLS
Dimerization
domain

DNA binding
domain
P50 P65 P50 P65
NF-κB

Figure 2.14 Activation of NF-κB. E4517/Borer/fig.2.14/401480/TB/R3-alw


ROS and some other stimuli can activate the proinflammatory TF NF-κB
through phosphorylation of two kinases, NIK and IκK. The p50/p65 heterodimer translocates to the nucleus to
activate transcription of antioxidant enzyme genes.
Adapted from Free Radical Biology and Medicine, Vol. 28, R.G. Allen and M. Tresini, “Oxidative stress and gene regulation,” pgs. 463-499,
copyright 2000, with permission from Elsevier.

permanently quiescent or may proceed to apoptosis. Transition between the phases is under
the control of the regulatory proteins cyclin and cyclin-dependent kinases (CDKs). CDKs
are constitutively expressed, but the cell cycle advances only after phase-specific cyclins,
activated by an external stimulus, bind to CDK. The cyclin–CDK complex can also target
some proteins for ubiquitination in the proteasomes.
Whether a cell enters apoptosis depends on the expression (or lack) of the survival genes
Bcl-2 and c-myc and on the mitochondrial response to p53 (figure 2.16). Damage to the
mitochondrial membrane leads to the release of cytochrome c into cytoplasm and activation
of proapoptotic caspase enzymes.
The goal of this chapter is to underscore the capacity of muscle, bone, and other tissues
to respond to external stimuli as well as mechanical and bioenergetic challenges of exer-
cise independently of circulating endocrine messengers. This does not negate cooperation
between tissue-specific mechanisms and hormones, as is the case of AMPK interaction
with leptin and adiponectin in the facilitation of fatty acid oxidation (Winder et al., 2000).
Hormones such as adiponectin, leptin, IL-6, CNTF (ciliary neurotrophic factor), and cat-
echolamines can activate AMPK in resting muscles, but AMPK activation during exercise
does not depend on circulating factors (Jorgenson & Rose, 2008).
G2

S M

Hormonal triggers of
Hormonal triggers of hypertrophy
proliferation:
G1 and differentiation:
IGF-1, bFGF, TNF , TGF
IGF-1, calcineurin, TGF, T3, GH,
G0 glucocorticoids, MRFs:
myogenin, MyoD, MRF4, and Myf5

Apoptosis

Figure 2.15 The cell growth cycle is influenced by ROS but is usually under hormonal control. DNA replication
occurs in the S phase, preparation for mitosis in the G2 phase, and mitosis in the M phase. Terminal differentiation
and protein synthesis occur in the G1 phase, and transition to the G0 phase leads to apoptosis.
Adapted, by permission, from E.A. Newsholme and A.R. Leach, 1983, Biochemistry for the medical sciences (New York: John Wiley and
E4517/Borer/Fig2.15/401481/alw/pulled-R2
Sons), 685. © 1983 John Wiley & Sons, Limited. Reproduced with permission.

Free radical
mediated changes
Receptor mediated events Oxidative
to mitochondrial
permeability damage
TNF, FAS, APO1 death receptor CD95

FADD Mitochondrion
GSH ROS
Nucleus
Bax/Bak P53
DNA damage
Procaspase-8 Bcl2
Bcl-x Nuclear targets
•Lamins
Cytochrome c •PARP
•DNAase (DFF)
Cytochrome c
Caspase-8
APAF-1
Procaspase-9

Caspase-9

Effector caspases Cytoplasmic targets


Caspases –3 & –7 •Cytoskeletal proteins
•Protein kinase C,δ,Rb,etc.

Figure 2.16 Induction by ROS of cellular apoptosis. ROS damage DNA, leading to the release of p53 into
cytoplasm. Damage by ROS to the mitochondrial membrane increases its permeability and leads to the release of
cytochrome c into cytoplasm and activation of proapoptotic caspase enzymes.
Adapted from Molecular Aspects of Medicine, Vol. 23, M.J. Jackson et al., “Antioxidants, reactive oxygen and nitrogen species, gene
induction and mitochondrial function,” pgs. 209-285, copyright 2002, with permission of Elsevier.

E4517/Borer/fig.2.16/401482/TB/R2-alw

49
50
(continued)

51
(continued)

52
Activation of Nonhormonal Signaling During Exercise 53

SUMMARY
Most of the endocrine and autonomic consequences of exercise discussed in this text repre-
sent endocrine and autonomic actions resulting from messenger–receptor binding. However,
exercise-associated signal transduction can be activated in the absence of the systemic influ-
ence of circulating hormones or of the binding of the hormones with the receptors. Muscle
contractions are accompanied by electrical, ionic, mechanical, metabolic, and oxidative
effects, each of which has the capacity to trigger an appropriate physiological response or
an adaptive change. These effects are often mediated by the same signaling pathways that
operate after hormone–receptor binding except that they are activated by specific exercise
effects at points distal to hormone–receptor binding.
Motoneuronal discharges and accompanying electromagnetic events can elicit patterns of
gene expression for the synthesis of contractile proteins suitable for support of a particular
pattern of muscle contractions. Intracellular Ca release that accompanies muscle contraction
can trigger both metabolic responses such as glycogenolysis or genetic programs for fast or
slow muscle fibers by binding to proteins with a high affinity for Ca. Ca BPs also stimulate
NO production in the muscle. Mechanical strain and microvibrations at frequencies charac-
teristic of motoneuron discharges affect bone turnover and stimulate muscle hypertrophy.
Muscle is one of several tissues that can sense and respond to energy need. Several enzymes,
most notably the AMPK, respond to low muscle-energy charge by blocking energy-costly
biosynthetic processes and growth and by mobilizing stored metabolic fuels. In addition, the
ROS that are produced by oxidative stress during exercise affect skeletal muscle contractile
force and affect transcription of genes for enzymes that can mount antioxidant defenses.
Physical activity is too important for the survival of the organism to rely exclusively on
hormonal mechanisms. For that reason, nonhormonal mechanisms evolved and add adaptive
redundancy in physiological responses to exercise.
This page intentionally left blank.
H APTER

3
C

Autonomic and
Hormonal Control of
the Cardiorespiratory
System
T
he central function of the cardiorespiratory system is to provide all cells in the human
body with the oxygen needed to generate energy from nutrients and to remove carbon
dioxide, the chief byproduct of metabolism. This system is exquisitely responsive to
metabolic demands as physical activity increases from rest to intense exercise. The ANS
and several hormones it activates are responsible for the control of rapid and automatic
adjustments in cardiorespiratory function, both in response to afferent error signals from
peripheral cardiorespiratory organs and from feed-forward activation by CNS command
neurons. Two divisions of the ANS that have generally opposing functions provide dual
efferent innervation of most organs important to physiological adaptation during exercise.
The S division of the ANS stimulates the cardiorespiratory function during rest and increases
it in response to the energy demands of exercise. The PS division reduces the activation of
cardiorespiratory function during rest. It also stimulates ingestion, digestion, and absorption
of food. This chapter highlights the roles of the ANS and auxiliary hormones in maintain-
ing homeostasis at rest and extraordinarily increasing oxygen supply for metabolism during
exercise. The important features outlined are the role of sensory input from peripheral
organs during exercise in elicitation of autonomic control of cardiorespiratory function
during exercise; the functional significance of dual autonomic innervation of the heart,
vasculature, and respiratory organs in elicitation of these reflexes; and coordinated control
of cardiorespiratory function by the ANS, auxiliary hormones, and central command.
This chapter also introduces the autonomic control of other functions that are impacted by
exercise and discussed in later chapters.

55
56 Advanced Exercise Endocrinology

AFFERENT SIGNALS TO THE ANS


ANS pathways are organized at two levels of complexity: those that serve for reflex homeo-
static adjustments of fluctuations in the internal environment and those that make reciprocal
connections with the CNS centers to modulate autonomic reflexes that support behaviors
such as exercise, feeding, drinking, and Na hunger. Integral components of all reflexes are
sensory afferent input, integrative CNS circuits, and efferent (in this case, autonomic) motor
nerve or endocrine components.
We do not consciously sense most signals arising from receptors in the internal envi-
ronment, the extracellular fluids of the body as defined by Claude Bernard. Chemical and
mechanical sensory afferents from the vasculature, heart, and lungs travel in PS glossopha-
ryngeal (9th) and vagus (10th) cranial nerves to the nucleus of the tractus solitarius (NTS)
(figure 3.1) in the dorsomedial MO but do not proceed to higher brain centers to produce
conscious detection of such changes. These signals achieve regulation of blood pressure,
partial pressure of oxygen and carbon dioxide, plasma ionic concentrations and acid-base
balance, and blood sugar and other metabolite concentrations through elicitation of hormonal
and autonomic reflexes. The afferents participating in these autonomic reflexes originate
from high-pressure arterial baroreceptors in the aortic arch and carotid sinus (figure 3.2) that
sense changes in arterial pressure; low-pressure cardiac baroreceptors in atrial walls and in
the great veins at their confluence with the atria that sense changes in venous return to the
heart; peripheral chemoreceptors in carotid bodies and the aortic arch that detect changes
in arterial pO2, pCO2, and pH; and slowly adapting and rapidly adapting pulmonary stretch
receptors. Additional sensory input to the NTS comes from the central chemoreceptors
on the ventral surface of the MO that monitor changes in pCO2, joint mechanoreceptors
that signal limb movement through undifferentiated nerve endings, muscle receptors that
transmit the information about the intracellular metabolic changes via group III myelin-
ated mechanoreceptor afferents and group IV unmyelinated chemoreceptor afferents, gut
afferents travelling in the sensory vagus, and taste afferents in afferent fibers in facial and
glossopharyngeal cranial nerves. From the NTS, integrative neurons make connections
with either the PS or S efferent neurons. The PS efferents involved in the reflex control of
the heart and respiration originate as preganglionic fibers in the nucleus ambiguus (NA)
in the medulla oblongata (MO), and others reflexively controlling digestion and absorp-
tion of food, in the dorsal motor nucleus of the vagus (DMN) (figure 3.1). S preganglionic
efferents originate in the intermediolateral (IML) column of the spinal gray (see figure 1.2)
and control autonomic reflexes through their trajectory through thoracic and lumbar spinal
nerves (see figure 1.1), and synapses with postganglionic neurons in the paravertebral or
preaortic ganglia before reaching their peripheral organ targets.
Some stimuli bridge the external and internal environments and allow conscious recogni-
tion of food, its ingestion, and its gastrointestinal digestive transit. These events are tracked
by several sensory systems that project to higher brain centers responsible for the control
of the motivation to seek and ingest food (discussed in chapter 6). Pain in the heart (angina
pectoris) or in portions of the gut is transmitted to the MO by afferent fibers in S nerves
comingled with a high proportion of somatic sensory fibers so that the conscious spatial
localization of pain in these visceral organs is referred to a large and diffuse body surface
area (Cervero & Foreman, 1990). Main sensory systems monitor changes in the external
environment including those representing sport competition tension or representing threats
that could cause tissue damage and body harm. Although the last two categories of stimuli
entail conscious processing of sensory information that informs deliberate behaviors, all
Nucleus
tractus solitarius Vagus nerve

Dorsal vagal
nucleus

XII

Stomach fundus (gastric acid)


Stomach antrum
Nucleus
Pancreas (insulin, glucagon)
ambiguus

Heart

Striated muscles of
pharynx, esophagus,
and larynx

Figure 3.1 Preganglionic vagal neurons that innervate the cardiorespiratory organs originate in the nucleus
ambiguus (NA). GI organs are innervated by vagal neurons from the dorsal motor nucleus of the vagus (DMV).
Adapted from CENTRAL REGULATION OF AUTONOMIC FUNCTIONS by A.D. Loewy & K.M. Spyer, copyright 1990 by Oxford University Press,
Inc. Used by permission of Oxford University Press, Inc.

E4517/Borer/fig.3.1/401525/TB/R2-alw

PG NG Carotid
NTS sinus Carotid baroreceptors
Common carotid artery
Sensory
fibers Aorta
NA Aortic baroreceptors

RVLM
Parasympathetic
vagus nerve
S-A node
A-V node

Spinal cord
(transverse
sections)
Sympathetic nerve
Sympathetic trunk

Figure 3.2 S and PS innervation of the heart. The preganglionic S neurons are shown as originating from spinal
E4517/Borer/fig.3.2/401527/TB/R2-alw
cord segments T1 to T4, synapsing in the cervical and thoracic ganglia of the paravertebral S chain, and terminating
in the atria and ventricles of the heart. The vagus nerve originates in the NA of the MO and innervates the atria and,
to a lesser extent, the ventricles. Baroreceptor afferents are shown from the carotid artery sinus and the aortic arch.
Adapted, by permission, from S.K. Powers and E.T. Howley, 2012, Exercise physiology: Theory and application to fitness and performance,
8th ed. (New York: McGraw Hill), 201. The McGraw-Hill Companies; K.T. Borer, 2003, Exercise endocrinology (Champaign, IL: Human
Kinetics), 31.

57
58 Advanced Exercise Endocrinology

of them are integrated with hormonal and autonomic reflexes. Thus ingestive and digestive
events trigger hormonal and autonomic reflexes that facilitate food processing, while the
perception and experience of danger trigger autonomic correlates of fight or flight physi-
ological responses as conceptualized by Walter Cannon.

INTEGRATIVE BRAIN CIRCUITS AND EFFERENT


COMPONENTS OF ANS REFLEXES
The NTS (figure 3.1) is the chief
brain center that collects sensory Cerebral cortex
information from these multiple
visceral modalities in a spatially
organized fashion (Loewy & A
Spyer, 1990). Because there is an
overlap between sensory inputs
from baroreceptors, chemorecep- PVN
tors, and lung receptors in the
commissural part of the NTS,
B
some reflex integration likely Kölliker-Fuse
occurs in this nucleus. Additional Raphe magnus
integration takes place in higher
A5 cell group
brain centers that have reciprocal
connections with the NTS. These C
A1 cell group
brain structures, in turn, have
projections to the S preganglionic Raphe obscurus
RVLM
neurons in the IML column of
thoracic and lumbar spinal gray Raphe pallidus
matter (figure 3.3) or to pregan- IML cell column
glionic PS neurons in the DMV
with projections to gut organs and
the NA, which has projections Heart
to the heart, vasculature, and Blood
respiratory organs (figure 3.1). An vessels
explanation of the dual autonomic
control of the heart, vasculature, Figure 3.3E4517/Borer/Fig3.3/401657/alw/pulled-R3-kh
Nuclei in the hypothalamus and the brain stem
and respiratory organs is helpful that activate spinal S neurons to the heart and blood vessels.
in understanding the integrated Adapted from CENTRAL REGULATION OF AUTONOMIC FUNCTIONS, edited
by A.D. Loewy & K.M. Spyer, copyright © 1990 by Oxford University Press,
function of the cardiorespiratory Inc. Used by permission of Oxford University Press, Inc.
organs during exercise.

Heart
Neural and hormonal control of the heart involves dual and reciprocal actions of the S and
PS divisions of the ANS and influences from several hormones.

S Control of Heart Function


Preganglionic S nerves to the heart originate in the IML region of the thoracic spinal seg-
ments T1 to T4 and make contact with postganglionic nerve fibers in three cervical ganglia
and four thoracic ganglia in the paravertebral chain. Postganglionic neurons in the cardiac, or
accelerator, nerves innervate the sinoatrial (SA) and the atrioventricular (AV) nodes, and the
Autonomic and Hormonal Control of the Cardiorespiratory System 59

sarcolemma of the cardiac muscle (figure 3.2).


The preganglionic neurons of both S and PS
divisions of the ANS use Ach as a neurotrans-
mitter that acts on nicotinic receptors on the
postganglionic neurons. Nicotinic receptors
(see chapter 1) are ligand-activated ion chan-
nels that open to allow flow of cations (Na, K,
Ca) in and out of the cell. The postganglionic
S nerve terminals are elongated structures that
contain neurotransmitter-filled vesicles (figure
3.4) that discharge NE into synaptic clefts on
the heart cell surface. NE activates the heart
pacemaker and muscle by acting on β1 AdR.
S nerve discharge is the principal S activa-
tor of cardiac and circulatory systems during
exercise and other stress. Some NE escapes
from the S synaptic clefts into circulation,
where it acts as a hormone. This NE spillover
from the heart and other target organs, rather
than adrenal NE, is the main contributor to the
circulating plasma NE concentration, which at
rest amounts to about 0.2 ng/ml. Circulating E
also contributes to the activation of the heart.
It originates from the adrenal medulla, an S
ganglion in which the postganglionic neurons
have developmentally been converted into Figure 3.4 S nerve ending in the iris of a cat. The
hormone-secreting endocrine cells. Approxi- muscarinic PS terminal (C) is shown to synapse on
mately 80% of adrenal medullary output is E the elongated S terminal (A).
and only 20% is NE. Although it activates β1 Springer and Experientia Vol. 24, 1968, pgs. 484-486, “Various
AdRs of the heart, E action occurs principally types of amino-storing vesicles in peripheral adrenergic nerve
terminal,” J.P. Tranzer and H. Thoenen, figure 4, with kind
through activation of β2 and β3 AdRs (see permission from Springer+Business Media B.V.
chapter 1 for more details).
NE and E are equally effective in stimulating α and β1 receptors. Therefore, they activate
with equal efficiency the heart muscle, which is predominantly populated with β1 recep-
tors and arterioles in which α1 receptors are highly responsive to catecholamines. On the
other hand, β2, and β3 receptors respond poorly or not at all to NE. This accounts for the
predominant role of circulating E in broncho- and vasodilatation, mobilization of metabolic
fuels, and stimulation of oxidative metabolism. E and NE stimulate heart contraction by
increasing both the chronotropic (frequency) and inotropic (force) functions. The chrono-
tropic activation by catecholamines is achieved through shortening of the latency of SA
pacemaker depolarization through earlier opening of Na, Ca, and K channels (figure 3.5a;
Vander et al., 1998). The increased inotropic activation by catecholamines (figure 3.5b) is
achieved through increased NE-induced Ca release in the heart muscle. Prolonged open-
ing of the Ca2+ channels ensures that a heart muscle contraction is sustained for about one
third of a second, or 300 times longer than a twitch of a skeletal muscle fiber. This feature
of heart muscle depolarization ensures sufficient time for filling the heart chambers with
blood. The inotropic effect is additive to the Frank-Starling mechanism through which the
increases in end-diastolic volume produce proportional increases in the strength of the heart
contraction (figure 3.5c). S nerves initiate heart contraction (systole), which takes up about
one third of the resting heart cycle.
60 Advanced Exercise Endocrinology

60
a,b,c are pacemaker potentials:

a = Control During stimulation


b = During sympathetic stimulation of sympathetic
Membrane potential (mV)

c = During parasympathetic stimulation nerves to heart

during contraction
Force developed
0

Control
Threshold
potential
−40
b a
c
−60

a Time b Time

E4517/Borer/Fig3.5a/448541/alw/r1 E4517/Borer/Fig3.5b/401629/alw/r2
Sympathetic
stimulation
200
Stroke volume (ml)

100
Control

Normal
resting
value
0 100 200 300
c Ventricular end-diastolic volume (ml)

Figure 3.5 (a) Shortening of the latency of heart pacemaker depolarization by S nerves, and lengthening by PS nerves.
E4517/Borer/Fig3.5c/401630/alw/pulled-R1
(b) Chronotropic and inotropic effects of S stimulation of the heart. (c) Additive increases in stroke volume contributed by
S stimulation and the stretch of the ventricular wall by increases in end-diastolic volume.
Reprinted from A. Vander, J. Sherman, and D. Luciano, 1990, Human physiology: The mechanisms of body function, 5th ed. (Boston: McGraw-Hill),
376, 378. © The McGraw-Hill Companies.

Activation of S nerves during exercise occurs early and accelerates with exercise inten-
sity. Adrenal secretion of E occurs later in exercise and requires greater relative effort
(figure 3.6). Cardiac responses to S activation can be modified by certain health conditions
because AdR populations are dynamic with respect to receptor type, number, and tissue
distribution. Prolonged exposure to catecholamines at high concentrations results in recep-
tor desensitization. This involves reduction in receptor number due to internalization and,
sometimes, uncoupling of receptors from components of signaling pathways. Denervation
of target organs by means of pharmacological catecholamine depletion or transection of S
nerves produces supersensitivity to catecholamines in both major types of AdR, usually due
to receptor upregulation. Such increased responsiveness to catecholamines is, for example,
seen in transplanted hearts. Supersensitivity also occurs when patients discontinue the use
Autonomic and Hormonal Control of the Cardiorespiratory System 61

E NE
1.0 6
E, NE (ng/ml)

0.8 4
NE
E
0.4 2

0 0
40 50 60 70 80 90 100
Maximal effort (%)
.
Figure 3.6 Changes in .catecholamine concentrations in response to 20 min of exercise at 50% and 70% of VO2max and 10
min of exercise at 90% of VO2max.
Adapted from Lugar et al. 1992; Deuster et al. 1989.
E4517/Borer/Fig3.6/401528/alw/pulled-R1

of beta blockers. Daily postural changes are also associated with changes in β AdR density.
Upright posture increases plasma concentration of catecholamines and reduces the number
of β AdR. Hyperthyroidism results in increases in β1 AdR on the heart muscle, whereas
hypothyroidism does the reverse. Hyperthyroidism is associated with increased heart
contractility and rate, skeletal muscle tremors, and hyperthermia. Elevation of circulating
corticosteroids also leads to increases in β1 AdRs in the heart muscle. In hypertension as
well as in aging, heart responsiveness to S stimulation is attenuated, most likely due to
myocardial β1 AdR downregulation or G protein uncoupling. Hypoxia is associated with
reduced β1 AdR density and increased plasma NE concentrations.

PS Control of Heart Function


Preganglionic neurons of the PS nerve vagus that synapse on the short postganglionic neu-
rons of the heart, trachea, bronchi, and glands of the respiratory tract originate in the NA
of the MO, also known as the ventral vagal motor complex (see figure 3.1). Ach released
from postganglionic vagal neurons acts on cholinergic MR on target organs (see chapter
1 for more details). Vagal control of HR increases the randomness of intervals between
successive heart beats. HR variability is relatively high during rest and is reduced during S
activation associated with assuming upright posture and physical activity (figure 3.7). HR
variability imparted through vagal participation in HR control reduces the risk of coronary
heart morbidity and mortality.
The postganglionic vagus predominantly innervates the SA node and, to a lesser extent,
the AV node, atrial muscle, Purkinje fibers, and ventricular muscle of the heart (Wisler
et al., 1992; see figure 3.2). The vagus also innervates the respiratory organs but does not
innervate the vascular tree. Short postganglionic neurons of the vagus also form synapses
on S nerve endings, where they can control NE release (see figure 3.4). The major action
of the vagus is to reduce HR and the force of heart contraction. Reduction of chronotropic
function is accomplished through hyperpolarizing the SA node cells and delaying their
discharge (see figure 3.5a). Reduction of inotropic function results from lower intracellular
Ca concentration as a consequence of vagal antagonism of S nerve action. PS nerves initiate
heart muscle relaxation (diastole), which takes up about two thirds of the resting heart cycle.
62 Advanced Exercise Endocrinology

Supine Upright

1.5 1.5
RR mean = 920 ms Variance = 1,412 ms2 RR mean = 724 ms Variance = 299 ms2
RR (sec)

.5 .5
1 Beats 200 1 Beats 200

7,000 VLF
2,000
VLF = 0.0 Hz P = 524 ms2 VLF VLF = 0.0 Hz P = 136 ms2
LF = 0.9 Hz P = 529 ms2 P = 59 nu LF = 0.1 Hz P = 128 ms2 P = 81 nu
PSD (ms2/Hz)

LF HF = 0.33 Hz P = 261 ms2 P = 29 nu HF = 0.36 Hz P = 18 ms2 P = 11 nu


LF/HF = 2 LF LF/HF = 7.1
HF

HF
0.0 0.0
0.0 .1 .2 .3 .4 .5 0.0 .1 .2 .3 .4 .5
Hz Hz

Figure 3.7 HR variability during supine (left) and upright (right) posture presented in temporal (top) and spectral
(bottom) domains. Lower HR during supine posture is associated with greater HRV and increased vagal tone, shown
as the high-frequency (HF) peak of spectral HR frequencies. Higher HR after standing reflects S activation, reduces
E4517/Borer/Fig3.7/401531/alw/R2
HRV, and presents as a spectral low-frequency (LF) peak.
Reprinted, by permission, from A. Mallian, 1999, “The pattern of sympathovagal balance explored in the frequency domain,” News in
Physiological Sciences 14: 111-117.

S and PS nerves and circulating E control ultradian (<24 h) and circadian (diurnal)
rhythmicity of heart contractions. Heart pacemakers possess an inherent ultradian rhythm
of spontaneous depolarization caused by a slow inward leak of Na and an outward leak of
K. The dominant SA node spontaneously depolarizes approximately 100 times per min,
and the subordinate AV pacemaker depolarizes about 60 times per min. Resting HR in most
individuals is considerably slower than the rate of pacemaker depolarization due to tonic
inhibition of the SA node by the vagus. Resting HR and respiratory rhythm result from
alternating rhythmic activation of S and PS discharges. This control is exerted by several
adjacent nuclear groups in the rostral ventrolateral medulla (RVLM) and the pons (see figures
3.3 and 3.16). Circadian rhythm of HR and blood pressure consists of partial withdrawal
of PS and an increase of S tone upon morning arousal and assumption of upright posture,
consistent with requirements for cardiorespiratory support of increased energy demands
during wakefulness compared with during sleep (figure 3.8).

Vasculature
Exercise elicits two contrasting changes in the vascular bed. In one change, the catechol-
amines and several endothelial and systemic hormones cause vasoconstriction in most
vascular beds, other than those in the exercising muscle. This contributes to redistribution
of blood from visceral organs to contracting muscles (figure 3.9). In the other change,
increased conductance in muscle vasculature, called exercise hyperaemia, is mediated by
the release of endothelial vasodilators.
160

150

140

130
SBP (mmHg)

120

110

100
Figure 3.8 The circadian rhythm of systolic
blood pressure in a healthy 62-yr-old post-
90
menopausal woman (broken line). The upper
and lower brackets represent 95% confidence
80 intervals for normal blood pressure ranges in
0:00
2:00
4:00
6:00
8:00
10:00
12:00
14:00
16:00
18:00
20:00
22:00
0:00

sexagenarian women.
Reprinted from Borer, Halberg, Cornelissen, 2003.
Time (clock hours) Unpublished data.

E4517/Borer/Fig3.8/401532/alw/R2

24

22
Skin
20

18 Heart, brain

16
Cardiac output (L/min)

Viscera
14

12
Muscle
10

4
Figure 3.9 Changes in cardiac output
2 and regional blood flow at different exercise
intensities.
0 Reprinted, by permission, from L.B. Rowell, 1974, “Human
0 0.25 1 2 3 . 4 cardiovascular adjustments to exercise and thermal
O2 uptake (L/min) VO2max stress,” Physiological Reviews 54: 75-159.

E4517/Borer/Fig3.9/401533/alw/pulled-R1
63
64 Advanced Exercise Endocrinology

S and Hormonal Vasoconstriction During Exercise


Arterioles and systemic arteries and veins are innervated by postganglionic S neurons
without a corresponding PS innervation. The preganglionic S nerve fibers synapse with
postganglionic fibers in the paravertebral chain ganglia along their entire length. NE from
S neural terminals induces contraction of smooth muscles in arterioles in the viscera, iris
(see figure 3.4), kidneys, and skin by acting on α1 and α2 AdRs. Combined activation of β1
receptors on cardiac muscle and α receptors on visceral vasculature during exercise leads
to a striking redistribution of blood from peripheral vascular beds to the contracting skeletal
muscle to ensure an increase in blood supply to the muscle (figure 3.9). Parallel increases in
exercise intensity and S activation pro-
duce graded increases in HR, systolic 125
blood pressure, pulse pressure (differ-
100 *

▲heart rate (bpm)


ence between systolic and diastolic
*
pressures), and mean arterial pressure 75
(diastolic pressure plus one third of the
50 *
pulse pressure), whereas vasodilatation
in contracting muscles accounts for the 25 *
absence of changes in diastolic pressure
(figure 3.10). 0
During exercise and hypotensive 100
▲systolic BP, diastolic BP (mmHg)

challenges, S nerves activate several *


75 Systolic BP
hormones that increase blood pres-
Diastolic BP *
sure. One of those is renin, the first 50
messenger in the renin–angiotensin– *
25
aldosterone reflex that, during the
nonstressed state, responds to declines 0
in the pressure of the blood flowing
−25
through the kidney, and during exercise
responds to S activation, to activate cor- −50
rective increases in blood pressure (see
125
chapter 4). Renin is secreted by cells in
▲MAP, pulse pressure (mmHg)

the JGA; the rate of its secretion signifi- 100 *


MAP
cantly increases at exercise intensities Pulse pressure
75
greater than 60% of maximal effort
*
(Tidgren et al., 1991). The same relative 50
effort is the threshold exercise intensity *
25
for increases in four other hydromineral * *
hormones, all of which help raise blood 0
pressure during exercise: angiotensin
−25
II, ANP, aldosterone, and AVP or ADH Pre 25% 40% 60% 80%
.
(figure 3.11). Workload (VO2max)
Angiotensin II contributes an
approximately 10 mmHg increase in Figure 3.10 Proportional increases in HR, systolic blood
systolic blood pressure and accelerates pressure, mean arterial pressure, and pulse pressure with
increases in exercise intensity.
heart function during maximal exercise
Adapted, by permission, from B.J. Freund, et al., 1991, “Hormonal,
or other emergencies (e.g., hemorrhage) E4517/Borer/Fig3.10/401537/alw/pulled-R2
electrolyte, and renal responses to exercise are intensity dependent,”
without influencing sodium reabsorp- Journal of Applied Physiology 70: 900-906.
Autonomic and Hormonal Control of the Cardiorespiratory System 65

tion in the kidney. By acting on AT1 recep- 25

▲angiotensin ll (pmol/L)
tors, angiotensin II produces sustained 20
increases in blood pressure during exercise 15
in at least four ways (figure 3.12). First,
10
it enhances S neural discharge from the
5
RVLM nucleus in the MO by acting on its
receptors in the paraventricular hypotha- 0
lamic nucleus (PVN) (Brooks, 1997; Steck-
elings et al., 1992). Second, it attenuates

▲ANP (pg/ml)
120
baroreceptor control of the heart. By attenu-
ating the reflex inhibition of heart function 80
to increases in blood pressure and shifting 40
the receptor operating range to higher blood
0
pressure values, angiotensin II instigates
simultaneous increases in both HR and
▲PRA (ng ANG ll/ml/h)
blood pressure during intense exercise 9
(Bishop et al., 1995). Third, angiotensin II 6
is a powerful vascular vasoconstrictor that
can increase systolic pressure by 50 mmHg 3
at a concentration as low as 1 millionth of 0
a gram. Finally, angiotensin II elicits the
release of the endothelial vasoconstrictor
▲aldosterone (ng/dl)

endothelin, ET-1 (d’Uscio et al., 1998). 20


At exercise intensities greater than 60% 15
of maximal effort, when the glomerular 10
filtration of the kidney is suppressed (see
5
chapter 4), AVP no longer facilitates antidi-
0
uresis and instead increases blood pressure.
At rest, AVP secretion is under tonic inhi-
4.0
bition by high-pressure carotid and aortic
▲AVP (µU/ml)

3.0
baroreceptors and by the left atrial stretch
receptors (figure 3.13) that signal increases 2.0
in blood pressure to the brain stem cen- 1.0
ters through vagal and glossopharyngeal 0
afferent nerve discharges (see figures 3.1 −1.0
and 3.15). The stimuli for AVP secretion Pre 25% 40% 60% 80%
.
Workload (VO2max)
during strenuous exercise are as follows:
attenuation of tonic baroreceptor inhibition
Figure 3.11 Release of angiotensin II, atrial natri-
over. AVP action on blood pressure at 90% ureticE4517/Borer/Fig3.11/401538/alw/pulled-R3-kh
peptide (ANP), renin, aldosterone, and arginine
of VO2max if there is a concurrent 10% vasopressin (AVP) during exercise of different intensi-
loss in PV and increase in angiotensin II ties.
concentration (Convertino et al., 1981); Adapted, by permission, from B.J. Freund, et al., 1991, “Hormonal,
increased plasma osmolarity, because both electrolyte, and renal responses to exercise are intensity
dependent,” Journal of Applied Physiology 70: 900-906; Data from
the osmolarity and AVP concentration Tidgren et al. 1991.
increase in parallel with increases in exer-
cise intensity (see figure 3.11) and dehydration; loss of PV; and metaboreflexes that trigger
myelinated and unmyelinated pain afferents and that are caused by compromised muscle
blood flow during isometric exercise.
66 Advanced Exercise Endocrinology

Angiotensin II

Thirst
Sodium hunger Aldosterone

Arteries

AT1
SFO
Vasoconstriction
AP
NTS
Kidney

Baroreflex
attenuation Sodium reabsorption

Figure 3.12 Biological actions of angiotensin II. During exercise, angiotensin acts on the brain and the smooth
muscle of the blood vessels to increase systolic blood pressure. At rest, it acts primarily on the kidney to reabsorb Na.
Reprinted, by permission, from K. Borer, 2003, Exercise endocrinology (Champaign, IL: Human Kinetics), 71.

AVP increases blood pressure by


Change in mean arterial pressure (mmHg)

125
enhancing neurotransmissionE4517/Borer/Fig3.12/401539/alw/pulled-R2
in S ganglia
(Peters & Kreulen, 1985) and by constrict-
100
ing the arterioles through its action on V1
receptors (Altura & Altura, 1984). AVP
also accelerates HR and increases blood 75
pressure during intense exercise through
stimulation of anterior hypothalamic neu- 50
rons and synergistic action with angiotensin
(Freund et al., 1988). AVP pressor actions 25
during intense stress are transient, in con-
trast to strong and sustained increases in
0
blood pressure by angiotensin II and cat- 0.026 0.26 2.6 26.0 260
echolamines (Ponchon & Elghozi, 1997). Infused AVP (ng · kg−1 · min−1)
Endothelin (ET-1) and angiotensin II are 1.7 10 42 130 250
two principal endothelial vasoconstrictors Plasma AVP (pg/ml)
that contribute in a dose-dependent fashion
to reduced blood flow in vascular beds other E4517/Borer/Fig3.13/401540/alw/pulled-R1
Figure 3.13 Vasopressor effect of AVP. Tonic inhibi-
than muscle during exercise of increasing tion of AVP or ADH pressor effects through vasocon-
intensity (Maeda et al., 2004; Maiorana et striction (solid diamonds) is reduced by deafferentation
al., 2003). of high-pressure baroreceptors (open circles) or removal
of their brain connections (solid squares). Dark shading
Exercise Hyperaemia indicates the normal range of AVP concentration.
Exercise hyperaemia, the increase in blood Adapted, by permission, from A.W. Cowley, E. Monos, and A.C.
Guyton, 1974, “Interaction of vasopressin and the baroreceptor
flow to the exercising muscle, occurs rap- reflex system in the regulation of arterial blood pressure in the dog,”
idly. The early phase of exercise hyperaemia Circulation Research 34(4): 505.
Autonomic and Hormonal Control of the Cardiorespiratory System 67

is completed in 5 to 7 s, and a more protracted increase starts after 20 s (Tschakovsky &


Joyner, 2008). The increase in muscle blood flow matches the metabolic needs of the muscle
and may exceed 100 to 300 ml of blood·100 g of muscle-1·min-1 (Joyner & Halliwill, 2000).
Despite a great deal of interest in what causes vasodilatation in exercising skeletal muscle,
this mechanism is not yet fully understood. A number of local metabolic, mechanical, and
hormonal factors certainly contribute to it. Researchers are relatively certain that in humans,
unlike in subprimate mammals, S cholinergic nerves that act on MR are not responsible for
exercise hyperaemia (Joyner & Halliwill, 2000; Tschakovsky & Joyner, 2008).
Among the local candidate metabolic triggers of exercise hyperaemia are K, adenosine,
prostacyclin, prostaglandin, bradykinin, and NO, all of which increase in concentration in
muscle extracellular space during exercise. Adenosine is a potential vasodilator in muscle
during exercise because its biosynthetic enzyme that uses AMP as substrate (AMP 5'-nucleo-
tidase) increases during exercise. Infusion of adenosine into the leg artery produces strong
vasodilatation. In addition, two of its three purinergic receptors, A2A and A2B, are present
in skeletal muscle (Rädegran & Hellsten, 2000). Vasodilatation that is initiated in resistance
arterioles that are sites of increased metabolic demand is thought to propagate upstream to
larger conduit arteries.
Of great interest is the possibility that exercise-associated release of NO is responsible
for exercise hyperaemia; this lipid-soluble gas has a potent vasodilatatory capacity. In the
muscle, NO is synthesized by two enzymes: neuronal (nNOS) and endothelial (eNOS) NO
synthetase. The synthesis requires L-arginine as a substrate and is catalyzed by NOS in
the presence of oxygen and cofactors NADPH, FAD, FMN, heme, and tetrahydrobiopterin
and the enzyme CaM (figure 2.3) (Stampler & Meissner, 2001). Vascular tone at rest is
established by a balance between the antagonistic influences of S vasoconstriction and
NO vasodilatation. However, the role of NO in vasodilatation in the exercising muscle is
presently unclear. The current view is that NO is not indispensable but may contribute to
vasodilatation and increased blood flow in the muscle in at least four ways (figure 3.14).
The first three mechanisms involve activation of eNOS through rhythmic contraction
and relaxation of muscles during exercise (the muscle pump; figure 3.14a), shear of the
faster-flowing blood through the vascular endothelium (figure 3.14b), and release of ATP by
the deoxygenated hemoglobin in the RBC, which then binds to the endothelial purinergic
receptors (figure 3.14c). The fourth mechanism entails activation of nNOS by the activity
of motoneurons and of eNOS by acetylcholine spilling over from the motor nerve endings
(figure 3.14d). The controversy regarding the role of NO in muscle vasodilatation during
exercise is in part caused by methodological challenge of rapid dilution of NOS blockers
infused into arteries during increased blood flow associated with muscle contraction.
Additionally, species differences exist in vasodilatatory control, muscle fiber differences
in distribution of nNOS (which is more prevalent in human Type I fibers than in Type II
fibers; Rädegran & Hellsten, 2000), and a wealth of interactions between individual local
factors, S nerves, and circulating E.
S nerves and circulating catecholamines also participate in regulation of blood flow of
contracting muscle during exercise. Some residual S vasoconstriction is preserved in skel-
etal muscles during exercise to maintain adequate systolic blood pressure with a limited
amount of blood. NO suppresses this noradrenergic vasoconstrictor tone by acting on α2
AdRs (Stampler & Meissner, 2001). Increases in systolic blood pressure during exercise
trigger the baroreflex and thus contribute to vasodilatation in leg circulation through S
withdrawal. E contributes to exercise hyperaemia by stimulating vasodilatation via vascular
β2 AdRs. Thus, the control of blood flow during exercise is characterized by integration
and redundancy of several local and systemic factors and needs to be elucidated further.
NO

eNOS

Smooth muscle

Muscle contraction

NO
E4517/Borer/fig.3.14a/401542/TB/R1
Smooth muscle Endothelium

eNOS
Shear stress
b

NO
E4517/Borer/fig.3.14b/448542/TB/R2-alw
Smooth muscle
eNOS

ATP binding to
purinergic receptor

RBC Low pO2 or


contraction-induced
RBC deformation
c

E4517/Borer/fig.3.14c/401543/TB/R2-alw
Somatomotor
nerve

Muscle
activation
nNOS

NO

Smooth muscle
eNOS
d

Figure 3.14 Possible mechanisms for the role of NO in exercise hyperaemia.


Reprinted, by permission, from M.E. Tschakovsky and M.J. Joyner, 2008, “Nitric oxide and muscle blood flow in exercise,” Applied
Physiology, Nutrition, and Metabolism 33: 151-161. E4517/Borer/fig.3.14d/401544/TB/R2-alw

68
Autonomic and Hormonal Control of the Cardiorespiratory System 69

Respiratory Organs
S and PS nerves form a pulmonary plexus (network) at the hilum of the lung (depression where
respiratory ducts begin) and then distribute around and along the airways and pulmonary
vessels to reach the alveoli and the pleura (Bittar, 2002). Afferent sensory nerve endings
from stretch receptors are present in the bronchial and alveolar walls and pleura and from
irritant receptors in the bronchial mucosa. These afferent fibers reach the sensory vagal
nuclei in the MO close to the respiratory control center. S nerves innervating the respira-
tory organs emerge from the thoracic segments T2 to T6 of the spinal cord to synapse in
ganglia in the paravertebral S chain that lies along the spinal cord. NE from postganglionic
fibers elicits bronchoconstriction and vasoconstriction and reduced secretion from bron-
chial glands by acting on alpha AdRs while E induces bronchodilatation, vasodilatation of
pulmonary and bronchial blood vessels, and reduced secretion from bronchial glands by
acting on β2 receptors.
In contrast to the S nerve activation of the heart and blood vessels, circulating E, rather
than NE from S nerve endings, is the principal activator of the respiratory system during
physical activity. Through its broncho- and vasodilatatory actions, E acting alone in a rest-
ing individual increases both ventilation and perfusion of the respiratory tree. Perfusion is
increased proportionately more than ventilation during E infusion, resulting in approximately
5 L of air to 8 L of blood-1·min-1. This represents a 60% reduction of the ventilation:perfusion
ratio and reduces the customary level of blood oxygenation. The combined activation of S
nerves, which cause bronchoconstriction and vasoconstriction, and the release of E during
exercise prevent the reduction of blood oxygenation during exercise compared to what would
result from the operation of E alone. An early and brief apnea and, in sensitive individu-
als, exercise-induced asthma results from the bronchoconstrictive effects of NE. The short
postganglionic PS neurons of the vagus trigger contraction of bronchial and bronchiolar
muscles and can thus precipitate an asthmatic episode. They also increase secretion of
bronchial glands and stimulate vasodilatation. A delayed increase in adrenal E secretion at
higher exercise intensities augments stimulatory effects of S NE and enhances ventilation
by causing bronchodilatation.

CONTROL OF CARDIORESPIRATORY FUNCTION BY


ANS AND HORMONES
The autonomic nerves coordinate the action of the heart, blood vessels, and respiration in
order to maintain steady blood supply to the heart and the brain. This is important in sev-
eral respects. First, because we live in a gravitational field, changing posture from supine
to upright affects the distribution of blood in circulation, which needs to be corrected to
ensure that important organs—the brain and the heart—subjected to reduced blood supply
when we stand up receive the necessary supply of oxygen. Second, breathing rate needs to
be adjusted to cardiovascular function at rest and particularly during exercise so that the
concentration of oxygen matches the changes in blood flow. Third, the same reflex that
regulates blood pressure during postural shifts needs to be reset to allow for the mainte-
nance of higher blood pressures during increased oxygen needs when we exercise. (How
the ANS helps produce these adjustments is discussed in turn.) Finally, a positive-feedback
endocrine mechanism sustains adrenal E release for the duration of exercise or other kind
of emergency and is terminated once the stress is over.
70 Advanced Exercise Endocrinology

Orthostatic Reflex
Coordination of HR and vascular diameter to maintain steady blood pressure is important
in bipedal humans as they switch from supine to upright position. When blood pressure
deviates outside of normal range during rest, an orthostatic reflex returns it to the normal
range. Gravitational pull drains blood from the upper regions of the body when we stand
up. Changes in blood pressure in the head and chest are detected by carotid and aortic baro-
receptors, which send this information to the NTS in the 9th and 10th PS afferent nerves
respectively, (see figure 3.2). When we stand up, the discharge rate from baroreceptors in
the carotid and aortic sinuses decreases in proportion to the decreases in blood pressure
(figure 3.15, left). The NTS integrates this information and initiates reflex orthostatic
correction through the activation of S discharge (shown in the left third of figure 3.15).
S activation of the heart and constriction of blood vessels increases the blood pressure to
the heart and the brain. The converse adjustment (shown in the right third of figure 3.15)
is reflexively achieved when blood pressure in carotid and aortic sinuses increases. Then
the S tone reflexively withdraws and vagus discharges from the NA increase. The result is
deceleration of HR and relaxation of vascular smooth muscles.

Blood pressure
lowered

Normal Elevated

Baroreceptors Afferents in
carotid sinus nerve
Vagus nerve

Heart rate Heart rate


raised lowered

S nerves to heart

Contractility Contractility
increased reduced

S nerves to vessels

Vessels Vessels
constricted dilated
Effect of reflex: Blood pressure Blood pressure Blood pressure
rises constant falls

Figure 3.15 The baroreflex regulates systemic blood pressure through adjustments in HR and peripheral
resistance. The discharge rate of baroreceptors is proportional to arterial blood pressure and efferent vagal firing,
whereas in the S cardiac and vascular nerves it is inversely proportional.
Springer and Autonomic functions in human physiology, 1985, page 125, G. Thews and P. Vaupel, figure 4.20, with kind permission from
Springer+Business Media B.V.
E4517/Borer/Fig3.15/401534/alw/pulled-R1
Autonomic and Hormonal Control of the Cardiorespiratory System 71

Coordination of Cardiovascular and Respiratory Functions


Autonomic reflex control of the heart pump is closely coordinated with the pulmonary
pump because circulatory transport of respiratory gases needs to be closely coordinated
with gas exchange and the perfusion of the lungs and tissues. During inactivity, these
reflexes homeostatically support resting metabolism, but during exercise, cardiorespira-
tory responses need to be increased in a coordinated fashion to provide additional supply
of oxygen. Control of the heart, blood vessels, and respiratory organs is exerted primarily
by a number of nuclei in the MO, identified in figures 3.1 and 3.3, and to a lesser extent
in the pons (Gebber, 1990). Their proximity reflects close coordination of cardiovascular
(left portion of figure 3.16) and respiratory (right portion of figure 3.16) control (Richter &
Spyer, 1990). Some cell groups, such as the noradrenergic A5 nucleus (section B in figure
3.3) in the RVLM and associated A1 and C1 adrenergic nuclei (the pressor area), possess
inherent rhythmicity and are responsible for rhythmic activation of the heart and inspira-
tion and for generation S vasomotor tone. The output from these brain centers activates the
dorsal and ventral parts of the PVN (figure 3.3) that control discharges of preganglionic
S neurons located in the IML segment of the spinal central gray matter that are synchro-
nous with the neural discharges in the RVLM (figures 3.3 and 3.16). They simultaneously
inhibit the respiratory and cardiorespiratory depressor centers that include DMV and the
more ventrally located NA. The motoneurons to the diaphragm exiting from cervical spine
segments C3 to C6, motoneurons activating intercostal muscles and arising out of thoracic
segments T1 through T12, and thoracolumbar spinal motoneurons to abdominal expiratory
muscles are all in close synaptic contact with S preganglionic neurons in the IML spinal
gray matter (figure 3.16, bottom section). In this way, pontine and medullary nuclei impose
and synchronize rhythmicity of HR and breathing.
Control of respiratory rhythm entails an interaction between pontine and medullary
neurons clustered in the dorsal (DRG) and ventral (VRG) respiratory groups (figure 3.16,
right side of brain sections; Richter & Spyer, 1990). In the dorsal pons (figure 3.16, top sec-
tion), locus coeruleus, parabrachial, and Kölliker-Fuse nuclei form a pneumotactic center
or DRG. In the MO, DRG also includes the dorsal group of premotor respiratory neurons
and interneurons in the NTS that control reflex respiratory rhythm and are influenced by
afferents from arterial baroreceptors, respiratory chemoreceptors, and lung stretch recep-
tors. These neurons connect with and modulate respiratory patterns of VRG neurons located
just posterior to the entry of the facial (seventh) cranial nerve. The VRG that is responsible
for generating inspiratory rhythms includes rostral and caudal respiratory nuclei and the
pre-Bötzinger and Bötzinger complexes (figure 3.16, right side of second brain section).
Functional connection between synchronous timing of heart contractions and inspiration
is based on the RVLM (figure 3.16, left side of second brain section) being located slightly
posterior to, and possibly functionally coupled to, the Bötzinger complex.
Control of diastole and expiration is mediated by brain regions activating the NA. The NA,
the site of depressor vagal outflow to the cardiovascular and respiratory organs (figure 3.1),
lies immediately dorsal to the VRG and exerts a tonic inhibitory influence over respiration
(figure 3.16, left side of second brain section). Detection of increases in pCO2 and H+ and
decreases in pO2 by peripheral chemoreceptors in carotid bodies and the aortic arch and by
chemosensitive neurons on the ventral surface of the medulla and in the RVLM triggers the
chemoreflex that homeostatically regulates arterial gas and H+ concentrations by increasing
the depth and rate of respiration. Peripheral chemosensitive afferents account for one-third
of the respiratory drive, and central chemosensitive neurons account for two-thirds.
72 Advanced Exercise Endocrinology

Locus coeruleus
Dorsal respiratory group
(DRG)
Lateral parabrachial and
..
Kollicker-Fuse nuclei

MLF
Nucleus ambiguus NA ..
Botzinger neurons
RP
Rostral ventrolateral
neurons (RVLM) IO
P

Baroneurons Solitary nucleus


(NTS)
Vagal cardiomotor II
neurons (DMV)

Ventral respiratory
Nucleus ambiguus group neurons (VRG)

Caudal ventrolateral Bronchomotor neurons


neurons

IML Intercostal
motoneurons

Figure 3.16 Medullary nuclei that control cardiovascular function (left) are adjacent to nuclei that control respira-
E4517/Borer/fig.3.16/401541/TB/R3-kh
tory function (right). The entry of the seventh cranial nerves is seen in the pontine (top) section.
Adapted from CENTRAL REGULATION OF AUTONOMIC FUNCTIONS, edited by A.D. Loewy & K.M. Spyer, copyright © 1990 by Oxford
University Press, Inc. Used by permission of Oxford University Press, Inc.

Role of Central Command in Regulation of Blood Pressure


During Exercise
As mentioned previously, baroreflex plays an important role in matching S vasoconstric-
tion to the control of HR in order to homeostatically regulate mean arterial pressure under
ordinary postural perturbations. However, during exercise, systolic and mean arterial
pressures increase above homeostatic levels in proportion to exercise intensity (figure
3.10). Clearly, during exercise, baroreflex does not prevent significant and proportional
increases in HR and in systolic blood pressure that are observed with increases in work
Autonomic and Hormonal Control of the Cardiorespiratory System 73

output (Freund et al., 1991). Why this

Mean systemic arterial pressure (mmHg)


145 240 w
happens has to do with the timing of
180 w
activation of different ANS centers
120 w
relative to afferent feedback from the 125
peripheral organs affected by exercise.
It is intuitively appealing to consider
60 w
that exercise first elicits S activation 105
because we are aware of anticipatory
increases in HR and respiration when
we are getting ready to exercise or are 85
facing a stressful situation. However,
these functions can also increase in Rest
65
response to inhibitory PS influence. 30 50 70 90 110 130 150 170
Much evidence suggests that PS with- Calculated carotid sinus pressure (mmHg)
drawal precedes S activation during
exercise, a concept that has been Figure 3.17 The shift of baroreflex operating point (symbols)
to higher blood pressures as exercise intensity increases from
called “central command” (Rowell,
60 to 240 watts.
1993). At the onset of exercise, the E4517/Borer/Fig3.17/401535/alw/R1
Reprinted, by permission, from Y. Papelier et al., 1994, “Carotid baroreflex
operating point of baroreflex (blood control of blood pressure and heart rate in man during dynamic exercise,”
pressure that triggers compensatory Journal of Applied Physiology 77: 502-506.
decreases in cardiac output) is reset to
a higher blood-pressure level by this shift in sympathovagal tone so that prevailing blood
pressure is perceived as a hypotensive error signal (figure 3.17; Papelier et al., 1993). The
near-instantaneous increase in HR at the onset of exercise and an increase in S activation
only after the heart has accelerated to about 110 beats per min has been attributed to the
operation of central neural command (vagal withdrawal) in these processes. According to
this concept, cardiorespiratory activation is simultaneous with the onset of physical activity.
The cardiorespiratory activation responds to the withdrawal of vagal neural outflow from the
NA to the medullary cardiorespiratory centers NTS and the rostral and caudal ventrolateral
medulla, and the RVLM responds to activation by the mesencephalic cuneiform nucleus and
lateral and posterior hypothalamic motor areas that are responsible for initiation of movement.

Maintenance of E Release by Adrenal Hormonal Positive


Feedback
Maintenance of high adrenal medullary output of E during intense exercise or other type of
emergency is maintained by positive feedback between the cortisol secretion of the adrenal
cortex and medullary E release (figure 3.18). As mentioned in chapter 1 (figure 1.6), synthesis
of E from NE in adrenal medullary cells is mediated by the stimulatory action of cortisol on
the E biosynthetic enzyme PNMT. This is facilitated by the vascular arrangement between
the cortex and the medulla as the arteries that supply adrenal cortex coalesce into venous
sinuses in the cortical zona reticularis and drain into the medulla, where they again coalesce
with medullary capillaries. When cortisol secretion is increased during intense or sustained
stress, it increases biosynthesis of E in the adrenal medulla. E, in turn, stimulates pituitary
ACTH secretion, and ACTH stimulates cortisol release. The positive feedback between
the two stress hormones is terminated once intense exercise or other stress subsides. This
reduces cortisol secretion. Its degradation in less than an hour reduces stimulation of E
biosynthesis, and the positive-feedback loop is broken.
PVN

T10
T11
CRF T12
L1
Median Hypophyseal Cortisol
eminence portal vessel Spinal cord

S trunk

Anterior
pituitary
NE E

ACTH

Adrenal gland

Figure 3.18 Autonomic and neuroendocrine


E4517/Borer/fig.3.18/401536/TB/R1
control of stress hormone secretion. The dorsal and ventral parts
of the PVN control preganglionic S nerve outflow in thoracicolumbar spinal cord including the activation of adrenal
medullary E release. The medial parvocellular part of PVN controls cortisol secretion from the adrenal cortex by
secreting CRF and thereby stimulating the release of ACTH from the anterior pituitary. During stress, E and ACTH
secretion is maintained through a positive-feedback loop in that cortisol stimulates E biosynthesis while E stimulates
the pituitary release of ACTH.
From CENTRAL REGULATION OF AUTONOMIC FUNCTIONS, edited by A.D. Loewy & K.M. Spyer, copyright © 1990 by Oxford University
Press, Inc. Used by permission of Oxford University Press, Inc.; PVN drawing reprinted by permission from K. Borer 2003, Exercise
endocrinology (Champaign, IL, Human Kinetics), 90.

74
(continued)

75
(continued)

Table 3.1 Clinical Preparations That Are Approved or Banned by the


United States Olympic Committee
Chemical name Trade name Receptor action Biological action
Allowed preparations
Albuterol Proventil, Ventolin β2 AdR agonist Bronchodilator
Clenbuterol β2 AdR agonist Bronchodilator
Fenoterol β2 AdR agonist Bronchodilator
Fluticasone Serevent β2 AdR agonist Bronchodilator
Formoterol Foradil β2 AdR agonist Bronchodilator
Indacaterol β2 AdR agonist Bronchodilator
Metaproterenol Alupent β2 AdR agonist Bronchodilator
Orciprenaline Alupent β2 AdR agonist Bronchodilator
Pirbuterol Maxair β2 AdR agonist Bronchodilator
Procaterol β2 AdR agonist Bronchodilator
Salbutamol Ventolin β2 AdR agonist Bronchodilator
Salmeterol Serevent β2 AdR agonist Bronchodilator
Terbutaline β2 AdR agonist Bronchodilator
Theophylline Slo-bid
Banned preparations containing
Ephedrine α1 and β2 AdR Decongestant,
agonist, α1 hypertensive
antagonist
Phenylephrine Afrin, Alka-Seltzer α1 and β2 AdR Decongestant,
Plus, Naldecon, agonist hypertensive,
Neo-Synephrine, pupillary dilatant
Sinex, Sudafed
Phenylpropanolamine Chlor-Trimeton, α1 and β2 AdR Decongestant,
Contac, Dimetapp, agonist hypertensive
Naldecon
Pseudoephedrine Actifed, Advil Cold α1 and β2 AdR Decongestant,
& Sinus, Ambenyl, agonist hypertensive
Benadryl, Dristan, addictive
Entex, Nyquil,
Seldane, Tavist,
Tylenol

76
Autonomic and Hormonal Control of the Cardiorespiratory System 77

SUMMARY
This chapter outlines the important role of the ANS and hormones in the control of car-
diorespiratory function during exercise and describes the pontine and medullary centers
that receive afferent information from peripheral baro-, chemo-, and mechanoreceptors.
Activation by RVLM of cardiovascular reflexes, timing of systole, and increases in blood
pressure are shown to be related to the initiation of inspiration by dorsal and ventral medulla
respiratory centers. Afferent information from lung stretch receptors and the mechano- and
chemoreceptors is also shown to control the PS depressor function that initiates diastole
and expiration. The importance of dual innervations of the heart, vasculature, and respira-
tory organs by the S and PS divisions of the ANS is discussed for a better understanding of
the activation of these organs during exercise. The systemic hormones renin, angiotensin
II, and AVP are shown to make an important contribution to increases in blood pressure
during exercise. Anatomical and functional coupling of the ANS centers that are responsible
for the control of heart function, vascular conductance, and respiration are described and
shown to be responsible for the synchronization of cardiovascular and respiratory functions
during exercise. This chapter also presents evidence that withdrawal of vagal tone through
action of the CNS at the onset of exercise is responsible for initial acceleration of HR and
peripheral vasoconstriction. Finally, a positive-feedback loop between cortisol and E is
described to account for the maintenance of emergency actions of the two stress hormones
during intense exercise.
This page intentionally left blank.
H APTER

4
C

Body Fluid Balance


W
ater is the only essential nutrient that cannot be stored in excess in the body.
Water plays four key physiological roles at rest and particularly during exercise.
Its first and universal function is to provide a medium for life-sustaining meta-
bolic cellular processes. Water makes up 71% of lean tissue cellular mass (ICF), and TBW
makes up about 60% and 52% of body weight in young men and women, respectively. The
remaining body water is extracellular (ECF), of which 79% is interstitial fluid surrounding
the cells (ISF) and 21% is plasma or the liquid portion of the blood (PV). Second, the water
component of plasma provides a route for delivering oxygen, nutrients, hormones, and other
essential molecules to cells throughout the body and for sending metabolic waste to the kid-
neys. This is accomplished by circulating blood. The third role of body fluids is to maintain
adequate pressure in the vascular system so that vital organs such as the brain and muscles
are supplied with the necessary molecules despite variation in the gravitational pull caused
by postural changes or increases in energy demands during exercise. The final vital function
of water is to allow for dissipation of heat load that accumulates in the body due to either
high ambient temperatures or addition of heat generated by muscle contraction. Heat load
is dissipated by the diversion of some of the circulating blood to the skin for heat exchange
by radiation and through water vaporization from the skin surface in the form of sweat.
Exercise presents a challenge to body fluid balance because the relative priorities for
the uses of body water shift. During exercise, accumulation of metabolic heat can be both
rapid and massive and become life threatening if Tc (core temperature) surpasses 40 °C.
Therefore, the thermoregulatory functions of body water can become more important than
maintenance of cellular hydration and PV. This conflict in the need for water during exercise
is exacerbated by a reduction or cessation of kidney functions involved in preserving water
and Na and by slow and relatively inefficient sensing by humans of the need for water and
Na after exercise. This chapter provides information from an endocrine and autonomic per-
spective on the thermoregulatory shifts and losses of body water as body heat load increases
and kidney function decreases during exercise and on the effects of cellular dehydration and
hypovolemia after exercise on thirst, Na hunger, kidney function, and body fluid restora-
tion. A better understanding of these mechanisms will allow one to develop individualized
strategies for preserving and restoring body fluids during and after exercise.

79
80 Advanced Exercise Endocrinology

INCREASES IN BODY HEAT LOAD DURING EXERCISE


ATP production for muscle contraction is only about 30% efficient (Bangsbo et al., 2001).
The remaining 70% of energy is converted to heat in the muscle and is transferred to the
blood. Because the specific heat of the human body is 0.83 kcal/kg, body temperature would
rapidly increase to lethal levels above 40 °C at moderate to high exercise intensities if no
compensating mechanisms existed. The two S reflexes that lead to heat dissipation and
reduce increases in core body temperature are the vasomotor reflex that increases blood
flow to the skin for radiative heat loss and the sudomotor reflex that results in body heat
loss through the evaporation of sweat. Because diversion of blood from exercising muscle
to the skin and the loss of body fluids through sweating compromise circulatory delivery
of oxygen and nutrients to the exercising muscle, the two reflexes do not produce perfect
thermoregulation. Rather, they attenuate rises in Tc associated with increases in exercise
intensity and energy metabolism. Tc increases about 0.75° for each 1 L/min increase in
oxygen consumption (Nielsen, 1938).
Physical performance declines with increases in Tc and is ended by fatigue at about 40.2
to 40.3 °C (Gonzalez-Alonso et al., 1999). Above this temperature, heat injury may impair
brain function and lead to death (Carter, 2008). Therefore, prevention or slowing of overheat-
ing through peripheral vasodilatation and sweating becomes a high priority during exercise.

THERMOREGULATORY CHANGES IN BODY WATER


DURING EXERCISE
Heat losses due to vasomotor and sudomotor reflexes are lower than exercise-induced accu-
mulation of body heat load. Therefore, Tc progressively increases as a function of exercise
intensity. Of the two heat-loss reflexes, sweating dissipates body heat to a much greater
extent than does radiative heat loss from the skin (figure 4.1).

Energy output
800
1
700
Heat production
600
2
500
Kcal/hr

3 Total heat loss


400
Evaporative heat
4
300

200
Convective and
5
radiative heat loss
100 6
Evaporative heat
0 from the lungs
0 30 60 90 120 150 180 210 240
Energy expenditure (watts)

Figure 4.1 Relationship between energy expenditure, heat production, and heat loss attributed to evaporative
sweating and radiation during exercise of increasing intensity.
E4517/Borer/Fig4.1/401546/alw/R2
Adapted, by permission, from P-O Ästrand, 2003, Textbook of work physiology: Physiological bases of exercise, 4th ed. (Champaign, IL:
Human Kinetics); Adapted from Nielson 1938.
Body Fluid Balance 81

Skin Blood Flow


Vasodilatation, increased blood flow to the skin, and sweating during exercise are under
dual control of both skin and Tc receptors, which relay the information to the anterior hypo-
thalamus and central command neurons in the motor cortex during exercise. The anterior
hypothalamic nuclei integrate the temperature information and initiate thermoregulatory
reflexes through the S neural outflow to the skin, a final common pathway for heat-loss
reflexes at the onset of exercise. Increases in skin blood flow in response to increasing Tc
during exercise are mainly mediated by a cotransmitter in cholinergic S nerves that does
not act on cholinergic MR (Joyner & Halliwill, 2000; Kellogg et al., 1995). Thus, increases
in skin blood flow due to increased Tc can be prevented by either sympathectomy (Roddie,
1983) or a chemical presynaptic blockade of cholinergic neurotransmission but not by MR
blockade with atropine (Kellogg et al., 1995). The usual cotransmitters in S nerves are
ATP and NPY and in cholinergic PS nerves, VIP. However, the identity of the transmitter
involved in active skin vasodilatation during exercise is unknown at the present. Activation
of nNOS during muscle contraction contributes only about 20% to 30% of the increases in
cutaneous circulation (Joyner & Halliwill, 2000) and thus is unlikely to be the major cause
of an exercise-associated thermoregulatory increase in skin blood flow.
The increase in skin blood flow facilitates heat loss during ambient heat and onset of
exercise and is reduced when increases in Tc are accompanied by cellular dehydration.
Reduction in skin blood flow appears to be triggered by increases in plasma osmolarity rather
than by PV losses because normalizing plasma osmolarity by consuming water increases
cutaneous blood flow whereas intravenous PV expansion does not (Montain & Coyle, 1992a).

Sweating
Sweat secreted via eccrine glands is essential to human thermoregulation in the heat. Whole-
body sweat rates from a single eccrine gland can reach 3.7 L/h and 10 nl/min (Wilson et al.,
2005). Therefore, in a hot environment humans can lose as much as 10 to 11 L/d of fluid,
a substance that they have no way of storing. Sweating from eccrine sweat glands begins
within seconds of the onset of exercise, before any increases in Tc are registered. This
suggests the mediation of sweating, in part, by central command neurons from the motor
cortex, by metaboreceptors during ischemic isometric exercise, and by S neural activation
prior to, and in addition to, the sensory input to the anterior hypothalamus from skin or Tc
(Shibasaki et al., 2006; figure 4.2). Increases in skin temperature lower the threshold for
the initiation of sweating in an anticipatory fashion to counteract increases in Tc (Nadel et
al., 1971). Stimulation of sweating in response to increases in Tc is mediated to a greater
extent by S activation of cholinergic MR (Joyner & Halliwill, 2000) rather than by S action
on alpha and beta AdRs (Sato, 1977; Shamsuddin et al., 2008). The sudomotor reflex is
endergonic and decreases body temperature by utilizing 580 kcal to vaporize 1 L of water
(Wenger, 1972).
Because S nerves also contribute to the maintenance of blood pressure during upright
posture and exercise, a hypothesis that the baroreflex and sweating response were activated
simultaneously was examined (figure 4.2) but not supported (figure 4.3; Wilson et al., 2005).
The two reflexes appear to be controlled separately.
Sweating is modulated according to the state of body hydration. Sweating produces both
cellular dehydration and hypovolemia because mean Na concentration in the sweat is between
20 and 60 mM and plasma Na concentration is 145 mM. Therefore through a relatively greater
loss of water than sodium, sweating causes hypertonicity and cellular dehydration, while
the loss of sodium from the plasma reduces PV. Na concentration in the sweat increases
82 Advanced Exercise Endocrinology

in proportion to the sweating rate and Central motor


the increasing body heat load (figure command
4.4). This is a consequence of slower Na Thermo-
reabsorption relative to Na secretion in regulatory
the sweat as the sweating rate increases center
(Buono et al., 2008).
Fluid and PV conservation pro-
cesses are activated during sweating.
Acute dehydration suppresses sweating,
whereas drinking water during a hyper-
osmotic state increases sweating (Taka-
mata et al., 1995). Losses of both plasma Eccrine
Baroreceptors sweat
water and Na decrease the sweating
glands
rate to minimize further losses of fluid
Sympathetic
and Na (Fortney et al., 1981). Cellular neurons
dehydration increases the Tc threshold
for initiation of sweating. Adaptation
to fluid losses after 10 days in the heat
lead to heat acclimation. In response
to such extended heat exposure, the
Na reabsorption capacity of the human
Muscle metaboreceptors
sweat gland increases and Na loss in
sweat is reduced by 15 mM indepen-
Figure 4.2 Neural control of sweating during exercise.
dently of the sweating rate (Buono et The motor cortex and metaboreflexes activate sweating
al., 2007). Therefore, heat-loss processes at the onset of exercise before Tc increases. Increasing
are reduced to minimize losses of water E4517/Borer/fig.4.2/401547/TB/R2-alw
body heat load also controls sweating through the anterior
and Na in response to both acute and hypothalamic temperature center. S nerves are the final
prolonged dehydration. common pathway for activation of sweating.
Reprinted, by permission, from M. Shibasaki, T.E. Wilson, and C.G.
Crandall, 2006, “Neural control and mechanisms of eccrine sweating
during heat stress and exercise,” Journal of Applied Physiology 100:
1692-1701.

1.2
Supine
Sweat rate (mg · cm−2 · min−1)

1.0 Tilt

0.8

0.6

0.4

0.2

0.0

36.8 37.0 37.2 37.4 37.6 37.8 38.0 38.2


Mean body temperature (°C)

Figure 4.3 The relationship between forearm sweat rate and mean body temperature is unaffected by the
baroreflex elicited by tilting of the body (open circles).
E4517/Borer/Fig4.3/401631/alw/R1
Reprinted, by permission, from T.E. Wilson, J. Cui, and C.J. Crandall, 2005, “Mean body temperature does not modulate eccrine sweat rate
during upright tilt,” Journal of Applied Physiology 98: 1207-1212.
Body Fluid Balance 83

70

Sweat [Na+] (mmol/L)


60
50 r = 0.73

40
30
20
10
0
0 0.2 0.4 0.6 0.8 1
Sweat rate (mg · cm−2 · min−1)

Figure 4.4 Na concentration in the sweat increases in proportion to sweating rate.


Reprinted, by permission, from M.J. Buono et al.,E4517/Borer/Fig4.4/401632/alw/R2
2008, “Na+ secretion rate increases proportionally more than the Na+ reabsorption rate
with increases in sweat rate,” Journal of Applied Physiology 105: 1044-1048.

CONSEQUENCES OF FLUID LOSS THROUGH


SWEATING
Two primary consequences of sweating are loss of PV (hypovolemia) due to greater water
losses than Na losses from the plasma compartment and cellular dehydration that results
from increases in plasma osmolarity. At rest, these body fluid changes stimulate endocrine
reflexes for conserving Na and water through the kidney. During exercise these reflexes
do not work, and the hypovolemia and cellular dehydration lead secondarily to circulatory
cardiovascular drift and restrict heat-loss processes.

Cardiovascular Drift
In addition to causing blood redistribution to skin and losses of water and Na through sweat-
ing, exercise-associated increases in body heat load result in cardiovascular drift. This drift
manifests as a decrease in stroke volume and an increase in HR that are proportional to the
degree of body fluid loss and increases in Tc (Montain & Coyle, 1992b). Like the control of
cutaneous blood flow, cardiovascular drift appears to be triggered by increases in plasma
osmolarity and Tc rather than by hypovolemia due to circulatory diversion to the skin and
decline of PV (Coyle & Gonzalez-Alonso, 2001; Montain & Coyle, 1992b).

Hypovolemia and Renal Reabsorption of Na at Rest


Na losses during sweating produce PV losses because the volume of ICF (intracellular fluid)
and ECF (extracellular fluid) compartments is determined by their ionic content and by the
semipermeability of cellular membrane to ions. Constituent ions in a fluid compartment do
not freely cross the membranes; therefore, they draw water into the compartment by osmotic
pressure. The osmolarity of the ICF compartment is on average 256 mM, 59% of which is
attributable to its K content. The osmolarity of plasma is 295 mM, 49% of which comes
from Na. Decreases in blood pressure that are associated with losses of PV are temporarily
compensated by S activation of the heart and peripheral vasoconstriction (see discussion of
baroreflex in chapter 3). A more complete correction of the lost PV requires water and salt
to be recaptured before it is excreted into urine and PV to be corrected through ingestion
of additional water and salt. Hormones angiotensin II (see figure 3.12) and AVP play a key
role in these actions.
84 Advanced Exercise Endocrinology

Na reabsorption, and to some extent water reabsorption, is mediated by angiotensin II in


a sedentary condition but not during exercise. Granular cells, the intrarenal baroreceptors,
release renin when they detect a decrease in filtration pressure in kidney arteries or a decline
in plasma Na. This is the first link in the renin–angiotensin–aldosterone endocrine reflex.
Renin-secreting cells are located in the macula densa, a contact area between the distal convo-
luted kidney tubule and the afferent arteriole, called JGA (juxtaglomerular apparatus). Renin
cleaves the prohormone angiotensinogen originating in the liver to form inactive polypeptide
angiotensin I. The endothelial ACE (angiotensin-converting enzyme) changes angiotensin I
to angiotensin II in vascular endothelia and inhibits synthesis of the vasodilator bradykinin.
Circulating angiotensin II stimulates AT1 receptors in the external zona glomerulosa of the
adrenal cortex to synthesize the mineralocorticoid aldosterone through the IP3 transduction
pathway (see figure 1.8). Decreases in PV through Na loss can increase aldosterone secre-
tion two- to sixfold. Angiotensin II also stimulates the PVN and SO hypothalamic nuclei
to release AVP for water reabsorption. Angiotensin II (see figure 3.12) is also a powerful
vasoconstrictor (discussed in chapter 3) and stimulates thirst and Na hunger by acting on
the CVOs (circumventricular organs) (discussed in “Thirst and Na Hunger After Exercise”).
Aldosterone preferentially binds to high-affinity type I mineralocorticoid receptors in
the kidney, colon, and salivary and sweat glands, where it increases Na reabsorption. In
the kidney tubule, reabsorptive action entails synthesis of alpha and beta subunits of an
Na permease pump that controls the rate-limiting step in Na reabsorption at the mucosal
surface. The Na–K pump then actively transports Na across the serosal membrane into the
ICF compartment on its way to capillaries. Aldosterone-mediated active transport of Na
establishes an electrochemical gradient that then facilitates passive K excretion into the
urine. The osmotic re-entry of water into capillaries associated with active Na reabsorption
facilitates restoration of PV and increases in blood pressure. An 11βHSD (hydroxysteroid
dehydrogenase) in the kidney, colon, and salivary and sweat glands prevents glucocorticoid
hormones, which are present in the blood in much higher concentrations than are miner-
alocorticoids, from binding to type I mineralocorticoid receptors and increasing PV and
blood pressure. This enzyme is inactivated by the glycyrrhizic acid in licorice, resulting in
significant increases in blood pressure.

Cellular Dehydration and Renal Reabsorption of Water at Rest


Because the concentration of NaCl is lower in sweat (0.2%-0.5% NaCl; Shirreffs & Maughan,
1997) than in plasma (0.9% NaCl), sweating causes increases in plasma osmolarity, which
drives water transfer from the ISF (interstitial fluids) and ICF into plasma compartments. As
plasma osmolarity increases both as a function of dehydration and exercise intensity (table
4.1; Montain et al., 1997), plasma-fluid losses remain constant at about 10% to 12% of total
fluid loss, whereas the fluid losses from the ISF and ICF compartments increase (table 4.2;
Costill et al., 1976). Plasma osmolarity increases further with water restriction, whereas the
effects of increased exercise intensity and water deprivation are additive (table 4.1).
Cellular dehydration is sensed by the osmoreceptors in the CVOs, vascular structures at
the interface between the circulatory and cerebrospinal compartments (figure 4.5). CVOs are
situated near the third and fourth cerebral ventricles, lack BBB, and are highly vascularized.
Three CVOs play an important role in preserving body water and stimulating drinking: the
OVLT (organum vasculosum of the lamina terminalis), the SFO (subfornical organ), and
the posterior pituitary gland.
OVLT and SFO are sensitive to a change in plasma osmolarity of as little as 1% to 2%.
The CVOs use the anterior hypothalamic nucleus medianus as a relay for activation of the
Body Fluid Balance 85

Table 4.1 Changes in Plasma Osmolarity in Response to 50 Min of Treadmill


Exercise at 30 °C and Dehydration
.
Exercise intensity (% VO2max)
Dehydration (% BW loss) Rest 25 45 65
0 281 284 286 287
3 286 287 291 293
5 294 294 296 299
BW= body weight.
Data from Montain et al. 1997.

Table 4.2 Distribution of Water Losses During Dehydration


As As % As %
TBW loss As % of % of PV loss of TBW ISF loss As % of ICF loss of TBW
(ml) BW TBW (ml) loss (ml) TBW loss (ml) loss
1,432 2.0 3 138 10 861 60 433 30
2,775 3.8 6 283 10 1,028 37 1,463 53
3,950 5.4 9 470 12 1,598 41 1,878 48
BW = body weight.
Data from Costill, Cote, and Fink 1976.

magnocellular neurons in the PVN and


SO nuclei for release of AVP, also called
ADH. The magnocellular neurons also
synthesize the hormone OXY, which
is involved in the milk-ejection reflex, cc
labor contractions, and some psycho- fx
logical reactions. The two hormones SFO PB
ac
are transported in the neurosecretory OVLT
vesicles in the neuronal axons through oc SCO
ME
the hypothalamo-hypophyseal stalk and
3rd
stored in axon terminals in the posterior ventricle PP AP
pituitary (figure 4.6). 4th
In response to cellular dehydration, ventricle
AVP is released from SO and PVN axon
terminals in the posterior pituitary into Figure 4.5 Location of the CVOs. Midsagittal view of
systemic efferent veins through the fenes- the human brain shows the location of CVOs (black). ac =
anterior commissure; AP = area postrema; cc = corpus
trated capillaries that characterize areas callosum; fx = fornix; ME = median eminence; oc = optic
devoid of BBB. AVP and OXY half-lives chiasm; PB = pineal body; PP = posterior pituitary.
are less than 5 min. At rest, the primary E4517/Borer/Fig4.5/401633/alw/pulled-R1
Reprinted from Neuroscience Letters, Vol. 57, S. Landas et al.,
action of AVP is antidiuresis, or produc- Demonstration of regional blood-brain barrier permeability in human
brain, 252, Copyright 1985, with permission from Elsevier Science.
tion of a smaller volume of concentrated
urine by stimulation of renal water reab-
sorption. Control by AVP of renal water reabsorption is responsive to both increases and
decreases in osmolarity (figure 4.7). Drinking a large quantity of water reduces plasma
osmolarity, suppresses AVP release, and increases urination. This is not an appropriate
strategy for restoring reduced PV.
86 Advanced Exercise Endocrinology

Paraventricular
nucleus
Supraoptic
nucleus
Hypothalamo-hypophyseal stalk
Superior
hypophyseal
artery
Posterior
Optic Median pituitary
chiasm eminence Inferior
hypophyseal
Hypophyseal artery
portal vessel

Anterior
pituitary Efferent
Intercavernous
veins
sinus
Efferent veins

Figure 4.6 Neurosecretion of posterior pituitary hormones. ADH is synthesized in paraventricular (PVN) and
supraoptic (SO) neurons, transported in their axons down the hypothalamo-pituitary stalk, and stored in axon
E4517/Borer/Fig4.6/401548/alw/pulled-R1
terminals in the posterior pituitary.
From TEXTBOOK OF ENDOCRINE PHYSIOLOGY, THIRD EDITION, edited by James E. Griffin and Sergio R. Ojeda, copyright © 1996 by
Oxford University Press, Inc. Used by permission of Oxford University Press, Inc.

AVP stimulates water reabsorption 1.6


by binding to V2 receptors on the apical
Release of vasopressin (100 mU/min)

cells in the distal convoluted kidney


tubules that are located in the external 1.2
cortical region and in the kidney tubule Dose-
Water dependent
collecting ducts that lie in the kidney
diuresis antidiuretic
medulla below the cortex (figure 4.8). response
0.8
Plasma that is filtered into tubule
lumen through the Bowman’s capsule
reaches the distal portions of the tubule Normal
after traveling and undergoing osmotic 0.4
Maximal
changes in the proximal tubule and the antidiuresis
loop of Henle. The binding of AVP to V2
receptors activates cAMP signaling and 0
phosphorylations to produce exocytotic 97 100 103 106
Plasma osmotic pressure (% normal)
insertion of aquaporin-2 protein into
the luminal plasma membrane of the
collecting duct. Figure 4.7 Control of AVP release by changes in plasma
osmotic pressure.
Hypertonicity of renal ISF is achieved
Springer andE4517/Borer/Fig4.7/401634/alw/pulled-R1
Chapman Hall, Hormones: From molecules to disease,
in the loop of Henle, where Na and Cl 1990, pg. 294, Hormones: a complex communication network, E.E.
ions are actively pumped out of the Baulieu, edited by E.E. Baulieu & P.A. Kelly, figure V1.14, with kind
permission from Springer+Business Media B.V.
ascending limb to produce increasingly
hypotonic urine and increasingly hyper-
tonic interstitial fluid (see figure 4.9). The walls of the descending loop of Henle are equipped
with aquaporin-1 molecules that provide high water permeability. Osmosis of water into
Body Fluid Balance 87

the hypertonic ISF renders the urine in


the descending limb progressively more ,
Bowman s
hypertonic. The interaction between the
Distal space
Na excretion by the ascending loop of tubule
Henle, the growing hypertonicity of the Glomerulus
descending loop, and the unidirectional
urine flow through the tubule gener- Afferent
ates an osmotic gradient in the ISF that arteriole
increases from the proximal end of the Proximal
tubule
descending loop, where it is 300 mM,
to the hairpin turn of the loop, where it
reaches 1400 mM. AVP acts to facilitate Efferent
arteriole
aquaporin water channel insertion into
the luminal wall of the collecting duct. Collecting
A central hydrophilic channel of aqua- duct
Descending
porin molecule allows water transit into loop of Henle
the hyperosmotic renal interstitium and
from there into the blood vessels of the
kidney cortex and outer medulla. Besides Ascending
loop of Henle Distal
its mediation of water reabsorption, AVP tubules
also promotes aquaporin gene expression from other
to enhance the water-absorbing capac- nephrons
ity of the kidney. Because the kidneys
excrete 600 mM of solutes daily, the Renal pelvis
smallest obligatory daily loss of fluid
through the kidney is 0.43 L. Figure 4.8 Features of a kidney tubule where AVP exerts
Reduction in PV interacts with the antidiuresis. Plasma is filtered into tubule lumen through the
increases in osmolarity to augment Bowman’s capsule. It undergoes osmotic changes in the
Henle’s loop, and the water reabsorption takes place when
the water-reabsorbing actions of AVP
AVP facilitates the insertion of aquaporin water channels
(figure 4.10). Hypovolemia lowers the E4517/Borer/Fig4.8/401549/alw/pulled-R1
into the collecting duct luminal membrane.
osmotic threshold for AVP secretion and Adapted, by permission, from A.J. Vander, J. Sherman, and D.S.
increases the sensitivity of AVP response Luciano, 1998, Human physiology: The mechanism of body function,
to increases in osmolarity. 7th ed. (New York: McGraw-Hill), 505. © The McGraw-Hill Companies;
Based on Dunn 1973.

CESSATION OF RENAL
REABSORPTIVE FUNCTION DURING EXERCISE
At rest, the principal roles of circulating blood are delivery of oxygen, nutrients, hormones,
and other vital components to the tissues and delivery of metabolic wastes to the kidneys. The
large volume of plasma that is filtered through the kidney underscores the importance of the
kidney in removing metabolic wastes and toxins and in regulating PV, plasma osmolarity,
and acid–base balance at rest. The combined weight of the two kidneys is only about 1% of
body weight, yet at rest they receive between 20% and 25% of the cardiac output. During
exercise, the relatively small total PV (3 and 2.6 L in young men and women, respectively)
has to support the increased requirement for oxygen and fuel delivery to the muscle and meet
the obligatory thermoregulatory requirement. As work effort increases, increased oxygen
and fuel delivery is achieved through increases in cardiac output; that is, the limited amount
of blood is circulated faster. At the highest exercise intensities, some S vasoconstriction is
NaCl NaCl H2O
H2O
NaCl
Descending Cortical
100 H2O
limb collecting
300
Distal tubule duct
NaCl
NaCl H2O
300 100 300 300

NaCl H2O
600 H2O 400 600
Ascending
limb Medullary
NaCl H2O collecting
900 H2O 700 900 900
duct

NaCl
NaCl
1,200 H2O 1,000 H2O
1,200 1,200

H2O
1,400 1,400 1,400

Figure 4.9 Antidiuretic action of AVP in the kidney tubule. The active transport of Na out of the ascending limb
of the loop of Henle dilutes the kidney tubule fluid and creates a gradient of Na concentration in the surrounding
interstitium. Plasma filtrate flowing unidirectionally down the descending limb of the loop of Henle is concentrated
E4517/Borer/Fig4.9/401550/alw/pulled-R1
as it loses water to hyperosmotic interstitium. The AVP response to hypertonic plasma is to activate the insertion
of aquaporin water channels into the collecting duct for reabsorption of water into the interstitium and the plasma.
Hypotonic plasma suppresses AVP secretion and allows hypotonic filtrate from the ascending limb of the loop of
Henle to be voided as dilute urine.
Adapted, by permission, from A.J. Vander, J. Sherman, and D.S. Luciano, 1998, Human physiology: The mechanism of body function (New
York: McGraw-Hill), 520. © The McGraw-Hill Companies; Based on Dunn 1973.

50

Small-
40 volume
Large- depletion
Plasma AVP (pg/ml)

volume
30 depletion

20

Euvolemia
10

0
275 285 295 305 315 325
Plasma osmolality (mOsm/kg)

Figure 4.10 Control of AVP secretion by hypovolemia and plasma osmolarity.


E4517/Borer/Fig4.10/401635/alw/pulled-R1
Adapted, by permission, from F.L. Dunn, et al., 1973, “The role of blood osmolality and volume in regulating vasopressin secretion in the rat,”
Journal of Clinical Investigation 52: 3212-3219.

88
Body Fluid Balance 89

maintained even in skeletal muscles in order to support increases in systolic blood pressure
against vasodilatation in contracting muscle. In addition, an obligatory 4% to 12% reduc-
tion in PV occurs due to fluid shift from plasma to the intracellular compartment of the
muscle, usually called exercise-induced hemoconcentration (Wilkerson et al., 1977). This
shift results from increases in systolic blood
. pressure and capillary filtration pressure . and
amounts to 3.7% of PV reduction at 40% VO2max and as much as 12.4% at 90% VO2max.
Increases in Tc necessitate diversion of some blood to the skin surface for heat loss through
radiation, and thermoregulatory sweating results in PV loss.
The limitation of small PV and its thermoregulatory losses during exercise is resolved
through S and hormonal constriction of renal arteries and the resultant suppression of
renal water loss. In addition to contribution of S stimulation of vascular alpha AdRs, ET-1
contributes to renal arterial vasocon-
striction during exercise (Maeda et
al., 2004; see chapter 3). When NE-
▲urine flow (ml/min)
1.0
induced vasoconstriction of afferent
0.5
renal arteries reduces renal blood flow
0
and GFR by more than 35%, the absorp-
−0.5 *
tive actions of hydromineral hormones *
in the kidney are reduced (Freund et −1.0
al., 1991). At lower exercise intensi- 0.5
ties, AVP can exert some antidiuretic 0
▲Cosm (ml/min)

action. Hypohydration during exercise −0.5


stimulates aldosterone and AVP secre-
−1.0 **
tion in the heat (Montain et al., 1997).
−1.5
These hormones do not engage in Na
reabsorption but can be activated by S −2.0 **
nerves and stimuli other than hypovole- −2.5
mia and cellular dehydration. Increased 1.5 *
▲CH 2O (ml/min)

S activation during exercise triggers the 1.0


*
renin–angiotensin–aldosterone reflex
5
through stimulation of beta AdRs on
0
granular cells when NE spillover in the
kidney increases. Angiotensin is formed −0.5
by a distinct serine protease pathway 50
▲GFR (ml/min)

during exercise. Aldosterone can also 0


be triggered by high concentrations of −50 *
ACTH during intense exercise or other
−100 *
forms of stress.
−150
At . exercise i ntensit ies above Pre 25% 40% 60% 80%
.
60% VO 2 max, urine f low, osmotic Workload (VO2max)
clearance, free water clearance, and
glomerular GFR decrease (figure 4.11). Figure 4.11 Effect of exercise intensity on renal function.
Therefore, during exercise of moderate Changes in urine flow, osmotic clearance (Cosm), free water
clearance (CH2O), and GFR as a function of different exercise
to high intensity, the hydromineral hor-
intensities during 20 min of cycle ergometry.
mones assume hypertensive and cardio-
Adapted, by permission, from B.J. Freund, et al., 1991, “Hormonal,
stimulatory actions rather than actions E4517/Borer/Fig4.11/401551/alw/pulled-R2
electrolyte, and renal responses to exercise are intensity dependent,”
that conserve body fluid (see chapter 3). Journal of Applied Physiology 70: 900-906.
90 Advanced Exercise Endocrinology

THIRST AND NA HUNGER AFTER EXERCISE


Drinking water and eating salty food are necessary for ultimate resolution of cellular dehydra-
tion and hypovolemia. Although we experience thirst in response to dehydration, our thirst
mechanism and specific hunger for salt are neither precise nor fast acting. A combined loss
of 4.5% of body water and 3% of body Na during exercise in a hot environment generates
a 3 mM increase in plasma osmolarity and a 6.5% PV loss. An immediate postexercise
increase in thirst is elicited by cellular dehydration and is accompanied by a fivefold increase
in plasma AVP. A rapid onset of thirst that follows exercise-induced dehydration slakes the
thirst for water within 5 h of water drinking by normalizing plasma osmolarity but corrects
only about one half of PV and body fluid deficit. Thirst sensation and plasma aldosterone
concentration return to the baseline (Takamata et al., 1994). However, a sixfold increase in
urine flow and sweating interfere with restoration of water balance. The inability to fully
correct a water deficit within a few hours after the onset of rehydration is called voluntary
dehydration. Voluntary dehydration
impedes optimal physical performance.
Exercise Rehydration
Thirst rating (% of full scale)

The immediate increase in thirst 100


after exercise-induced dehydration **
80
elicits only a 2.5-fold increase in plasma
aldosterone that returns to baseline upon 60
water drinking. Correction of hypo- 40 * *
volemia is corrected about 15 h after 20
exercise-induced dehydration when a
0
more sustained thirst and about three- −10 −5 0 5 10 15 20 25
fold increase in plasma aldosterone are
6
observed (figure 4.12) (Takamata et al., **
5
1994). Thus, hypovolemia stimulates
AVP (pg/ml)

4
thirst and renal sodium-retention reflex
3
with a delay, and usually a relatively
2
high PV loss of between 8% and 10%
1
is needed to elicit thirst by hypovole-
mia and elicitation of angiotensin II 0
−10 −5 0 5 10 15 20 25
(Fitzsimmons, 1998). Angiotensin II
Plasma aldosterone (pg/ml)

400
triggers thirst in response to changes *
in PV by binding to AT1 receptors on 300 * *
the SFO cells and by activating neuro-
nal circuits controlling drinking. This 200
high threshold of thirst elicitation by 100
angiotensin II prevents variations in PV
stimuli caused by postural changes and 0
−10 −5 0 5 10 15 20 25
movement from triggering thirst before Time after the onset of rehydration (hr)
any real PV losses are incurred.
Due to the delayed and inadequate FigureE4517/Borer/Fig4.12/401552/alw/pulled-R1
4.12 Thirst and hydromineral hormone responses
mechanism of human rehydration, con- to exercise-induced dehydration. The timing of subjective
scious strategies have been developed rating of thirst (upper panel), AVP (middle panel), and
for monitoring fluid loss during exer- aldosterone (lower panel) in response to hyperosmotic–
cise; for diagnosing the magnitude of hypovolemic dehydration induced by exercise.
the loss by measuring changes in body Adapted, by permission, from A. Takamata, et al., 1994, “Sodium
appetite, thirst, and body fluid regulation in humans during rehydration
weight, urine loss, and concentration; without sodium replacement,” American Journal of Physiology:
and for adjusting the intake of water and Regulatory, Integrative and Comparative Physiology 266: R1493-R1502.
Body Fluid Balance 91

electrolytes accordingly (Maughan & 60


Shirreffs, 2010; Sawka et al., 2007).

Palatability rating (% of full scale)


A concomitant preference for 40
a more concentrated solution of
NaCl accompanies sweating-induced 20
decl ine in plasma sod ium a nd
increases in aldosterone concentration 0
(figure 4.13). Intracerebroventricular
administration of angiotensin II has
−20
been demonstrated to increase Na
hunger (Fitzsimmons, 1998). There-
−40
fore, the inverse relationship between Control
plasma Na concentration and human
−60
preference for salt solutions (figure 148 150 152 154 156 158
4.13; Takamata et al., 1994) demon- Plasma Na+(meq/kg H2O)
strates the probable involvement of
angiotensin II in behavioral restoration Figure 4.13 Effect of Na depletion on salt preference.
of PV by an increase in Na hunger. Palatability rating of 1 M (5.5%) NaCl solution as an inverse
A joint action of angiotensin II and function of plasma Na concentration.
E4517/Borer/Fig4.13/401636/alw/pulled-r2
aldosterone on the SFO receptors is Adapted, by permission, from A. Takamata, et al., 1994, “Sodium appetite,
necessary for the development and full thirst, and body fluid regulation in humans during rehydration without
sodium replacement,” American Journal of Physiology: Regulatory, Integrative
expression of Na hunger and rebound and Comparative Physiology 266: R1493-R1502.
thirst (Fitzsimmons, 1998).

HYPERHYDRATION Central
nervous Salt appetite
AND HYPONATREMIA system Water intake

Deliberate intake of salty foods causes AVP


transient hypertonicity of plasma and
cellular dehydration thirst. Drinking RAP
fluids increases PV. Correction of
this problem calls into action a hor- Venous ANF
mone ANP (atrial natriuretic peptide) return
released from the right atrium in

response to its distension by increased
BP Plasma
PV. In the kidney, ANP blocks Na volume Adrenal cortex
reabsorption by inhibiting renin release Aldosterone
from the JGA, angiotensin formation, Cortisol
?
and aldosterone secretion from the
adrenal cortex. ANP also blocks Peripheral vasculature
Selective vasomotion
water reabsorption by antagonizing Starling forces ( HCT)
AVP in the distal tubule and collect- Kidney
ing duct. Overall, it increases GFR UNaV
UV
and urination (figure 4.14). ANP can Renin
also act to lower blood pressure in the
vasculature and the brain. It facilitates
the transcapillary fluid shift between Figure 4.14 Natriuretic actions of ANP in response to
hyperhydration.
E4517/Borer/Fig4.14/401553/alw/pulled-R2
plasma and the ISF compartment and
From TEXTBOOK OF ENDOCRINE PHYSIOLOGY, THIRD EDITION, edited
has vasodilatatory properties. In the by James E. Griffin and Sergio R. Ojeda, copyright © 1996 by Oxford
brain, ANP inhibits the release of AVP University Press, Inc. Used by permission of Oxford University Press, Inc.
92 Advanced Exercise Endocrinology

from the SO and PVN hypothalamic nuclei, attenuates the baroreflex through actions on
NTS, and suppresses thirst and Na hunger by acting on the SFO. Thus, both peripheral and
central actions of ANP reduce Na and water overload.
However, in extreme sports such as marathon and ultramarathon running, small mal-
functions in endocrine fluid regulatory mechanisms may cause major shifts in body fluid
balance. Deliberate excessive intake of water during exercise that lasts more than 4 h is
thought to be the main cause of the development of hyponatremia, or overdilution of plasma.
Hyponatremia is associated with excessive drinking of water during exercise predominantly
in female distance runners, who run relatively slowly and lose modest quantities of sweat
and experience reduction in urine formation and diuresis during exercise. Other variables
that may contribute to hyponatremia are the use of nonsteroidal anti-inflammatory drugs
and running in unusually hot or cold environments. In hyponatremia, plasma osmolarity can
decrease to less than 135 mmol/L (Bürge et al., 2011; Murray & Eichner, 2004). Hypotonic
plasma can cause cerebral and pulmonary edema and neurological symptoms that include
headache, fatigue, and confusion.
Hormonal dysregulation in hyponatremia may involve an inappropriate increase in the
secretion of AVP, leading to water retention and decreased urination and decreased secre-
tion of aldosterone causing reduced sodium retention. According to one view (Noakes et
al., 2005), three independent mechanisms explain why some athletes develop hyponatremia
during and after prolonged exercise: overdrinking due to biological or psychological factors,
the failure to suppress AVP secretion in the face of an increase in TBW, and the failure to
increase plasma Na concentration. Recent research by Bürge and colleagues (2011) impli-
cates both AVP and aldosterone in adjusting of postrace PV after an ultramarathon. In 50
ultramarathon runners who did not develop hyponatremia, postrace serum Na was negatively
correlated with fluid intake during the race. The ultramarathoners drank 0.6 L/h, a midrange
of recommended fluid intakes (0.4-0.8 L/h), which suggests that avoiding overdrinking helps
prevent hyponatremia. Their Na intake during the race was unrelated to postrace serum Na.
No association was found between postrace serum Na and increased postrace aldosterone.
However, the change in aldosterone was significantly and positively associated with postrace
change in transtubular K gradient and the change in K:Na ratio in urine, suggesting that this
hormone has a role in adjusting postexercise PV. AVP appeared to be directly implicated
in the maintenance of normal plasma osmolarity during the ultramarathon because AVP
concentration was significantly and positively correlated with plasma osmolarity.

STRATEGIES FOR FLUID MANAGEMENT IN EXERCISE


An understanding of the imperfect human mechanism of thirst and Na hunger and the con-
straints that thermoregulatory needs place on body fluids is necessary in order to understand
one’s individual responses to exercise and to develop personalized strategies for mitigating
dehydration during exercise. This understanding is needed because exercise-associated heat
load requires diversion of blood to the skin and loss of body fluids through sweating, both
of which contribute to fatigue and impair performance (Gonzalez-Alonso et al., 1999). On
the other hand, excess water that is ingested before exercise cannot be stored. It reduces
plasma osmolarity and leads to diuresis (see figures 4.7 and 4.16). During exercise, excess
ingestion of water can cause hyponatremia due to the inhibition of renal diuresis.
The best overall strategy is to realistically assess an individual’s rate of sweating and the
increase in Tc during exercise. Weight loss is the simple index of body water loss, and Tc
can be monitored with ingestible thermistor probes (Maughan et al., 2007). An assessment
of the magnitude of water (see figure 4.3) and Na (see figure 4.4) losses can be made using
Body Fluid Balance 93

the rate of sweating and Tc elevation. If the goal is to prevent overheating and fatigue during
exercise, drinking NaCl solutions that contain replacement amounts of Na and water is
desirable (Coyle, 2004; Noakes, 2003; Sawka et al., 2007). If the goal is to rapidly rehydrate
following exercise-associated fluid loss, intake of NaCl solutions in concentrations ranging
between 0.14% (23 mM) and 0.36% (61 mM) and in volume 1.5 to 2 times as large as the
losses of body fluid will ensure that body fluid balance is restored in part by correcting
the losses of PV (figure 4.15; Maughan et al., 1996, Shirreffs et al., 1996) and in part by
suppressing the rate of urine output (figure 4.16; Merson et al., 2008).

1,500 1,500

1,000 1,000
Net fluid balance (ml)

Net fluid balance (ml)


500 500

0 0

−500 −500

−1,000 −1,000

−1,500 −1,500

−2,000 −2,000
xerc ise xer cise 0 1 2 4 6
xercis
e
xercis
e 0 1 2 4 6
Pre-e Poste Pre-e Poste
a Time (hr) b Time (hr)

Figure 4.15 Influence of volume and Na content of fluids ingested postexercise on attainment of net body fluid balance.
The Na contentE4517/Borer/Fig4.15a/401554/alw/R2 E4517/Borer/Fig4.15b/401545/alw/R2
was either (a) 23 mM or (b) 61 mM. The volumes ingested were 0.5 (solid circles), 1 (open triangles), 1.5
(solid rectangles), and 2 (open rectangles) times the volume lost.
Reprinted, by permission, from S.M. Shirreffs et al., 1996, “Post-exercise rehydration in man: effects of volume consumed and drink sodium content,”
Medicine and Science in Sports Exercise 28: 1260-1271.

1.80
1 mmol/L
1.60
31 mmol/L
Cumulative urine output (L)

a
1.40 40 mmol/L a
50 mmol/L
1.20
1.00
0.80
0.60
0.40
0.20
0.00
Postdrinking 1 2 3 4
Time point (hr)

Figure 4.16 Cumulative urine output as a function of the concentration of NaCl ingested in a volume 150% that
of body weight lost after exercise. The concentrations ranged from 0.02% to 0.3%.
E4517/Borer/Fig4.16/401637/alw/R3-kh
Springer and European Journal of Applied Physiology, Vol. 103, 2008, pgs. 585-594, “Rehydration with drinks differing in sodium
concentration and recovery from moderate exercise-induced hypohydration in man,” S.J. Merson, R.J. Maughan, and S.M. Shirreffs, figure 2,
with kind permission from Springer Science+Business Media B.V.
94
Body Fluid Balance 95

SUMMARY
This chapter discusses the essential role of body water in survival and especially during
exercise. Exercise challenges body fluids to prevent an excessive increase in body heat load.
This is accomplished through autonomic vasomotor and sudomotor reflexes. The vasomo-
tor reflex redistributes blood to the skin for heat loss by radiation, whereas the sudomotor
reflex helps reduce core body temperature through evaporation of sweat. Because sweating
causes greater losses of body water than of Na, exercise causes both increases in plasma
osmotic pressure causing cellular dehydration, and losses of Na, leading to a reduction in PV.
During rest and the recovery from exercise-induced disturbance in body-fluid balance, a
neuroendocrine reflex that involves increased secretion of ADH (AVP) normalizes plasma
osmotic pressure by increasing renal reabsorption of water. A separate neuroendocrine
renin–angiotensin–aldosterone reflex restores PV by increasing renal reabsorption of Na. In
addition to increasing PV, this reflex increases blood pressure through the vasoconstrictive
effects of angiotensin. Neither reflex operates during exercise because of S redistribution
of blood from the kidneys and other visceral organs to the muscle. For that reason, opera-
tion of the two reflexes as well as thirst and Na hunger are of particular importance during
recovery from exercise. In contrast to ungulates, human capacity to detect imbalances in
body fluid is imperfect and slow. Cellular dehydration and increased osmotic pressure are
corrected first by an early postexercise increase in thirst that is associated with increased
release of AVP. PV normalization occurs several hours later in response to parallel increases
in thirst, Na hunger, and secretion of aldosterone. Because humans slowly and inadequately
sense body fluid disturbances due to exercise, rehydration strategies have been developed
that prescribe volumes and electrolyte composition of rehydrating fluids needed to more
rapidly normalize plasma osmolarity and volume.
96 Advanced Exercise Endocrinology

Hyponatremia is a complication that is encountered when individuals who may have


a minor neuroendocrine insufficiency ingest excess water during exercise that lasts more
than 4 h. The inability of these individuals to prevent a decline in plasma osmotic pressure
appears to be a consequence of inadequate reduction in secretion of AVP and in associated
suppression of water reabsorption.
H APTER

5
C

Hormones and Fuel


Use in Exercise
E
nergy is essential to life, including all forms of movement. Powerful and redundant
physiological mechanisms evolved to sense energy shortage, procure and store
nutrients, and facilitate the release of metabolic fuels from chemically inert stores as
needed. Researchers have extensively studied the role of hormones and the ANS in energy
storage, release, and utilization at rest and during physical activity, although they do not
yet fully understand how energy status is sensed, what controls nutrient intake, and how
energy balance is maintained.
This chapter outlines current knowledge on the role of hormones and the ANS in the
control of energy storage, release, and utilization at rest and during exercise. All hormones
and autonomic messengers have multiple actions, usually including metabolic and anabolic
or catabolic actions because nutrients serve as both sources of energy and structural building
blocks. Conversely, structural proteins in the body also serve as a source of energy. This
chapter focuses on the metabolic roles of insulin and the principal counterregulatory hor-
mones and autonomic transmitters that oppose its action: glucagon, E, NE, ANP, cortisol,
and GH. Because aerobic and resistance modes of exercise engage energy expenditure and
mechanical loading to a different extent, it is reasonable to expect that hormonal responses
to the two types of exercise, as well as their reliance on hormonal versus muscle-specific
actions, will differ. For that reason, this chapter discusses hormonal responses to aerobic
exercise and RE separately. Adaptations to the two modes of exercise and the interactions
of hormones and nutritional support in these adaptations are discussed in chapter 8.

HORMONAL MEDIATION OF ENERGY BALANCE


The antagonism between insulin—the primary function of which is nutrient uptake, energy
storage, and facilitation of carbohydrate metabolism—and counterregulatory hormones—
which mobilize and facilitate the utilization of stored metabolic fuels—is based on the
incompatibility of their signaling pathways (discussed in chapter 1). Through the enzyme
PDE, insulin inactivates cAMP, the second messenger produced by almost all counter-
regulatory hormones (figure 1.14), and thus antagonizes their catabolic action. Insulin also

97
98 Advanced Exercise Endocrinology

triggers intracellular dephosphorylations that inactivate catabolic energy-releasing enzymes


that are initiated when these enzymes are phosphorylated by counterregulatory hormones.
The pleiotropic anabolic and catabolic actions of these hormones are discussed in chapters
6 and 8. However, metabolic and anabolic or catabolic actions of hormones and autonomic
transmitters are related because habitual bioenergetic challenges to some body functions
during exercise produce the structural and functional adaptations discussed in chapter 8.
In addition, many hormone actions during aerobic exercise employ signaling cascades that
begin with hormone-receptor binding and usually engage as the final common pathways
the enzymes activated directly by muscle contractions that were discussed in chapter 2. The
reader is therefore urged to reread chapters 2 and 8 alongside this chapter in order to fully
understand the integrated role of hormones in human exercise physiology. A discussion of
the signaling and metabolic actions of the principal hormones involved in the regulation
of fuel metabolism will allow a better understanding of how their actions accommodate
energy needs during exercise.

Metabolic Actions of Insulin


Insulin is arguably the most important controller of metabolism. Therefore, it is useful to
review its actions and the reasons why suppression of its secretion is adaptive during exercise.
Insulin is produced by beta cells in pancreatic islets of Langerhans, whereas the chief gluco-
regulatory antagonist of insulin, glucagon, is produced by pancreatic alpha cells. Pancreatic
beta cells act as glucose sensors. The presence of glucose at a concentration greater than about
5 mM in extracellular fluid activates the high-Km glucose-phosphorylating enzyme GK and
low-affinity, high-Km glucose transporter GLUT-2 to initiate insulin exocytosis in beta cells
(Evans et al., 2009). Insulin also is released in response to the appearance in circulation of
absorbed amino acids. The metabolic effects of insulin are nutrient uptake; biosynthesis of
the storage molecules glycogen, TG, and protein; stimulation of carbohydrate metabolism;
and inhibition of mobilization of stored nutrients and of protein degradation (figure 5.1).
Recruitment of insulin receptor substrate (IRS) to IR initiates two distinct signaling path-
ways; one activates PKB (or Akt) by way of membrane phospholipids PIP2 and PIP3 and the
other stimulates MAPK and its constituent enzymes ERK1/2 by way of membrane-bound
Ras. Akt activation facilitates glucose uptake through translocation of GLUT-4 glucose
transporters to the membrane surface, inhibits lipolysis through activation of PDE, and
stimulates glycogen synthesis through dephosphorylation of GSK3β. Protein metabolism
is affected through activation of PI3K and MAPK signaling.
Insulin signaling starts with the autophosphorylation of tyrosines on the intracellular
portion of its receptor (IR) after hormone–receptor binding occurs. This allows IR to recruit
adaptor proteins containing SH2 and SH3 domains including IRS, Grb-2, PI3K, and MAPK.
Several IRS proteins are found in different organs. IRS-1 stimulates glycogen synthesis in the
liver, glucose uptake into the muscle, and, together with IRS-3, fat synthesis in adipose tissue.
IRS-2 primarily stimulates hepatic lipogenesis and proliferation of pancreatic beta cells and,
to a lesser extent, appetite suppression and, along with IRS-1, glycogen synthesis in the liver.
Both IRS-1 and IRS-2 inhibit protein degradation and facilitate protein synthesis through
activation of the Akt–mTOR (figure 2.1) and MAPK (figure 5.1) signaling pathways. IRS-4
restrains actions of IRS-1/2 (Thirone et al., 2006; Toniguchi et al., 2006). As described in
the “Insulin Receptor” section in chapter 1, IRS proteins 1 through 4 also undergo tyrosine
phosphorylation. Subsequent to hydrolysis of membrane PIP2 by phosphatidylinositol-3-kinase
(PI3K) to PIP3 (figure 5.1), additional effector proteins are recruited to the cell membrane,
one of which is protein kinase B (PKB or Akt).
Hormones and Fuel Use in Exercise 99

Insulin

Insulin receptor

GLUT-4 PIP 2
PY PY PY
PIP 3

IRS

IRS
Glucose Grb-2 SOS
P13K
Ras
Akt/PKB GTP

GDP
AS160 GSK3 PDE 3β mTOR Raf
Raptor
p70S6K
GS PKA MEK1/2
Protein
GLUT-4 synthesis
vesicle Glycogen HSL
synthesis MAPK/
ERK1/2
Lipolysis Cellular
Glucose proliferation
uptake

Nucleus

Akt ERK

DNA FOXO3A FOXO4 FOXO1 AP-1

Apoptosis Growth

Figure 5.1 Insulin signaling pathways. Insulin signaling may proceed down several different enzymatic cascades.
E4517/Borer/fig.5.1/401555/TB/R4-KH
Phosphotyrosines on IR recruit and phosphorylate adaptor proteins (insulin receptor substrate [IRS], Grb-2, PI3K,
and MAPK) that have an SH2 or SH3 domain.

When Akt is phosphorylated, it phosphorylates and activates other effectors that then
mediate insulin actions. After its activation, Akt mediates insulin-dependent glucose uptake,
an action that is not necessary for muscle glucose uptake during exercise; promotes glycogen
synthesis that is facilitated after glycogen-depleting exercise followed by high carbohydrate
intake through dephosphorylation of GSK3β; increases carbohydrate oxidation by stimulating
pyruvate dehydrogenase (PDH) kinase expression and PDH activity; and inhibits lipolysis,
glycogenolysis and gluconeogenesis, and fatty acid muscle uptake and oxidation through a
TF FOXO that acts on nuclear DNA.
The second insulin signaling pathway entails activation of MAPK. Effectors upstream
of MAPK include the Grb-2–SOS complex, which binds to IRS, and the membrane-bound
Ras G protein, which acts as a signaling switch when phosphorylated. Ras then activates
Raf and the kinase MEK. MEK phosphorylates and activates MAPK, which consists of
ERK1 and ERK 2. The same pathway may be activated by GH (figure 1.13). IRS-1 and
IRS-2 inhibit protein degradation and facilitate protein synthesis through the Akt–MTOR
(figures 2.1, 2.6, and 2.7) and MAPK (figure 5.1) signaling pathways.
100 Advanced Exercise Endocrinology

Insulin-Dependent Glucose Uptake


Of all insulin effects, the one that is best documented is the facilitation of cellular glucose
uptake. Initial tyrosine phosphorylation of IRS protein leads to activation of PI3K (figure
5.1, left). The PI3K heterodimer comprises a p85 regulatory subunit and a p110 catalytic
subunit (figure 2.1), both of which get phosphorylated and further activate the PIP3 and
PDK-1 (3'-phosphoinositide-dependent kinase-1) by phosphorylation (figure 2.1). PDK-1 in
turn activates PKB (Akt), which acts on the Akt substrates AS160 and TBC1D1 to facilitate
translocation of the glucose transporter GLUT-4 to the cell membrane (Cartee & Funai,
2009). Both substrates have sites at which insulin-stimulated enzyme Akt acts, but TBC1D1
also has an AMPK activation site (figure 5.2) and is implicated in non-insulin-dependent
GLUT-4 translocation during muscle contraction (Cartee & Funai, 2009; Funai & Cartee,
2009).
The liver, brain, and pancreas import glucose by non-insulin-dependent mechanisms.
(Chapter 2 discusses non-insulin-dependent glucose uptake by the muscles.) Glucose trans-
port into the liver and pancreatic islets of Langerhans is mediated by GLUT-2 (Clark &
Newgard, 2007), and glucose transport into the brain neurons is mediated by GLUT-3 via
the blood-brain barrier (BBB) that requires GLUT-1 (Simpson et al., 2008). The affinity
of GLUT-2 for glucose in the liver and pancreas is 5 to 10 mM, appropriate for response
to postprandial plasma glucose increases from 4 to 5 mM to 9 to 10 mM. The affinity of
GLUT-3 is higher than that of GLUT-2, and its transport capacity is five times greater than
that of GLUT-4. This gives brain neurons preferential access to glucose, given that the local
plasma glucose concentration is low at 1 to 2 mM (Simpson et al., 2008).

Stimulation by Insulin of Carbohydrate Metabolism


The stimulation by insulin of carbohydrate metabolism (figure 5.3) includes direct and
indirect actions. Direct actions include the stimulation of four enzymes that unidirectionally
regulate glycolysis and the consignment of pyruvate to the oxidative Krebs cycle (Clark &
Newgard, 2007). First, insulin increases the affinity of hepatic GK for glucose beyond its

AMPK site Akt site

S237 T596

Human TBC1D1 PTB PTB CBD GAP

1 96-153 301-373 721-757 800-994 1168

Akt sites

S588 S666
S318 S341 S570 T642 S751

Human AS160
(TBC1D4) PTB PTB CBD GAP

1 121-180 367-439 832-876 918-1121 1299

Figure 5.2 The two Akt substrates involved in insulin-stimulated translocation of GLUT-4 to the cell membrane.
Borer/E4517/Fig.5.2/401557/TimB/r2-alw
Both TBC1D1 and AS160 have sites at which Akt acts; TBC1D1 also has an AMPK binding site.
Reprinted, by permission, from G.D. Cartee and K. Funai, 2009, “Exercise and insulin: Convergence or divergence at AS160 and TBC1D1?”
Exercise in Sport and Sciences Reviews 37: 188-195.
Hormones and Fuel Use in Exercise 101

Glucose Insulin +
Glycogen synthesis
Insulin
+ +
Glycogen
Glucose Pentose shunt
Glucose
G6P NADP +
uptake NADPH
F6P Gluconolactone 6P
Insulin NADP
F1,6P2
NADPH
TG
TrioseP α glycerophosphate
Glycolysis + +
NAD+ Fatty acids Lipogenesis
NADH NAD+ ADP
Insulin
NADPH NADPH ATP
PEP
Insulin NADP+ Malonyl CoA +
Pyruvate Malate
AcCoA
NAD
Pyruvate ATP ADP NADH
+
Insulin OAA OAA
AcCoA
Citrate
Carbohydrate Citrate
+
oxidation
Mitochondrion
Krebs cycle

Figure 5.3 Stimulation by insulin of carbohydrate and fat metabolism. Insulin stimulates Embden-Meyerhof
glycolysis (left), pentose monophosphate shunt (PMS, top right), carbohydrate oxidation in the mitochondrion
(bottom), and free fatty acid (FFA) E4517/Borer/fig.5.3/401558/TB/R2-alw
and TG synthesis (lower right).
Adapted, by permission, from G.A. Leveille and D.R. Romsos, 1974, “Meal eating and obesity,” Nutrition Today 9(6): 4-9.

Km (6 mM). Then, GK (and the corresponding HK enzyme in the muscle) traps glucose
in hepatocytes by phosphorylating it to G6P. Insulin rapidly increases GK availability in
the cytoplasm by triggering its dissociation from the GK regulatory protein that otherwise
sequesters this enzyme in the nucleus. Insulin also stimulates transcription of GK and HK
genes.
It is important to appreciate the difference in glucose handling by GK in the liver and by
HK in the muscle. The Km of muscle HK is 0.1 mM and the Km of hepatic GK is 10 mM.
Therefore, the affinity of HK for glucose is about 100 times greater than the affinity of GK
for glucose. This gives muscle the capacity to take up, phosphorylate, and sequester within
its cells glucose at hypoglycemic concentrations while GK is activated by postprandial
increases in plasma glucose above fasting concentrations. HK is near maximally inhibited
by its product G6P, whereas GK is not under this negative feedback control. This gives
muscle the capacity to import glucose as needed for metabolism or glycogen synthesis while
glucose uptake by the liver in response to high glucose concentrations is not so limited.
The second direct action of insulin on carbohydrate metabolism is activation of PFK1
that converts F6P to F1,6P2. PFK1 is regulated by three factors: cellular energy charge
(ATP:AMP ratio), plasma concentration of citrate generated by TCA cycle, and fructose
2,6-bisphosphate (F2,6P2). After feeding, concentrations of plasma glucose and insulin
102 Advanced Exercise Endocrinology

increase. Increased glucose uptake and activity of GK facilitate glycolysis down the PMS
(figure 5.4, right) as well as down the Embden-Meyerhof glycolytic pathway (figure 5.4,
left). Xylulose 5 phosphate, a substrate in the PMS, activates a protein phosphatase that
through dephosphorylation increases kinase activity of a bifunctional enzyme. This stimu-
lates glycolysis by increasing both the concentration of F2,6P2 and the affinity of PFK for
F6P by lowering the ATP negative feedback over PFK activity (Clark & Newgard, 2007;
figure 5.4, bottom). Concurrently, fructose 1,6-bisphosphate (F1,6P2) which converts F1,6P2
back to F6P, is deactivated. Insulin also facilitates transcription of PFK and of bifunctional
enzyme genes.
The third enzyme through which insulin directly regulates glycolysis is PK. It catalyzes
conversion of phosphoenolpyruvate (PEP) to pyruvate (figure 5.3). Like PFK, it is respon-
sive to the ATP:AMP energy charge. Its activity is facilitated by F1,6P 2 and by protein
phosphatase that increases the concentration of F2,6P 2. The same protein phosphatase
also activates the ChREBP (carbohydrate-responsive element-binding protein), a TF that
increases the expression of the hepatic PK gene and promotes fat synthesis (Iizuka &
Horikawa, 2008).

Glucose Insulin

Insulin + GLUT-2
Glucose
Glucokinase

G6P Pentose shunt Xyl5P

Bifunctional
enzyme (+)
Xyl5P-regulated
Pi phosphatase
Glycolysis F1,6- F6P
PFK (+) P
stimulated BPase F1,6P2 F6P
F6P
F2,6P2 F2,6P2

F2,6P2
NADP+
Pyruvate
Malate

Pyruvate ATP ADP


+
Insulin OAA OAA
AcCoA
Citrate
Citrate
Carbohydrate
oxidation Mitochondrion
Fed state

Figure 5.4 Control of glycolysis by insulin. The E4517/Borer/fig.5.4/401559/TB/R2-alw


postprandial increase in glucose and insulin stimulates the
PMS and activates a protein phosphatase. This enzyme dephosphorylates a bifunctional enzyme and increases
its kinase activity. The result is an increase in the synthesis of F2,6P2, a regulator of PFK activity, and an increase
in glycolysis through activation of PFK.
Springer and Mechanisms of Insulin Action, 2007, pg. 96, Hepatic regulation of fuel metabolism, C. Clark and C.B. Newgard, edited by A.R.
Saltiel and J.E. Pessin, figure 2, with kind permission from Springer Science+Business Media B.V.
Hormones and Fuel Use in Exercise 103

The final enzyme through which insulin directly regulates carbohydrate metabolism
is PDH. It catalyzes oxidation and decarboxylation of the glycolytic end-product pyruvate
to acetyl CoA, a metabolic step that commits carbohydrate irreversibly to oxidation in the
Krebs TCA cycle. Pyruvate can produce a limited amount of energy anaerobically when
converted to lactate by lactate dehydrogenase, but it releases a full complement of the ATP
contained in glucose molecular bonds when it is oxidized in the TCA cycle. Both glucose
uptake and glucose oxidation are useful ways to reduce hyperglycemia produced by high
consumption of carbohydrate. Pyruvate also can be rerouted toward glucose by gluconeo-
genesis, as discussed in the “Gluconeogenesis” section.

Insulin and Biosynthesis of Fuel Stores


In the fed condition, insulin stimulates conversion of glucose to glycogen and TG in the liver,
muscle, and adipose tissue. This is partly driven by the facilitation by insulin of glucose
uptake and stimulation of GK. G6P is a substrate for glycogen synthase, which builds α-1,4
glycosidic bonds between glucose molecules to promote the growth of a glycogen strand.
A branching enzyme adds side carbohydrate chains by way of α-1,6 glycosidic bonds. At
the molecular level, insulin promotes glycogen synthesis through PI3K and Akt. The latter
phosphorylates GSK3β and reduces its kinase activity, which is inhibitory to glycogen
synthesis (see figures 2.1, 2.7, and 2.8). Glycogen synthesis is also facilitated by protein
phosphatase 1 that joins with scaffolding proteins to produce glycogen targeting subunits
in both liver and muscle.
Insulin also facilitates fat synthesis and TG storage in muscle and the adipose tissue by
both direct and indirect actions (figure 5.3). As is the case with glycogen synthesis, insulin
promotes fatty acid synthesis through increased glucose uptake, support of carbohydrate
metabolism, and release of enzymes involved in fatty acid and triacylglycerol synthesis.
Insulin-stimulated carbohydrate metabolism indirectly provides three substrates necessary
for fatty acid and TG synthesis. First, the Embden-Meyerhof glycolytic pathway (see figure
5.3, left) produces 3PG (3-phosphoglycerate), which is necessary for the production of the
activated form of glycerol for TG synthesis. Second, insulin produces NADPH reducing
equivalents, which are necessary for supplying energy for FFA synthesis in the PMS (figure
5.3, top right). Oxidation of pyruvate by PDH in the mitochondrion produces acetyl CoA,
the building block for the growing FFA chain (figure 5.3, bottom right). Third, insulin
stimulates transcription and activity of ACC (acetyl CoA carboxylase), the enzyme that
initiates FFA synthesis (Zang et al., 2005). Its metabolic product is malonyl CoA, which is
a potent inhibitor of CPT I (carnitine palmitoyltransferase I), the gate-keeping enzyme for
transport of long-chain fatty acids into the mitochondrion for FFA oxidation. The assem-
bly of fatty acids and activated α-GP into TG is catalyzed by fatty acid synthetase (FAS).
Insulin mediates transcription and activation of this enzyme as well (Radenne et al., 2008).
The transcriptional effect of insulin on genes involved in lipogenesis (including genes for
GK, PK, SCD-1 [stearoyl-CoA desaturase], ACC, and FAS) is stimulated by SREBP-1c
(sterol regulatory element-binding protein-1c) and antagonized by glucagon. The activity of
SREBP-1c TF accounts for about one half of this gene translational effect. The other half
is contributed by the TF ChREBP.
Although insulin does facilitate de novo TG synthesis in adipose tissue and the liver
as described previously, the activity of endothelial enzyme LPL (lipoprotein lipase) is the
principal avenue of fat accumulation and storage in the adipose tissue of humans. After
a human consumes a meal containing fat, adipose tissue LPL is activated by insulin and
hydrolyzes FFA from TG packaged in lipoproteins produced by the intestine (chylomicrons)
and the liver (VLDL).
104 Advanced Exercise Endocrinology

Suppression by Insulin of Fuel Mobilization and Lipid Oxidation


Insulin blocks mobilization of FFA from TG and the breakdown of glycogen to glucose.
Lipolysis is suppressed at insulin concentrations as low as 13 µU or 96 pM. Therefore, any
decline in insulin concentration below this threshold, as occurs during exercise and during
the postabsorptive intermeal interval, is a powerful lipolytic stimulus (Groop et al., 1992).
The antilipolytic actions of insulin are mediated through its degradation of cAMP by way
of activated PDE 3B (figure 5.1; Eriksson et al., 1995). PDE activation is mediated by cGMP
after hormonal stimulation of Gq regulatory protein (figure 1.14), and PDE 3B is activated
by phosphorylation by Akt (figure 5.1). PDE inactivates cAMP by converting it to its inac-
tive 5' AMP form and thus reduces lipolysis. Insulin suppresses hepatic glucose production
at the somewhat higher concentration of 24 μU/ml (Nurjhan et al., 1986) by blocking the
expression of the gluconeogenic enzyme PEPCK through the PDE mechanism and through
inhibition of glucose-6-phosphatase (G6Pase), the enzyme that is necessary for converting
G6P to a more diffusible free glucose (Sasaki et al., 1984). Insulin inhibits lipid oxidation
at still-higher concentrations of 44 μU or 324 pM (Campbell et al., 1992), again through the
degradation of cAMP by PDE and in part by inhibiting the activity of CPT I via malonyl
CoA (Zang et al., 2005).

Mobilization of Metabolic Fuels During Exercise


A shift from nutrient uptake and storage toward mobilization of stored fuels and utilization
of carbohydrate and FFAs entails inhibiting insulin secretion and actions by activating the
appropriately named (insulin) counterregulatory hormones glucagon, the catecholamines,
and, to a lesser extent, GH and cortisol. Catecholamines inhibit insulin secretion by stimu-
lating α2 AdRs on pancreatic beta cells. Collectively, counterregulatory hormones protect
against the development of hypoglycemia by opposing insulin hypoglycemic action and
increasing hepatic glucose output and blood glucose concentration, initially through break-
down of glycogen and later by gluconeogenesis. Concurrently, counterregulatory hormones
reduce insulin antilipolytic actions and facilitate lipolysis and lipid oxidation.

Glycogenolysis
When plasma glucose concentrations decrease below the regulated range, secretion of
glucagon, cortisol, E, and GH is triggered. When plasma glucose concentrations increase
above the regulated range, secretion of glucagon, cortisol, E, and GH is suppressed. E and
glucagon increase hepatic glycogenolysis as well as gluconeogenesis, and E also stimulates
glycogenolysis in the muscle, a tissue devoid of gluconeogenic enzymes. Glucagon plays a
more important role in glucoregulation than does E (Lavoie et al., 1997), and E responds to
changes in glucose concentration during exercise more than does NE (Galbo et al., 1977b).
GH and cortisol stimulate gluconeogenesis in part by upregulating gluconeogenic genes, and
cortisol increases gluconeogenesis primarily from amino acids as a substrate (Gerich, 1988).
Glycogenolysis is initiated by activation of adenylyl cyclase and release of cAMP via the
Gs type of G regulatory protein after glucagon binds to its receptors in the liver or E or its
receptors in the liver and muscle (figure 1.13). Glycogenolysis is also initiated by Ca signal-
ing that is triggered by either glucagon or muscle contractions (figure 2.2). Activation of
either signaling pathway produces sequential phosphorylations of phosphorylase kinase and
hepatic and muscle phosphorylase to convert the latter into its active form. This allows the
breakdown of glycogen and release of glucose 1-phosphate, which then enters glycolysis by
conversion to G6P. An increase in G6P concentration through glycogenolytic or gluconeo-
Hormones and Fuel Use in Exercise 105

genic actions of glucagon or catecholamines activates a low-Km (2-3 mM) enzyme glucose
6 phosphatase that dephosphorylates G6P and permits the release of glucose by diffusion
from the liver into plasma (Van Schaftingen & Gerin, 2002). Concurrent phosphorylation of
glycogen synthase A and its conversion to the inactive B form prevents glycogen synthesis.
The result of these hormonal events is a several-fold increase in hepatic glucose production
during exercise and a moderate production during an overnight fast (about 2 mg·kg of body
weight-1·min-1, or about 5 to 6 g/h; 70%-80% derived from glycogenolysis and about 20%
derived from gluconeogenesis).

Gluconeogenesis
Gluconeogenesis is synthesis of glucose from pyruvate, lactate, glycerol, and amino acids.
Gluconeogenesis is activated in the liver after its glycogen is depleted and in the kidney
during a prolonged fast when high titers of counterregulatory hormones and low insulin
concentration stimulate lipolysis and a rise in FFA availability (figure 5.5). This hormonal
environment blocks production of malonyl CoA, which is an allosteric inhibitor of CPT I.
Disinhibition of CPT I and an increase in FFA availability stimulate FFA oxidation and
the production of acetyl CoA, NADH, and ATP. The latter three substrates, along with low
insulin concentration, inhibit PDH activity and shift the metabolism from carbohydrate to
fat oxidation (Randle et al., 1965). Acetyl CoA is a potent stimulus for the gluconeogenic
enzyme pyruvate carboxylase (PC). ATP produced by fat oxidation can be used in energy-
consuming gluconeogenic PC and PEPCK (phosphoenolpyruvate carboxykinase) reactions.
Glucagon stimulates gluconeogenesis by opposing insulin action through its suppres-
sion of the concentration of F2,6P2 (see figure 5.5) while insulin inhibits gluconeogenesis
by stimulating production of F2,6P2 (see figure 5.4) and by blocking the expression of the
gluconeogenic enzyme PEPCK (Mounier & Posner, 2006). After stimulating the production
of cAMP-dependent PKA, glucagon causes phosphorylation of the bifunctional enzyme and
increases its phosphatase action, which then catalyzes degradation of F2,6P2 to F1,6P. This
suppresses the activity of PFK (see figure 5.5) and initiates the reverse gluconeogenic flow
of glycolytic intermediates to glucose. As nutritional or hormonal status changes in favor
of either carbohydrate or lipid utilization, the consignment of pyruvate to the TCA cycle or
gluconeogenesis shifts accordingly. The final step in hepatic glucose output resulting from
gluconeogenesis is the dephosphorylation of G6P by glucose 6 phosphatase and the release
of glucose into circulation.
Glucocorticoids increase hepatic glucose production through gluconeogenesis from
amino acids. They do so by suppressing cellular amino acid uptake for protein synthesis by
extrahepatic tissues and increasing amino acid availability in circulation for uptake by liver
for gluconeogenesis. During prolonged moderate-intensity exercise and after depletion of
glycogen stores, low plasma insulin and increased plasma cortisol concentration facilitate
protein breakdown. By binding to the GRs, which then act as a gene TFs, glucocorticoids
increase the transcription of the gluconeogenic enzyme PEPCK. Glucocorticoids also facili-
tate stimulation of gluconeogenesis in the liver by glucagon and E. This increases binding
of CREBP to the CRE and initiates the transcription of PEPCK.
The liver and muscle collaborate during exercise to provide two gluconeogenic substrates
through operation of the Cori and glucose–alanine cycles (figure 5.6). The Cori cycle pro-
vides the liver with lactate and pyruvate substrates for the gluconeogenic pathway. Alanine
is produced by the contracting muscle from pyruvate through acquisition of an amino group
from glutamine. Release by the muscle of alanine is proportional to exercise intensity and
thus is a result of transamination from branched-chain and other amino acids and is unrelated
106 Advanced Exercise Endocrinology

Glucose
Glucagon
GLUT-2
Glucose
Glucose-6-phosphatase

G6P Pentose cAMP


shunt
ATP (+)
PKA
P
F1,6- F6P F6P F6P
PFK F2,6P2
BPase F1,6P2 F2,6P2
Glyconeogenesis
Bifunctional
enzyme F2,6P2
PEP NADPH
OAA Malonyl CoA
NADP+
Pyruvate Malate –
AcCoA
NAD
Pyruvate ATP ADP NADH
OAA OAA
AcCoA
Liquid Citrate
oxidation Citrate

Mitochondrion
Krebs cycle
Fasted state

Figure 5.5 Control of hepatic gluconeogenesis by glucagon. A decrease in blood glucose and concurrent increase
E4517/Borer/fig.5.5/428888/TB/R2-alw
in fat oxidation stimulates pancreatic glucagon secretion, which then activates cAMP. This leads to the phosphory-
lation of bifunctional enzyme and an increase in its phosphatase action and degradation of F2,6P2 to F1,6P. This
suppresses the activity of PFK and initiates the reverse gluconeogenic flow of glycolytic intermediates to glucose.
Springer and Mechanisms of Insulin Action, 2007, pg. 96, Hepatic regulation of fuel metabolism, C. Clark and C.B. Newgard, edited by A.R.
Saltiel and J.E. Pessin, figure 2, with kind permission from Springer Science+Business Media B.V.

to muscle amino acid composition (Ahlborg et al., 1974). In the liver, transamination of the
amino group from alanine to α-ketoglutarate produces pyruvate and glutamate, thus again
providing a gluconeogenic substrate. Glucagon facilitates the Cori cycle by stimulating the
gluconeogenic enzymes PEPCK and F1,6-biphosphatase and facilitates the glucose–alanine
cycle by increasing hepatic amino acid uptake (Felig, 1973; Kietzman et al., 1998). High
titers of catecholamines and other counterregulatory hormones also increase the supply of
glycerol and lactate for gluconeogenesis.

Lipolysis
By default, the limitation in the size of carbohydrate stores leads to metabolism of lipids
during times of energy shortage. Liver contains between 70 and 90 g of glycogen and muscle
contains between 16 and 45 g/kg, depending on the levels of training and muscle glycogen
supercompensation. Physiological limitations to carbohydrate use include the capacity of the
liver to produce only about 75 g of glucose by glycogenolysis and gluconeogenesis during 2
to 4 h of exercise of different intensities (Ahlborg & Felig, 1976, 1977) and the inability of
muscle to release glucose into circulation due to the absence of glucose 6 phosphatase. By
Hormones and Fuel Use in Exercise 107

FFAs TG
Adipose
Glycogen tissue

Glycerol
– +
(P) (GS) Insulin Glycogen
GL
Glucose
+ –
GIP + GL
+
)
(HK +
(GK) GIP
G6P –
(G6P) Glucagon
Cortisol G6P

+
BCAA P Alanine + F6P
+ +
(LDH) (PFK) FP2
(PDH) GNG
Lactate FI,6P2
Gl

– + OAA
uc

KCAC se (PEPCK)
o

-a
Co la 3PG
ri c nin
yc ec +
le yc (PC)
Muscle le PEP
(PK)
Liver
P

Figure 5.6 Hormonal control of carbohydrate metabolism in the muscle and liver. Glucagon stimulates, and insulin
inhibits, hepatic glycogenolysis and gluconeogenesis. Glucagon also facilitates hepatic amino acid uptake. During
exercise, the muscle provides gluconeogenic substrates to the liver through the operation of the glucose–alanine
E4517/Borer/fig.5.6/401560/TB/R3-alw
and Cori cycles. Glycerol that results from adipose tissue lipolysis also is a gluconeogenic precursor. Cortisol
provides additional gluconeogenic amino acid substrates by facilitating protein degradation.
Reprinted, by permission, from K. Borer, 2003, Exercise endocrinology, (Champaign, IL: Human Kinetics), 99.

contrast, adipose tissue can store in excess of 200,000 kcal of energy, and the vastus lateralis
muscle contains about 1.45 g/kg (5.4 mM/kg) of intramyocellular triglycerides (IMTGs;
Pan et al., 1997). Therefore, lipolysis—the enzymatic breakdown of TG into three FFAs
and glycerol in the adipose tissue and their release into circulation and the breakdown of
IMTGs to FFAs and glycerol in the muscle—is an important source of ATP during aerobic
exercise and other conditions of negative energy balance.
Lipolysis in humans is under negative control primarily by insulin and activation of
alpha AdRs and under positive control by the catecholamine action on beta AdRs, and by
ANP (Birkenfeld et al., 2005; Sengenes et al., 2000) and GH (Møller & Jorgenson, 2009).
Insulin-associated lipolysis is centrally regulated by different sets of autonomic neurons
in the brain stem that inhibit insulin secretion and is initiated when insulin concentration
decreases below 13 µU or 96 pM. E stimulates lipolysis at plasma concentrations as low as
three times E’s basal level (75-125 versus 25 pg/ml; Galster et al., 1981). Researchers have
found clear evidence in animals (Bartness & Song, 2007) and more tenuous evidence in
humans (Dodt et al., 2003) of direct S innervation of the subcutaneous or intra-abdominal
fat compartment and activation of lipolysis. Others maintain that E, the release of which
108 Advanced Exercise Endocrinology

is stimulated by preganglionic S fibers, is the principal lipolytic catecholamine (Lafontan


& Langin, 2009). In humans, the catecholamines initiate lipolysis through two actions:
inhibiting insulin secretion and binding to β1 and β2 AdRs throughout adipose tissue and
to β3 receptors in visceral fat (Krief et al., 1993). The latter type of beta AdRs is reported
to be activated by high concentrations of catecholamines after β1 and β2 receptors become
desensitized (Granneman, 1995). All three receptors activate Gs regulatory protein and
adenylyl cyclase, which in turn stimulates PKA to phosphorylate the hormone-sensitive
lipase (HSL) and perilipins (proteins coating the surface of the lipid droplets; figure 5.7).
Catecholamines also can inhibit lipolysis by stimulating α2 AdR, which act via the Gi regula-
tory protein to inhibit the formation of cAMP (figure 1.15). In addition, ANP activates HSL
through production of cGMP as a second messenger and activation of PKG (figure 5.7).
The degree of lipid mobilization of regional adipose tissue depends on the adipose tissue
distribution of alpha and beta AdRs. The former are more numerous in subcutaneous adipose
tissue, and the latter are more numerous in the visceral fat depot. Of the three lipases that
participate in TG breakdown to FFA and glycerol—the adipose triglyceride lipase (ATGL),
HSL, and monoacylglycerol lipase (MGL)—HSL is considered to be the rate-limiting step
in lipolysis and is regulated by hormones. Perilipin phosphorylation alters the lipid droplet
surface to allow HSL access to the TG and activates the ATGL to break down TG to DAG.
HSL has high affinity for DAG and hydrolyses it to monoacylglycerol (MAG). MGL then
hydrolyses MAG to glycerol and FFAs (figure 5.7). A moderate fast triggers lipolysis by
lowering insulin concentration and action and by enhancing adipose tissue responsiveness to
catecholamines. A more severe and prolonged calorie restriction stimulates greater lipolysis
as a result of low insulin concentration and increased release and action of catecholamines,
ANP, and GH. In addition, increased expression of HSL and ATGL adds to facilitation of
lipolysis under more extreme energy depletion.
GH is an additional lipolytic hormone. The lipolytic action of GH may to a large extent
account for its suppression of protein oxidation during exercise (discussed in chapter 8). The
mechanism of GH-induced lipolysis is less understood. Because between 70% and 80% of

ANP E, NE Insulin

NPR-A βAR IR

cGMP cAMP PDE PKB

PKG PKA

FFAs
ATGL HSL HSL MGL glycerol
Perilipin
Triglycerides
(lipid droplet)

Figure 5.7 Hormonal stimulation of HSL. Catecholamines stimulate HSL by acting on β1 and β2 AdRs and by
E4517/Borer/fig.5.7/401561/TB/R2-alw
stimulating the Gs regulatory protein, cAMP, and PKA. ANP does so by activating cGMP and PKG. Lipolysis is
inhibited by insulin acting via PDE 3B and by α2 AdR, which act via the Gi protein to inhibit the formation of cAMP.
Adapted from Progress in Lipid Research, Vol. 48, M. Lafontan and D. Langin, “Lipolysis and lipid mobilization in human adipose tissue,” pg.
278, copyright 2009, with permission from Elsevier.
Hormones and Fuel Use in Exercise 109

GH secretion occurs at night, GH contrib- 14


utes to increased FFA utilization during

Growth hormone (µg/L)


12
an overnight fast. GH pulsatile pattern 10
(figure 5.8) is necessary for stimulation
8
of lipolysis during a prolonged fast (Surya
6
et al., 2009). Lipolysis by GH is activated
after a delay of about 2 h (Gravholt et al., 4
1999; Møller & Jorgenson, 2009), indicat- 2
ing that GH facilitates lipolysis through 0
the synthesis of necessary enzymes and of 7 a.m. 11 a.m. 3 p.m. 7 p.m. 11 p.m. 3 a.m. 7 a.m.
Time of day
beta AdRs, which increase the sensitivity
of adipose tissue to catecholamines. Figure 5.8 The daily pulsatile GH pattern in lean adults
GH lipolytic action is not mediated that is necessary for the stimulation of lipolysis.
by HSL. One year of GH therapy in GH- E4517/Borer/Fig5.9/401562/alw/R1
Reprinted, by permission, from S. Surya et al., 2009, “The pattern of
deficient individuals did not change the growth hormone delivery to peripheral tissues determines insulin-like
expression of HSL mRNA in adipose growth factor-1 and lipolytic responses in obese subjects,” Journal of
Clinical Endocrinology and Metabolism 94: 2828-2834.
tissue (Khalfallah et al., 2001). Nor does
GH lipolysis depend on the activities of
PKA, PKC, PLA, PLC, or MAPK (Asada et al., 2000). On the other hand, a whole-genome
microarray analysis of gene expression profiles in the adipose tissue of men with hypopi-
tuitarism who were treated for 1 mo with 0.5 mg/d of recombinant GH revealed that GH
affects lipolysis through the upregulation of a novel TG hydrolase containing a patatin-like
phospholipase domain (PNPLA3), the suppression of a novel lipid droplet-associated pro-
tein (CIDEA) that promotes lipid storage, and the downregulation of a 11 β hydroxysteroid
dehydrogenase type 1 (11β-HSD1) that facilitates conversion of cortisone to cortisol (Zhao
et al., 2011). These genomic actions were correlated with increased whole-body lipolysis
and fat oxidation. Downregulation of 11β-HSD1 reduces the expression of proinflammatory
components of the extracellular matrix (e.g., integrins and adhesion proteins) and the TGFβ
signaling pathway. GH administration also increased expression of IGF-I in the adipose
tissue, leading to increased adipocyte proliferation but to suppression of adipocyte differ-
entiation. GH also upregulated enzymes that promote synthesis of DAG, acetyl-coenzyme
synthetase (ACSS), and phosphatidic acid phosphatase (LPIN1). Thus, GH alters adipocyte
lipid turnover toward destabilization of TG deposition into lipid droplets. This makes lipid
droplets available for hydrolysis by the novel TG hydrolase and promotes formation of DAG
and hydrolysis of TG (figure 5.9).

Counterregulatory Hormones and Insulin Resistance


The catecholamines, together with cortisol and GH, stimulate hepatic glucose production but
suppress insulin-dependent glucose uptake. This is a consequence of their interference with
activation of insulin signaling (Smith et al., 1997; see chapter 1) and facilitation of lipolysis
and fat metabolism (Krag et al., 2007). A concurrent increase in plasma glucose concentration
and insulin resistance occurs when energy is needed during exercise and other conditions
of energy need. In such emergency situations the key organs—brain and muscle—are able
to take up the limited supply of glucose by insulin-independent mechanisms.

HORMONES IN FUEL MOBILIZATION AND UTILIZATION


DURING AEROBIC EXERCISE
The metabolic switch to a catabolic process of fuel mobilization and utilization during
exercise is mediated by the suppression of insulin secretion and the increased release of
110 Advanced Exercise Endocrinology

FFA FFA
+

ACSS
GH

(c)
DAG FACoA
+

LPIN1
GH

PNPLA3
+
GH
(b)
PA
TG

− DAG
GH CIDEA Perilipin
(a) Lipid
droplet

Adipocyte

Figure 5.9 Lipolytic action of GH. Downregulation of a lipid droplet-associated protein (CIDEA) destabilizes TG
E4517/Borer/fig.5.9/401563/TB/R2-alw
deposition into lipid droplets in adipocytes. Upregulation of the novel TG hydrolase PNPLA3 facilitates lipolysis.
Concurrent upregulation of acetyl-coenzyme synthetase (ACSS) and LIPIN promotes biosynthesis of DAG and
facilitates TG turnover.
Reprinted, by permission, from J.T. Zhao et al., 2011, “Identification of novel GH-regulated pathway of lipid metabolism in adipose tissue: A
gene expression study in hypopituitary men,” Journal of Clinical Endocrinology and Metabolism 96: E1188-E1196.

catecholamines, glucagon, cortisol, and GH. The hypoglycemic and antilipolytic actions of
insulin during exercise are prevented by suppression of its secretion, which is achieved by
the binding of circulating catecholamines or NE delivered from the S nerve terminals to
alpha AdRs on pancreatic islet beta cells (Galbo et al., 1977a; Porte et al., 1966). At lower
and intermediate exercise intensities, secretion of glucagon, cortisol, E, and GH are pre-
dominantly governed by metabolic need and are subject to homeostatic feedback control.
Decreases in plasma glucose (or FFA availability, in the case of GH) stimulate secretion of
these hormones, whereas hyperglycemia or high FFA concentrations suppress the secretion
of these hormones (Casanueva et al., 1987; Galbo et al., 1977b; Sotsky et al., 1989). How-
ever, secretion of counterregulatory hormones increases in proportion to exercise intensity
and associated energy expenditure, as illustrated by GH (figure 5.10; Pritzlaff et al., 1999)
and catecholamine (see figure 3.6) release. The threshold exercise intensity necessary to
increase GH pulse amplitude appears to be 30% of maximal effort and may increase 100-
fold, whereas the latency is about 10 min after the onset of exercise. Prolonged daytime
exercise that lasts longer than 4 h may increase nocturnal GH secretion in the second half
of sleep (Jenkins, 1999). .
At low exercise intensity of between 25% and 50% of VO2max, S tone increases
and E. concentration doubles. During moderate-intensity exercise of between 50% and
75% VO2max, plasma E and NE.concentrations increase between four- and sixfold. At high
exercise intensities above 75% VO2max, plasma E and NE concentrations can reach levels
that are 17 to 20 times higher than they are at rest (figure 3.6). Thus, as exercise intensifies,
activation of S nerves and secretion of counterregulatory hormones become exponential and
assume the feed-forward characteristics of an emergency response. This is particularly the
case with secretion of adrenal NE or its spillover from S nerve endings because this mes-
senger is not responsive to declines in, or suppressed by infusion of, metabolic substrates
(Galbo et al., 1977b; Sotsky et al., 1989). Instead, S activation during exercise is responsible
Hormones and Fuel Use in Exercise 111

16
Control
14
.25LT
.75LT
12
Mean serum GH (μg/L)
LT
1.25LT
10
1.75LT
8 Onset of
exercise
6

0
7:00 7:40 8:20 9:00 9:40 10:20 11:00 11:40 12:20 13:00
Clock time

Figure 5.10 Effect of acute exercise intensity on GH secretion, expressed relative to the lactate threshold (LT).
Reprinted, by permission, from C.J. Pritzlaff et al., 1999, “Impact of acute exercise intensity on pulsatile growth hormone release in men,”
E4517/Borer/Fig5.11/401565/alw/R1
Journal of Applied Physiology 87: 498-504.

for overall coordination of cardiorespiratory and hormonal responses to stress, including


suppression of insulin and stimulation of glucagon, cortisol, and, very probably, GH secretion.
During exercise of short duration at moderate intensities, the counterregulatory hormonal
milieu facilitates lipolysis and high fat utilization in predominantly Type I oxidative muscle
fibers as well as some hepatic glycoge-
nolysis and carbohydrate metabolism. Muscle glycogen
300
Muscle triglycerides
During the first 30 min of exercise, fat
Plasma FFA
Fuel utilization (cal · kg−1 · min−1)

metabolism accounts for 75% to 90% of Plasma glucose


energy utilization (figure 5.11) and plasma
glucose and intramuscular lipids each 200
provide between 10% and 15% of energy
(Romijn et al., 1993, 2000).
At low and moderate exercise intensi-
ties, lipolysis can increase 400-fold above 100
the basal level under the influence of
higher catecholamines and ANP and a
decline in plasma insulin concentration
0
(Lafontan & Langin, 2009). Visceral 25 65
.
85
fat has a higher proportion of lipolytic Percentage of VO2max
β1 and β2 AdRs than of antilipolytic α2
Figure 5.11 Utilization of metabolic fuels originating
AdRs compared with other fat depots, and in the liver, muscle, and adipose tissue as a function of
abdominal subcutaneous (SC) fat shows a exercise intensity. At low exercise intensities, muscle
similar receptor advantage compared with utilizes circulating metabolic fuels. The reliance on car-
E4517/Borer/Fig5.11/401638/alw/R2
femoral and other peripheral SC fat (Arner bohydrate and intramuscular fuels as a source of energy
et al., 1990). This makes the visceral and progressively increases as exercise intensity increases.
abdominal SC adipose fat depots most Reprinted, by permission, from A. Romijn, 1993, “Regulation of
endogenous fat and carbohydrate metabolism in relation to exercise
responsive to catecholamine stimulation intensity and duration,” American Journal of Physiology, Endocrinology
of lipolysis. However, contribution to and Metabolism 265: E380-E391.
112 Advanced Exercise Endocrinology

circulating FFAs by visceral adipose tissue is small (5%-20%) compared with the predomi-
nant contribution by the larger abdominal SC fat depot (Jensen, 2008; Nielsen et al., 2004).
Although GH, ACTH, and cortisol are secreted at moderate to high exercise intensities,
their contributions to exercise metabolism are auxiliary rather than essential because the
blockade of catecholamine action during exercise fails to produce a compensatory increase
in these hormones (Franz et al., 1983)..Similarly, suppression of GH secretion by octreotide
during 30 min of exercise at 70% of VO2max had no effect on glucose, FFA, and glycerol
release (Chalmers et al., 1979). Nevertheless, exercise-associated increases in GH secretion
lead to enhanced postexercise lipolysis from abdominal SC adipose tissue (Enevoldsen et
al., 2007; Pritzlaff et. al., 2000). GH lipolysis reaches peak levels about 2 h after 20 min of
exercise at 70% of VO2max or in response to GH infusion in a pattern that mimics that of
exercise (Wee et al., 2005).
Glucagon and catecholamines, and reduced insulin concentrations, control hepatic glu-
cose production during exercise. In the liver, glycogenolysis and hepatic glucose production
are initiated primarily by E through activation of both alpha and beta AdR by the signal-
ing mechanisms discussed in chapter 1. The role of catecholamines in glycogenolysis is
secondary to their lipolytic role because the threshold E concentration for hepatic glucose
production (150-200 pg/ml) is double the concentration needed for lipolysis (Galster et al.,
1981). During exercise, a catecholamine-induced decrease in plasma insulin increases liver
sensitivity to glucagon. Any reduction in plasma glucose elicits glucagon release, and small
increases in plasma glucagon concentration increase the rate of glucose appearance (Lins et
al., 1983). A decrease in insulin and an increase in glucagon play a decisive role in hepatic
glucose production during the first 90 min of moderate-intensity exercise (Wasserman &
Cherrington, 1996). Declining liver glycogen, sustained α1 adrenergic stimulation, and a
decline in plasma glucose stimulate glucagon secretion, which facilitates splanchnic uptake
of gluconeogenic precursors and gluconeogenesis, whereas beta adrenergic stimulation
increases FFA concentration. After 3 h
of low- to moderate-intensity exercise,
8
glucose production (mg · kg−1 · min−1)

plasma glucose and muscle glycogen


Exercise-induced increase in

stores in Type I muscle fibers are low. Of


the 75 g of glucose produced during 4 h of 6
low-intensity exercise, all but 15 to 20 g is
Glucagon
generated by gluconeogenesis (Ahlborg
et al., 1974). Splanchnic uptake of gluco- 4
neogenic precursors and their conversion
to glucose during prolonged moderate- Insulin
intensity exercise account for about 47% 2
of hepatic glucose output. However, this
process fails to produce enough glucose Epinephrine
to meet peripheral glucose needs, and 0 60 120
plasma glucose concentration decreases Time (min)
(Ahlborg & Felig, 1977). The catechol-
amines assume a predominant role in Figure 5.12 The time course of hormonal control of
hepatic glycogenolysis and gluconeo- hepatic glucoseE4517/Borer/Fig5.13/401640/alw/R1
production during prolonged moderate-
intensity exercise.
genesis at high exercise intensities and
Reprinted, by permission, from D.H. Wasserman and A.D. Cherrington,
during prolonged glycogen-depleting 1991, “Hepatic fuel metabolism during muscular work: Role and
exercise (Marliss et al., 1991; Wasserman regulation,” American Journal of Physiology 260: E811-E824.
& Cherrington, 1996; figure 5.12).
Hormones and Fuel Use in Exercise 113

Hormonal Regulation of Recovery From Low- to Moderate-


Intensity Aerobic Exercise
Hypoglycemia and depletion of liver and muscle glycogen during prolonged submaximal
exercise trigger endocrine responses that influence fuel metabolism during recovery from
such exercise (figure 5.13). Plasma glucose and insulin concentrations remain low and
increase slowly during the initial stage of recovery. Glucose uptake and glycogen resynthe-
sis in the previously exercised muscle are increased three- to fivefold for the next several
hours and take precedence over hepatic glycogen resynthesis (Ahlborg & Felig, 1982; Fell
et al., 1980). Preferential muscle glycogen resynthesis under such conditions is facilitated
by increased intestinal absorption of glucose, increased muscle sensitivity to insulin for
about 48 h after exercise (Mikines et al., 1988), lower Km of muscle HK relative to liver
GK, greater sensitivity of glycogenolysis than gluconeogenesis to inhibition by insulin
(Chiasson et al., 1976), and high residual concentrations of circulating catecholamines and
glucagon that stimulate hepatic glycogenolysis in addition to the prevalent gluconeogenesis.
Muscle glycogen resynthesis is aided by hepatic release of glucose, about 70% of which is
derived from gluconeogenesis (Ahlborg & Felig, 1982). Any acute postexercise ketosis is a
consequence of low plasma insulin and hepatic glycogen depletion (Koeslag, 1982). When
muscle glycogen concentration decreases to less than 30 mM, its resynthesis becomes insulin
independent and proceeds at rates that are proportional to the magnitude of its depletion
(Casey et al., 1995). As muscle glycogen content approaches repletion, its resynthesis again
becomes insulin dependent (Price at al., 1994).

Glycogen depletion
Glycogenolysis (30%) Insulin sensitivity
+
+ Glycogen synthesis

Hypoglycemia G6P
Glucagon
+
Gluconeogenesis (70%)
Insulin
+
Epinephrine Glycogenolysis

Lactate
+

Glycogenolysis

Previously resting muscle

Figure 5.13 Hormonal control of recovery from prolonged low- to moderate-intensity aerobic exercise.
Decreased plasma insulin and increased concentrations of counterregulatory hormones characterize the recovery
E4517/Borer/fig.5.14/401566/TB/R1
from muscle and liver glycogen-depleting exercise. This favors glucose uptake by insulin-sensitive muscle and
continued hepatic glucose production.
Reprinted from O. Bjorkman, and J. Wahren, 1988, Glucose homeostasis during and after exercise. In Exercise, nutrition, and energy
metabolism, edited by E.S. Horton and R.L. Terjong (New York: MacMillan Publishing Company), 111. Reproduced by permission of O.
Bjorkman.
114 Advanced Exercise Endocrinology

Intramuscular TG also are depleted during prolonged low- to moderate-intensity aerobic


exercise. Activity of LPL in the vascular endothelia of the muscle is increased for 3 to 4 h
after termination of exercise (Lithell et al., 1984) and is stimulated by depletion of IMTG
and high titers of glucagon, cortisol, and catecholamines.

Moderate- to High-Intensity Aerobic Exercise


.
At moderate exercise intensities of about 65% of VO2max, approximately one half of the
energy cost of exercise is initially provided by carbohydrate, mostly from muscle glycogen
and about one fifth from hepatic glucose production. The other half of the energy costs
arises from fat utilization, equally provided by IMTGs and peripheral fat depots. In absolute
terms, fat utilization is maximal at a moderate exercise intensity of about 65% (Romijn et
al., 1993, 2000). This pattern of fuel use is appropriate for Type IIa muscle fibers that have
capabilities for both oxidative and anaerobic metabolism (figures 5.11 and 5.14).
In addition to the influence of reduced plasma insulin and increased secretion of counter-
regulatory hormones, increased release of the cytokines leptin and adiponectin from adi-
pose tissue contributes to FFA mobilization and lipid use from IMTG at moderate exercise
intensities (Thamer et al., 2002; Tomas et al., 2004). The effect is in part a consequence of
activation by these cytokines of AMPK and in part a consequence of S activation. Another
cytokine, IL-6, is produced by exercising muscle and can stimulate hepatic glucose produc-
tion, glucose uptake via PI3K, lipolysis in adipose tissue, and fat oxidation via cytokine
receptors and STAT3 signaling (figure 5.15). IL-6 release is proportional to exercise duration
rather than intensity (Febbraio et al., 2004; Pedersen & Febbraio, 2008).
As intensity of exercise increases above 75% of maximal effort, counterregulatory
hormone concentrations and actions increase in parallel. Reliance on utilization of fuels
endogenous to the muscle increases, and carbohydrate oxidation and glycolysis become the
major energy sources in Type IIx fast-twitch glycolytic muscle fibers. Now, 72% of energy
is derived from intramuscular fuel—about 60% is derived from muscle glycogen and about
12% is derived from IMTGs. Only 28% of energy is derived from circulating metabolites—
about 14% each from glucose and FFA (Romijn et al., 1993; figure 5.11).

25 200
Plasma glucagon (pg/ml)
Plasma insulin (µU/ml)

20
150
15

10
100
5

0 50
−50 −10 0 30 60 120 −50 −10 0 30 60 120
a Time (min) b Time (min)
(continued)
E4517/Borer/Fig5.15a/401568/alw/pulled-R1 E4517/Borer/Fig5.15b/401570/alw/pulled-R1
Figure 5.14 Endocrine control of fuel use during intermediate-intensity exercise. Hormonal (above) and metabolic (below)
changes that accompany 1 h of exercise at 50% to 70% of relative effort. Declines in (a) plasma insulin and increases in
(b) plasma concentrations of glucagon, (c) E and NE, (d) GH, and (e) cortisol are shown in the upper panels. Changes in
plasma concentrations of (f) lactate, (g) glucose, and arterial (open circles) and venous (solid circles) concentrations of (h)
FFA and (i) glycerol are shown in the three lower panels. The exercise starts at 0 min.
Figures a, c, d, e, f, g, h, and i adapted, by permission, from V. Hodgetts et al., 1991, “Factors controlling fat mobilization from human subcutaneous
adipose tissue during exercise,” Journal of Applied Physiology 71: 445-451. Figure b adapted, by permission, from H. Galbo, N.J. Christensen, and J.J.
Holst, 1977, “Glucose-induced decrease in glucagon and epinephrine responses to exercise in man,” Journal of Applied Physiology 42: 525-530.
10.0 80

Plasma GH (µU/ml)
8.0
Plasma (pM/ml)

60
6.0 NE
40
4.0 E

2.0 20

0 0
−50 −10 0 30 60 120 −50 −10 0 30 60 120
c Time (min) d Time (min)

E4517/Borer/Fig5.14c/401571/alw/pulled-R2 E4517/Borer/Fig5.15d/401572/alw/pulled-R1
1,000 8.0

Plasma lactate (mM/L)


Plasma cortisol (µM/ml)

800 6.0
600
4.0
400
2.0
200

0 0
−50 −10 0 30 60 120 −50 −10 0 30 60 120
e Time (min) f Time (min)

E4517/Borer/Fig5.15e/401573/alw/pulled-R1 E4517/Borer/Fig5.15f/401574/alw/pulled-R1
5.5 4,000
Plasma glucose (mM/L)

Plasma FFA (µM/L)

5.0 3,000

4.5 2,000

4.0 1,000

3.5 0
−50 −10 0 30 60 120 −50 −10 0 30 60 120
g Time (min) h Time (min)

E4517/Borer/Fig5.14g/401575/alw/pulled-R2 E4517/Borer/Fig5.14h/455918/alw/pulled-R2-kh
800
Plasma glycerol (µM/L)

600

400

200

0
−50 −10 0 30 60 120
i Time (min)

Figure 5.14 (continued) E4517/Borer/Fig5.14i/455917/alw/pulled-R2-kh

115
116 Advanced Exercise Endocrinology

Liver
Blood vessel
IL-6
Contraction Increased hepatic
IL-6
glucose production
IL-6
during exercise

IL-6
IL-6

IL-6Rα/gp130Rβ IL-6

IL-6
P13-K p-STAT3

p-Akt p-AMPK
Adipose tissue

Glucose Fat
uptake oxidation
Increased lipolysis

Figure 5.15 Stimulation of glucose uptake and production and fat oxidation by contraction-induced release of
E4517/Borer/fig.5.10/401563/TB/R1
IL-6 by the muscle.
Reprinted, by permission, from B.K. Pedersen and M.A. Febbraio, 2008, “Muscle as an endocrine organ: Focus on muscle-derived
interleukin-6,” Physiological Review 88: 1379-1406.

Plasma insulin concentrations are low during exercise and are insufficient to stimulate
PDH, the rate-limiting enzyme that commits pyruvate to the TCA cycle. Activation of car-
bohydrate oxidation and PDH at high exercise
intensities is proportional to relative exercise .
5 90% VO2max
intensity. It relies on early release of Ca as muscle
(mmol/kg wet wt/min)

contraction is initiated, feedback from signals of 4 .


PDH activation

65% VO2max
decreasing cellular energy charge, and pyruvate
3
concentration (Howlett et al., 1998; figure 5.16).
Carbohydrate and lipid metabolism at high exer- 2
cise intensities is also increased by E acting on β2 .
1 35% VO2max
AdR in the muscle to simultaneously stimulate
glycogenolysis, production of LA, and glucose 0 2 4 6 8 10
oxidation (Kjaer et al., 2000) as well as HSL and Time (min)
lipolysis (Donsmark et al., 2005).
E and cortisol, two hormones that mediate Figure 5.16 Activation of PDH at aerobic
increased hepatic gluconeogenesis and muscle exercise of different intensities.
E4517/Borer/Fig5.17/401567/alw/R1
glycogenolysis, engage in a positive-feedback Reprinted, by permission, from R.A. Howlett et al., 1998,
“Regulation of skeletal muscle glycogen phosphorylase and
loop when intensity of exercise or other stress PDH at different exercise power outputs,” American Journal
is high. E then increases cortisol secretion by of Physiology 75: R418-R425.
Hormones and Fuel Use in Exercise 117

stimulating the pituitary release of ACTH, and cortisol stimulates the synthesis and release of
E (see figure 3.18). The metabolic consequence of this emergency S and endocrine response
to higher-intensity exercise is hyperglycemia (figure 5.17). A high level of S activation causes
alpha adrenergic vasoconstriction of vascular beds other than those in muscle. This sequesters
FFA in the adipose tissue; the FFA will be released when exercise ceases and S tone decreases.
A postexercise rebound increase in FFA concentration is caused by the removal of S circulatory
restraint over FFA transport out of adipose tissue. Increased postexercise fat metabolism is in
part due to stimulation by GH, the release of which coincides with FFA rebound (Henderson
et al., 2007), and in part a result of increased IMTG utilization and E-stimulated blood flow to
adipose tissue (Horowitz & Klein, 2000).
Muscle lipid metabolism at high exercise intensities is activated by muscle contraction
through activation of Ca-binding and energy-sensing enzymes that increase the activity of
ATGL in an intensity-dependent manner. In both Type I oxidative and Type II muscle, the
effects of E and contractions on stimulation of HSL activity are partially additive (Langfort
et al., 2003). Thus, at higher exercise intensities, endogenous signals of energy need appar-
ently play a greater role in the regulation of muscle fuel metabolism than do hormones.

Low-intensity exercise
40 1.4
1.2
µmol ∙ kg−1 ∙ min−1

30 1.0
mmol/L

0.8
20 25% 0.6
10 0.4
0.2
0 Exercise 0 Exercise
0 60 120 180 0 60 120 180
Intermediate-intensity exercise
40 2.0
µmol ∙ kg−1 ∙ min−1

30 1.6
mmol/L

1.2
20 65%
0.8
10
0.4
0 Exercise 0 Exercise
0 60 120 180 0 60 120 180
High-intensity exercise
60 1.4
50 1.2
µmol ∙ kg−1 ∙ min−1

85%
40 1.0
mmol/L

0.8
30
0.6
20 0.4
10 0.2
0 Exercise 0 Exercise
0 30 60 90 0 30 60 90
Time (min) Time (min)

Figure 5.17 Influence of exercise intensity on the rates of glucose appearance (left) and of glycerol (open circles)
E4517/Borer/Fig5.18/401577/alw/pulled-R1
and FFA (solid circles) appearance into plasma (right).
Adapted, by permission, from J.A. Romijn, 1993, “Regulation of endogenous fat and carbohydrate metabolism in relation to exercise intensity
and duration,” American Journal of Physiology 265: E380-E391.
118 Advanced Exercise Endocrinology

Hormonal Regulation of Recovery From Moderate- to High-


Intensity Aerobic Exercise
A rebound increase in plasma insulin concentration that results from a decline in S tone
occurs within minutes of termination of high-intensity exercise (figure 5.14) and is accom-
panied by a rapid disappearance of catecholamines from plasma (Hagberg et al., 1979).
Increased plasma insulin concentration facilitates glucose uptake by the recovering Type
IIa and Type IIx muscle fibers (figure 5.18). Glucose uptake by muscle remains three to four
times higher than at rest for about 1 h, after which it declines rapidly (Wahren et al., 1973).
Splanchnic glucose output decreases and reaches basal level within 40 min. During
recovery from intense exercise, gluconeogenesis contributes about 40% of hepatic glucose
output. This metabolic pattern is facilitated by continuing high glucagon concentration, high
gluconeogenic precursors, and increased splanchnic blood flow. Oxidative metabolism of
lipids, ketone bodies, and LA is gradually replaced by oxidation of mixed fuels as concen-
trations of insulin and counterregulatory hormones return to resting postabsorptive levels.

Previously exercised muscle

Glycogenolysis (60%)

Glycogen synthesis
Basal glucose
levels G6P
– +
Gluconeogenesis (40%) Insulin
Glucagon
+

Lactate
(elevated levels)

Figure 5.18 Hormonal control of recovery from short, high-intensity exercise.


E4517/Borer/Fig5.19/401578/alw/pulled-R1
Reprinted from O. Bjorkman and J. Wahren, 1988, Glucose homeostasis during and after exercise. In Exercise, nutrition, and energy
metabolism, edited by E.S. Horton and R.L. Terjung (New York: MacMillan Publishing Company), 111. Reproduced by permission of O.
Bjorkman.

HORMONES IN FUEL MOBILIZATION AND UTILIZATION


DURING ANAEROBIC EXERCISE OR RE
The well-known effects of resistance training on muscle strength and hypertrophy have
directed much of the research into the ability of RE to stimulate the release of the anabolic
hormones GH, T, and IGF-I and the catabolic hormone cortisol and toward research into
the relationship between the changes in the concentrations of these hormones and protein
synthesis and muscle hypertrophy. The relevance of this pattern of hormone release to muscle
Hormones and Fuel Use in Exercise 119

hypertrophy is discussed in chapter 8. The present section describes the main endocrine
responses to different types of RE; the contribution of reduced muscle oxygen availability
to hormone secretion during RE; the effects of RE on myofiber hormone receptor, enzyme,
and gene expression; and some metabolic consequences of RE.

Endocrine Responses to RE
Description of endocrine responses to RE is complicated by the often discontinuous nature
of such exercise, the use of the upper or lower limbs or both, and the different patterning
and types of movement. Several generalizations facilitate discussion of hormone responses
to different types of RE. First, the core goal of RE is to increase muscle strength. To this
end, exercise stimulus consists of shorter episodes of mechanical loading of specific limbs or
joints rather than repetitive and prolonged locomotor or similar propulsive exercise. Because
of the relatively high intensity and brevity of RE, this type of movement is predominantly
anaerobic and is usually quantified by the magnitude of LA release. It is reasonable to expect
that shorter, intermittent bouts of high-resistance movement would produce a different
pattern of hormone secretion than would energy-expensive, prolonged episodes of aerobic
exercise. Second, RE can be characterized by three main patterns of mechanical loading:
one focused on increasing strength (SRE), one focused on increasing muscle power (PRE),
and one focused on maximizing muscle hypertrophy (HTRE). The heaviest load that can be
lifted at one time—1RM—is a measure of RE intensity and of muscle strength. The three
RE patterns differ in intensity, volume, duration, and speed of movement; they also elicit
different patterns of hormone secretion (Kraemer & Ratamess, 2005).

Hormone Secretory Responses to RE Protocols


SRE protocol applies the highest loading intensity—about 85% to 90% of RM, usually in
3 to 5 sets of 3 to 5 repetitions—and inserts extended rest periods of 3 to 5 min between
the sets (Fleck & Kraemer, 2004). Increases in absolute load to maintain relatively high
load intensity ensure continued increases in strength. The intensity of this approach is high;
however, the total volume of exercise (i.e., number of muscles loaded and number of sets,
repetitions, and exercises) is not particularly high. PRE requires high-velocity total-body
movements such as jump squats at a lower intensity of about 30% to 50% RM, substantial
loading volume (3-5 sets of 5-6 repetitions), and moderate rest periods of about 2 to 3 min.
The greatest amount of interest and research has focused on HTRE, which usually engages
a large muscle mass (e.g., the lower limbs) at moderate intensities of 60% to 80% of RM and
includes up to 15 repetitions per set. Given that between 4 and 8 sets are usually performed
and that each lasts up to 1 min, this approach is most metabolically challenging and most
fatigue-inducing because the rest periods between the sets are short (1 min). As in SRE,
relative load in HTRE is progressively increased and the number of repetitions per set is
reduced to promote increases in strength.
When the three patterns of RE performed with the lower limbs were compared in the
same study (McCaulley et al., 2009), HTRE generated an approximately sixfold increase
in LA concentration, whereas SRE generated an approximately fourfold increase in LA
concentration and PRE generated no change. Only HTRE produced significant increases
in total T (32%), cortisol (12%), and SHBG (20%); SRE and PRE elicited no endocrine
responses. In most studies contrasting SRE and HTRE (reviewed in Kraemer & Ratamess,
2005), GH response to acute SRE was either absent or low, but GH response to an HTRE
protocol was significantly higher. When the relative contribution of an HTRE protocol
consisting of a high number of moderate-intensity repetitions is compared with an SRE
120 Advanced Exercise Endocrinology

protocol consisting of half as many repetitions at twice the intensity and only 15% difference
in total work, the LA response is 2 times higher and GH response is about 8 times higher in
the HTRE protocol than in the SRE protocol (Kraemer et al., 1990; figure 5.19). Both the
volume and the stressfulness of the protocol contribute to the high GH response in HTRE
protocol because increasing the rest period from 1 to 3 min (figure 5.19, F) or reducing the
repetitions from 10 to 5 (figure 5.19, E) abolish the eightfold elevation of GH (figure 5.19,

1,000
GH AUC

800

A 5RM, 3 min rest D 10RM, 1 min rest


600 B 10 reps, 3 min rest E 5 reps, 1 min rest
C 5RM, 1 min rest F 10RM, 3 min rest

400

200

0
A B C D E F

Strength protocol Hypertrophy protocol

32.0
Serum human growth hormone (µg/L)

28.0
D
24.0

20.0

16.0
C F
12.0
A B E
8.0

4.0

0
Pre Mid 0 5 15 30 60 90 120 Pre Mid 0 5 15 30 60 90 120
Exercise Postexercise Exercise Postexercise
(min) (min)

Figure 5.19 Robust GH secretory response to the HTRE training protocol compared with the GH secretory
response to the SRE training protocol. SRE (lower left, solid line) elicited one eighth as much GH as did HTRE
(lower right, solid line). In both protocols, 1 min rest periods (C and D, upper panel) and large exercise volume (D
vs. E, upper panel) contributed to the magnitude of GH response.
E4517/Borer/Fig5.19/401579/alw/R3-kh
Adapted, by permission, from W.J. Kraemer et al., 1990, “Hormonal and growth factor responses to heavy resistance exercise protocols,”
Journal of Applied Physiology 69: 1442-1450.
Hormones and Fuel Use in Exercise 121

D). Conversely, reducing the rest period in the SRE protocol approximately doubles the
GH response (figure 5.19, C), whereas reducing the load to 10RM with a long rest period
is ineffective (figure 5.19, B). Therefore, of the commonly measured endocrine responses
in RE, GH appears to be the most sensitive to metabolic load and exercise volume (Leite
et al., 2011; Uchida et al., 2009).
Three additional features of GH secretion in response to RE are noteworthy. First, GH
is secreted as variants with a range of molecular sizes (<30 kD, 30 to 60 kD, and >60 kD).
The most commonly measured variant has a molecular weight of 22 kD. A single bout of
heavy RE stimulates the release of fractions of low and middle molecular weight that are
detectable by RIA but not the release of the biologically active hormone measured by one
of the two available bioassays (Kraemer et al., 2008; Kraemer & Ratamess, 2005). And, as
is the case with the stimulatory influence of estrogen on GH secretion in general, HTRE
protocol elicits a higher GH response when estrogen levels are increased by use of oral con-
traceptives (Kraemer et al., 2008). Second, bed rest or lack of gravitational pull suppresses
bioassayable GH secretion for as long as muscle activity is suppressed (McCall et al., 1997,
1999). GH response to RE is restored several days after the resumption of muscle activity,
suggesting that some form of afferent feedback related to muscle activity is necessary for the
maintenance of RE-induced GH response. Third, nocturnal GH secretory pattern is altered
in response to a high-volume (50 set) HTRE protocol (Tuckow et al., 2006). Nocturnal GH
secretory burst frequency and disorderliness of secretion (approximate entropy) are increased
and hormone burst mass and secretory rate and half-time are reduced, whereas the total
amount of GH secreted in a day is not affected.
Finally, although it is usually assumed that GH secretion is responsible for increased
hepatic IGF-I release, the secretory responses of the two hormones are dissociated in exer-
cise. HTRE protocol strongly stimulates GH secretion (figure 5.19; Kraemer et al., 1990);
however, the same acute exercise parameters have little effect on IGF-I or its GH-stimulated
BP, IGFBP-3 (Manetta et al., 2002). Furthermore, resistance training increases GH responses
to acute RE (Kraemer & Ratamess, 2005), but IGF-I concentration often decreases during
the early weeks of exercise training (Rarick et al., 2007), does not respond to acute RE,
and becomes elevated only after longer exposure to training (Koziris et al., 1999; Nindl &
Pierce, 2010).
Given that cortisol leads to muscle atrophy but also has a possible gluconeogenic action,
researchers are greatly interested in its secretory pattern during RE. An HTRE protocol
involving the lower limbs increases plasma concentrations of ACTH more in untrained young
men than in older men, whereas cortisol concentration is increased to the same extent in
both age groups (Kraemer et al., 1999).
In contrast to GH response to the two training protocols, total testosterone (T) response
was greater in the SRE protocol than in the HTRE protocol. The intensity of exercise was
more stimulatory, and the duration of rest periods was less important for T response to the
SRE protocol than was the case for GH response in the same study (Kraemer et al., 1990;
figure 5.20).
Additional factors that modify the magnitude of GH, T, and cortisol secretory responses
to different RE protocols include hypoxia, anticipation of stress, mass of muscle recruited,
gender, age, and dietary supplements. Reduced oxygen availability to exercising muscle,
whether produced by restriction of blood flow during moderate-intensity RE (Reeves et
al., 2006) or the exposure to 13% rather than 20% atmospheric oxygen (Kon et al., 2010),
significantly increases the magnitude of LA production and GH, E, and NE secretion but
has little impact on T or cortisol secretion. Anticipatory increases in E, NE, and DA before
122 Advanced Exercise Endocrinology

600
Testosterone AUC
A 5RM, 3 min rest D 10RM, 1 min rest
B 10 reps, 3 min rest E 5 reps, 1 min rest
400 C 5RM, 1 min rest F 10RM, 3 min rest

200

−200
A B C D E F

Strength protocol Hypertrophy protocol


C E
D
31.0

29.0
Serum testosterone (nmoI/L)

A
27.0

25.0

23.0

21.0

19.0

17.0
B F
15.0

Pre Mid 0 5 15 30 60 90 120 Pre Mid 0 5 15 30 60 90 120


Exercise Postexercise Exercise Postexercise
(min) (min)

Figure 5.20 Robust total T secretory response to an SRE protocol relative to an HTRE protocol. In response to
higher exercise intensity (5RM) in the SRE protocol, total T response was about double that (A and C top left and
solid circles, bottom left) obtained with the HTRE protocol (D-F, top right). Difference in the rest period (3 min in A
and F vs. 1 min in C-E) did not change total T response, whereas cutting the intensity of SRE in half (B) abolished it.
E4517/Borer/Fig5.20/401580/alw/R4-KH
Adapted, by permission, from W.J. Kraemer et al., 1990, “Hormonal and growth factor responses to heavy resistance exercise protocols,”
Journal of Applied Physiology 69: 1442-1450.

and during high-volume HTRE protocol suggest that that this most likely represents the
contribution of general stress to endocrine response (French et al., 2007). Similarly, loss of
between 2.5% and 5% of body mass by dehydration represents another stressor that increases
cortisol and NE responses to RE (Judelson et al., 2008). The volume of muscle used in RE
is positively related to the magnitude of GH response (Kraemer & Ratamess, 2005; West et
al., 2010). Endocrine responses to RE in some instances are smaller in females than in males
Hormones and Fuel Use in Exercise 123

(Kraemer et al., 1991). Both sexes display increased GH responses to an HTRE protocol,
but females fail to increase GH in response to an SRE protocol. IGF-I concentrations in
both sexes increase in response to both SRE and HTRE protocols, but increases in total T
in response to both protocols are significantly greater in males than in females. Both GH
and T responses to RE decline with age; after middle age, T responses are proportional to
muscle strength (Baker et al., 2006; Roberts et al., 2009).
Finally, some dietary supplements alter hormonal responses to RE. Administration
of caffeine (a sympathomimetic agent) increases T response to RE in a dose-dependent
fashion, suggesting that catecholamines may mediate this endocrine response (Beaven et
al., 2008; Fatouros et al., 2010). Carbohydrate supplementation was seen to reduce corti-
sol response to RE in some studies, suggesting that increased secretion of this hormone
may play a gluconeogenic role (Kraemer & Ratamess, 2005). Similarly, amino acid and
dipeptide supplementation reduces E, NE, and GH responses to RE, suggesting again that
a biological role of these endocrine responses may be to provide amino acid substrates for
gluconeogenesis (Goto et al., 2011; Sharp & Pearson, 2010).

Effects of RE on Muscle Hormone Receptors, Enzymes,


and Genes
Androgen receptors (ARs) in skeletal muscle are upregulated by high plasma titers of T
and natural or electrically stimulated contractile activity. The mechanism of their anabolic
actions is described in chapter 1 and is further discussed in chapter 7. HTRE causes a short-
term decline in ARs despite the increases in T secretion, whereas GRs undergo no change
(Vingren et al., 2009). On the other hand, potentiation of circulating T concentration by
high-volume upper-body exercise before lower-limb RE resulted in a relative increase in
exercising muscle AR (Spiering et al., 2009). Protein–carbohydrate supplementation before
and after training blunts the magnitude of initial AR decline (Kraemer & Ratamess, 2005).
Although RE may not alter GRs in contracting skeletal muscle, it apparently increases
systemic activity of 11β-HSD type 1, the enzyme that converts cortisone to biologically
active cortisol. The functional significance of this effect of RE may be to increase cortisol
concentration in loaded skeletal muscle and limit exercise-induced inflammatory response
(Dovio et al., 2010). Heavy RE that is associated with postexercise muscle soreness and
enzymatic signs of muscle damage is not associated with changes in AR or their expression;
however, the expression of IGF-IEa and mechano growth factor (muscle version of IGF-I)
is increased by 6 d postexercise and may represent acute regenerative processes after RE-
induced muscle injury (Ahtiainen et al., 2011).

Metabolic Effects of RE
Researchers have not studied the metabolic consequences of RE as extensively as they
have studied the metabolic consequences of endurance exercise. However, three areas have
attracted interest: post-RE stimulation of resting metabolism, changes in glucose tolerance
and insulin sensitivity, and clearance of postprandial lipemia. In a study that compared HTRE
with SRE and with an intermediate RE protocol, 1 h of exercise at the lower intensities of
the HTRE and intermediate RE protocols elicited an LA response that was 2 times higher
and cortisol and FFA concentrations that were 3 times higher than those of the SRE protocol
(Fatouros et al., 2009). The HTRE and intermediate RE protocols also were associated with
a 5% increase in resting energy expenditure (REE) that peaked 12 h after the termination of
exercise, whereas the SRE protocol was associated with a 7.5% increase. REE returned to
124 Advanced Exercise Endocrinology

baseline by 24 h postexercise after the HTRE and intermediate RE protocols, whereas REE
elevation after the SRE protocol, along with an increase in the concentration of adiponectin
seen from 12 to 24 h, persisted through 48 h.
Studies have examined the capacity of RE to increase glucose tolerance and insulin
sensitivity, often using diabetic or obese subjects. An acute HTRE bout consisting of 3 sets
of 8 to 12 repetitions reduces whole-body insulin action, as shown by the reduced glucose
infusion rate necessary to maintain euglycemia during a euglycemic–hyperinsulinemic
clamp immediately after exercise in healthy subjects (Howlett et al., 2007). This effect
is accompanied by reduced mobilization of glucose transporters (as reflected by reduced
phosphorylation of AS160 protein) to muscle membrane. By contrast, a higher-intensity SRE
protocol appears to lead to better glucose tolerance, as assessed immediately after exercise
with an oral glucose tolerance test (OGTT), compared with an HTRE protocol of equal
energy expenditure (Miller et al., 2007). Similar findings were reported in two studies of
subjects with T1D in which a single bout of SRE consisting of 5 sets of 6 repetitions had no
effect on insulin sensitivity 12 and 36 h postexercise as measured by euglycemic–hyperin-
sulinemic clamp (Jimenez et al., 2009). In another study using subjects with T1D, a single
bout of HTRE consisting of 3 sets of 10 repetitions resulted in no change in glucose clear-
ance but an apparent increase in insulin clearance (Fluckey et al., 1994).
During 2 h postexercise, a single bout of HTRE consisting of 10 sets of 10 repetitions
increases glucose transporter translocation to the muscle membrane and glucose uptake by
170% (Dreyer et al., 2008). At 24 h postexercise, insulin sensitivity as measured by an insulin
tolerance test is enhanced after a single bout of HTRE (Koopman et al., 2005). Reduction of
whole-body insulin sensitivity after an acute HTRE protocol no longer occurs when HTRE
is repeated on 3 alternate days (Howlett et al., 2007). Thus, RE appears to cause an acute
reduction and a postexercise increase in glucose tolerance and insulin action.
Interest in the capacity of RE to lower postprandial lipemia stems from the association
of high plasma lipids with endothelial dysfunction, atherogenesis, and metabolic syndrome.
The effectiveness of RE in lowering plasma TG and VLDL cholesterol is typically tested by
having the subjects ingest a high-fat meal 14 to 16 h after performing RE. An SRE protocol
of eccentric exercise that produced muscle soreness reduced postprandial lipemia by 12%
despite the brevity and low energy expenditure in this protocol (Pafili et al., 2009) Other
studies, typically using an HTRE protocol, found that a similar (~20%) reduction in post-
prandial lipemia was produced by exercise varying in energy expenditure by a factor of 2
(180 vs. 335 kcal; Zafeiridis et al., 2007) and by exercise varying in intensity by a factor of
2 (Singhal.et al., 2009). In a comparison of 3 sets of HTRE and 90 min of aerobic exercise
at 30% of VO2max, both requiring 400 kcal of energy expenditure, RE reduced postprandial
lipemia by 28%, whereas aerobic exercise was ineffective (Magkos et al., 2008).

SUMMARY
This chapter first discusses hormonal control of metabolism under resting conditions and
outlines the key role of insulin in stimulating sugar and amino acid uptake; storing nutrients
as glycogen, fat, and protein; carbohydrate metabolism; and suppressing metabolic fuel mobi-
lization. This chapter then discusses the role of the counterregulatory hormones glucagon,
cortisol, GH, and catecholamines in the mobilization of glycogen stores by glycogenolysis
and gluconeogenesis and the mobilization of fat stores by lipolysis. The hormonal shifts,
including suppression of insulin secretion and activation of counterregulatory hormones,
are discussed separately for aerobic or endurance exercise and anaerobic exercise or RE.
125
126 Advanced Exercise Endocrinology

This chapter describes different patterns of hormone secretion to account for different pat-
terns of fuel use as a function of intensity of aerobic exercise. The role of hormones in the
shift of exercise metabolism from lipid to carbohydrate oxidation is shown to diminish with
increasing exercise intensity. Interaction between tissue depletion of fuel stores and endocrine
responses to these changes during recovery from shorter, more intense exercise is compared
with recovery from longer, glycogen-depleting exercise. This chapter outlines the hormonal
responses to three RE protocols and shows that GH secretion is greater in response to the
HTRE protocol than in response to the SRE and PRE protocols, and T secretion is greatest
in response to SRE protocol. Finally, this chapter discusses the effects of resistance training
on expression of hormone receptors, genes, and enzymes in the muscle and the metabolic
effects of RE on outcomes other than muscle hypertrophy.
H APTER

6
C

Hormonal Control of
Energy Expenditure
and Intake
T
he prevalence of excess body weight increased dramatically in recent decades in
developed countries. In the United States alone during 2009 and 2010, more than 60%
of the adult population was overweight, about 36% was obese, and 3% was morbidly
obese (Flegal et al., 2012). Overconsumption of energy-rich food and decreased physical
activity are lifestyle factors that contribute to the prevalence of excess body weight. Normal-
weight individuals maintain a stable adult body mass (or setpoint) through intermittent
feeding, physical activity, and changes in metabolism. How they do so is not yet completely
understood, but the process clearly involves the interaction between regulatory brain centers,
autonomic reflexes, systemic hormones, and tissue enzymatic activity. All are reactions
to changes in energy availability both in the external environment and in our bodies and
influence feeding, physical activity, and metabolic processes. This chapter outlines current
information and controversies on how these neural, hormonal, and metabolic factors shape
the defense of the body mass setpoint on one hand and contribute to its upward movement
toward obesity on the other. Factors that contribute to controversies and some confusion
about the regulation of body mass include the following: uncertainty about the comparative
importance in this process of internal energy availability, level of adiposity, food intake,
or body mass setpoint; disagreements about descriptors and controls of feeding behavior;
misunderstanding of the role of physical activity in the regulated process; and unsatisfac-
tory conceptual integration of homeostatic (need based) and nonhomeostatic (socially or
environmentally influenced) controls in body-mass regulation. This chapter also assesses
several approaches to weight loss that focus on exercise.

127
128 Advanced Exercise Endocrinology

ENERGY-REGULATING MECHANISM
In at least one third of Americans and larger proportions of the adult population in less-
advanced countries, when prepubertal statural growth ceases, body mass remains remarkably
stable over extended periods of the life span despite fluctuations in the levels of individuals’
energy consumption and expenditure. Intermittent food intake by these individuals matches
energy expenditure that includes resting metabolism, which accounts for the cost of auto-
matic life-sustaining processes and structural maintenance, as well as physical movement
and work. This suggests that an energy-regulating mechanism or energy equation exists
that matches calories lost with calories eaten.

Energy Equation as a Mechanism of Energy Regulation


Based on the composition of adipose tissue (83% fat, 2% protein, and 15% water), an energy
equation evolved that connects a gain of 1 kg of fat to overeating about 7550 kcal and asserts
that losing the fat requires simply expending or withholding the same number of calories.
Many studies have tested this concept by devising different combinations and rates of reduc-
tion of energy availability using food restriction or exercise. The relative ease of body fat
accretion and the significant difficulty of maintaining body fat and body mass loss below an
either normal or obese body mass plateau show that the equation does not work this simply.
Weight gain is facilitated by choosing easily available and highly palatable, energy-dense
food over equally available food of lower energy density and high fiber content. On such
cafeteria choices, rats (Sclafani & Springer, 1976), first-year college students (Cluskey &
Grobe, 2009; Levitsky et al., 2004), and more than two thirds of American adults (Flegal et
al., 2012) tend to regularly gain body fat. Social interaction, supply of large food portions,
and a number of other environmental factors (Levitsky, 2005; Speakman et al., 2011) also
support food intake and weight gain. Driving the putative energy equation in the opposite
direction is very difficult. Although deliberate withholding of food and increased energy
cost of mandated exercise can drive body fat and body mass down, increased hunger,
decreased S and hormonal activation of metabolism, and increased insulin sensitivity and
its energy storage actions all tend to slow weight and fat loss and accelerate the return to
predeprivation body mass levels whenever restraints over feeding and mandated exercise
are removed (Johannsen et al., 2012).
The difficulty of maintaining losses in body mass and fat demonstrates the operation
of regulation that defends lean tissue, body fat, and bone against losses of these body
components through compensatory increases in hunger, food seeking and ingestion, and
other metabolic adjustments that favor regaining any losses (Adolph, 1947). This regulation
becomes expressed after cessation of statural growth and involves coincident appearance
of four developmental events that characterize adult mammals: the onset of high volumes
of spontaneous physical activity, sexual maturity, cessation of skeletal growth, and scaling
of food intake to unit body mass (Kennedy, 1957; Kennedy & Mitra, 1963a). The weight
plateau of a mature female rat oscillates around a stable setpoint in parallel with changes
in the levels of estrogen and spontaneous locomotor activity (Kennedy & Mitra, 1963b).

Homeostatic Lipostatic Concept of Body Fat Regulation


The homeostatic lipostatic model of body fat regulation grew from the energy equation con-
cept through a confluence of discoveries about brain substrates and hormones that control
feeding and body fat storage. Gordon Kennedy identified the ventromedial hypothalamus
(VMH) and arcuate nucleus (ARC) in the medial basal hypothalamus as the sites that are
Hormonal Control of Energy Expenditure and Intake 129

involved in the previously mentioned four features of adult energy regulation (Kennedy,
1950, 1953). After damage to these centers, rats attained a new, higher weight and body fat
plateau through transient increases in food intake and permanent decline in spontaneous
levels of physical activity. Subsequent experiments with parabiotic rats suggested that a
circulating factor arising in the adipose tissue of an obese lesioned rat suppressed feeding
and produced weight loss in a nonlesioned parabiotic partner (Hervey, 1959). This nascent
lipostatic concept of body mass regulation was reinforced by two additional developments:
discovery of genetically obese rodents and discovery of the hormone leptin (Zhang et
al., 1994) secreted from the adipose tissue in proportion to the mass of adipose tissue or
adipocyte size (Considine et al., 1996). It soon was established that ob/ob mice developed
obesity because of their inability to produce leptin and that obese and diabetic db/db mice
and fa/fa rats lacked functional leptin receptors (LepRb) in the medial basal hypothalamus.
Systemic or intracerebroventricular administration of leptin suppressed feeding and reduced
body fat mass in ob/ob mice (Pelleymounter et al., 1995), genetically obese fa/fa rats, and
humans bearing a similar genetic defect (Farooqi et al., 1999). Furthermore, a parabiotic
circulatory union between an ob/ob mouse unable to produce leptin and an obese diabetic
db/db mouse lacking LepRb receptors but oversecreting leptin caused the ob/ob mouse to
lose weight and die of starvation (Coleman, 1973; Coleman & Hummel, 1969). Mediation by
leptin of changes in feeding and body fat mass in leptin-deficient rodents and humans has
therefore prompted a homeostatic concept
of body fat regulation in neurologically
intact animals and humans that is mod-
eled after engineering negative feedback
control. This concept posits that the sta-
bility and defense of body mass against
either upward or downward excursion is
based on the regulation by leptin of body
fat mass through its inhibitory action + –
on the neurons in the ARC and VMH, Energy Energy Leptin
expenditure intake Insulin
a view that persists to date (Guyenet & +
Schwartz, 2012; Schwartz et al., 2000).
The key tenets are that increases in fat
mass trigger increased basal leptin (and
insulin) secretion and that these hor- +
mones, in turn, prevent obesity through
homeostatic suppression of feeding and – Fat stores
increased energy expenditure via meta-
bolic thermogenesis and increased physi- Figure 6.1 The prevailing homeostatic concept of
E4517/Borer/fig.6.1/401582/TB/R3-alw
cal activity. Conversely, decreases in body body fat regulation. Leptin is released from distributed
fat mass lead to decreased leptin secretion, adipose tissue sites in proportion to their mass or size of
adipocytes. This signal, as well as basal insulin concen-
which is postulated to increase hunger and
trations that also change in proportion to body fat level,
decrease metabolic and behavioral energy is assumed to directly act on the LepRb leptin receptors
expenditure (figure 6.1). According to this and IR in the VMH. There, they inhibit energy intake and
view, lesions of ARC neurons destroy increase spontaneous physical activity when adiposity is
leptin receptors (LepRb), the principal high and increase food intake and reduce spontaneous
target of leptin negative feedback. The activity when body fat level is reduced.
hyperphagia that results from such neural Reprinted, by permission, from M.W. Schwartz and K.D. Niswender,
2004, “Adiposity signaling and biological defense against weight gain:
damage is considered to be a chronic and Absence of protection or central hormone resistance?” Journal of
dominant change that produces obesity. Clinical Endocrinology and Metabolism 89: 5889-5897.
130 Advanced Exercise Endocrinology

However, several major problems arise with this view: rapid development of dietary
obesity in parallel with increasing concentrations of leptin upon exposure of animals to
high-fat diets (Frederich et al., 1995), which represents a failure of putative antiobesity
negative feedback and is attributed to obesity-induced resistance to leptin action (Fam et
al., 2007; Schwartz & Niswender, 2004); failure of physiological doses, and up to tenfold
higher doses, of exogenous leptin to cause substantial body fat losses in obese humans who
are able to synthesize leptin (Heymsfield et al., 1999); misunderstanding of the relationship
of physical activity with body fat mass or its role in the regulation of body mass; and focus
on body fat as the key regulated component of body mass even though it represents only
between 15% and, at most, 50% of body composition.

Regulation of Food Intake


The constancy of food eaten under steady-state conditions in mature animals that maintain
a stable weight plateau—unlike high food intakes in growing animals—has prompted the
concept that animals sense caloric deficit and match it with calories eaten. One way to test
this hypothesis is to provide food of different energy densities but equal sensory proper-
ties and palatability to see whether feeding behavior compensates for difference in caloric
content of the diet. In a study in which human volunteers received the same type of food
for 11 wk, the volume of food eaten was identical for diets that contained either 25% or
35% fat (Kendall et al., 1991). The daily difference between equal portions of the two diets
(~200 kcal) helped the group receiving the lower-fat diet to lose weight compared with the
group receiving the higher-fat diet. This suggests that in meal-to-meal eating, energy-replete
humans do not track calories in their food but instead eat a constant volume of food.
Another test of the hypothesis of calorie sensing and tracking is to vary energy expen-
diture through regular physical activity. In a study comparing the effects of energy cost of
exercise and exercise intensity on ratings of hunger, energy expenditure through exercise
was shown to acutely decrease. hunger sensation (King et al., 1994). The expenditure of
about 360 kcal at 36% of VO2max had no effect on the ratings of hunger before.the midday
meal (figure 6.2a, open circles), whereas expenditure of 340 kcal at 72% of VO2max led
to a transient 45% suppression of hunger. The . size of energy cost of exercise also played a
role. The expenditure of 540 kcal at 75% of VO2max suppressed hunger approximately two
times more than did expenditure of about half as many calories at the same exercise intensity
(figure 6.2b). Because exercise energy expenditure should have produced compensatory
increase in hunger rather than hunger suppression, the hypothesis of automatic tracking of
dietary calories was again not supported.
Further, if calories lost during exercise were tracked, a delayed compensatory increase
in food consumption would be expected. As shown in figure 6.3a, daily energy intake
after two levels of exercise energy expenditure in the study just described (see figure 6.2b;
King et al., 1994) did not differ from intake on a sedentary day. When calories expended
were subtracted from calories eaten, fewer daily calories were consumed after exercise
than in the sedentary trial (figure 6.3b). No increases in food intake were detected during
subsequent days.
If human hunger was responsive to changes in energy availability, reducing the amount
of daily physical activity should lead to isocaloric reduction in hunger and food intake.
Instead, reducing the opportunity for physical activity in the short term by confining the
study subjects to a small metabolic chamber during 1 d did not affect the amount of food
eaten. Instead, it produced a significant increase in energy balance that favored fat deposi-
tion when the diet was energy dense and high in fat (Shepard et al., 2001).
Hormonal Control of Energy Expenditure and Intake 131

140 140

120 120
Mean rating (mm)

Mean rating (mm)


100 100

80 80

60 60

40 40
Exercise Exercise
20 20
Breakfast Lunch Breakfast Lunch
0 0
8:30 9:00 11-11:30 12:00 12:15 1:00 8:30 9:00 11-11:30 12:0012:05 12:10 12:15 Pre Post
a.m. a.m. a.m. a.m.
a Time of day b Time of day

Figure 6.2 Suppression of hunger by energy cost and. intensity of exercise. (a) Hunger was 45% lower when 340 kcal
were expended during 30 min of cycling at about 72% . of VO2max (solid circles), whereas hunger was unaffected when 360
kcal were expended during 1 h of cycling at 36% of VO2max (open circles) or during a sedentary trial (rectangles). (b) Com-
E4517/Borer/Fig6.2a/401583/alw/R1
pared with the sedentary
. trial (rectangles), hunger was 60% lower when 540E4517/Borer/Fig6.2b/401642/alw/R2
kcal were expended during 52 min of exercise
at about 75% of VO2max (solid circles) and 40% lower when 300 kcal were expended during 30 min of cycling at the same
intensity (open circles).
Adapted, by permission, from N.A. King, V.J. Burley, J.E. Blundell, 1994, “Exercise-induced suppression of appetite: Effects on food intake and implications
for energy balance,” European Journal of Clinical Nutrition 48: 715-724.

3,000 2,000

2,000
Intake (kcal)

Intake (kcal)

1,000

1,000

0 0
a Sedentary Low volume High volume b Sedentary Low volume High volume

Figure 6.3 No postexercise compensation for calories expended. (a) Daily energy intake during the exercise experiment
E4517/Borer/Fig6.3a/401584/alw/R2
described in figure E4517/Borer/Fig6.3b/401643/alw/R1
6.2b. (b) Relative intake reflects the difference between kilocalories consumed and kilocalories expended
exercising in the same study.
Adapted from King, Burley, and Blundell 1994.

Finally, whether hunger ratings and amount of food eaten during a meal in nondeprived
subjects are affected by changes in premeal energy availability can be tested by changing the
number of calories available in the previous meal. If a 100 kcal breakfast is compared with a
500 kcal breakfast, one would expect that the smaller meal would generate stronger hunger
and higher midday energy consumption than would the larger morning meal. In addition, if
the calories withheld from the small meal were provided intravenously, one would expect
that the internal energy sensors would adjust the hunger sensation and quantity of food eaten
later. The small meal (Rest trial) produced greater hunger ratings (figure 6.4a) and lower
132 Advanced Exercise Endocrinology

satiation scores (figure 6.4b) than did the large breakfast (Sed trial). However, total parenteral
(intravenous) nutrition (TPN) administration of 365 cal after the small meal (Rest-TPN trial)
did not attenuate increased hunger (Borer et al., 2009). Similarly, fullness ratings were lower
after the small breakfast and unaffected by TPN, and higher during exercise compared with
fullness ratings in the sedentary trial with the large breakfast, regardless of intravenous
nutrient restoration (figure 6.4b). Under all three conditions, the hunger (figure 6.4a) and
fullness (figure 6.4b) ratings were identical at 13 h, as was the amount of food consumed
during this ad libitum midday meal (figure 6.4c). These results confirm that sensations of
hunger and satiation respond to the amount of nutrients transiting through the mouth and
GI tract but are unresponsive to internal signals of energy availability. Hunger and fullness
ratings and the amount of food eaten at the next meal are the same regardless of preceding
energy availability and intermeal differences in these appetite sensations. This study also
confirmed the capacity of exercise to acutely suppress hunger (figure 6.4a) and increase
satiation (figure 6.4b) in postmenopausal women as it did in young men (King et al., 1994).
Both psychophysical sensations remained refractory to intravenous nutrients, and variations
in energy availability had no effect on the amount of food eaten at the next meal (figure 6.4c).

100
Meal Meal
90 Rest
Rest-TPN
80
Sed
Hunger ratings (percent)

Infusion
70 EX
EX-TPN
60
50
40
30
20
10 Exercise

0
6:00

7:00

8:00

9:00

10:00

11:00

12:00

13:00

14:00

15:00

16:00

17:00

a Time of day
(continued)

Figure 6.4 Modulation of (a) hunger and (b) satiation ratings by orogastric transit of food and exercise and absence
of caloric tracking of internal premeal energyE4517/Borer/Fig6.4a/401585/alw/R2
balance. (a) Hunger was increased in response to a small morning
meal (Rest, open circles) and decreased during 2 h of exercise at 46% of maximal effort (EX, open rectangles) that
expended 565 kcal after a large breakfast. In both trials, hunger was insensitive to an intravenous 68% calorie
replacement in both reduced-energy groups (Rest-TPN, solid circles; EX-TPN, solid rectangles) relative to the
sedentary control in which subjects ate a large meal (SED, solid triangles). (b) Converse effects were observed on
satiation ratings. (c) Despite the negative energy balance at the time of the midday meal (top) that was produced by
the small breakfast (Rest) and exercise (EX), the amount of food eaten and available in unlimited amount at lunch
was the same in all five conditions (center). The meal did not erase residual differences in energy balance (bottom).
Reprinted, by permission, from K.T. Borer, E. Wuorinen, and K. Ku, 2009, “Appetite responds to changes in meal content while ghrelin, leptin,
and insulin track changes in energy availability,” Journal of Clinical Endocrinology and Metabolism 94: 2290-2298.
100
Meal Meal
90
Infusion
Fullness ratings (percent) 80
70
Exercise
60
50
40
Rest
30 Rest-TPN
20 Sed
EX
10 EX-TPN
0
6:00

7:00

8:00

9:00

10:00

11:00

12:00

13:00

14:00

15:00

16:00

17:00
b Time of day

Energy availability before the mid-day meal


500 E4517/Borer/Fig6.4b/401586/alw/R2
Energy availability

400
300
(Kcal)

200
100
0
Rest Rest-TPN Sed EX EX-TPN
−100
−200
Food eaten during the mid-day meal
900
800
700
600
Kcal eaten

500
400
300
200
100
0
Rest Rest-TPN Sed EX EX-TPN

Energy availability after the mid-day meal


1,250
Energy availability

1,000

750
(Kcal)

500

250

0
c Rest Rest-TPN Sed EX EX-TPN

Figure 6.4 (continued)

E4517/Borer/Fig6.4c/401587,588,644/alw/R2
133
134 Advanced Exercise Endocrinology

DESCRIPTORS AND CONTROLS OF FEEDING


BEHAVIOR
Human eating is governed by biological, psychological, and social factors, and in some
instances these influences interact. Some dissonance in research on feeding reflects incon-
sistent use of terminology that describes psychophysical sensations associated with food
seeking and meal eating, and additional dissonance is caused by differences of opinion
regarding how feeding behavior is controlled.

Inconsistent Use of Descriptors of Feeding Behavior


The term hunger indicates a biological drive to seek and ingest food when under negative
energy balance; the drive generally is little affected by food palatability. The term appetite
refers to the motivation to consume food based on its attractive odor, taste, and visual pre-
sentation. This latter type of eating often is described as hedonic and is based on sensory
pleasure associated with a food stimulus. In obese humans and animals, such intake is
exaggerated when the sensory properties of food are positive and is greatly suppressed
when food is unpalatable. This feeding response is termed finickiness. Alliesthesia, the
term coined by Michel Cabanac, indicates increased desirability or sensory salience of food
resulting from energy restriction (Cabanac & Duclos, 1973). This type of feeding therefore
combines a physiological need with hedonic appreciation. Humans also eat palatable food
when they experience neither hunger nor appetite because consumption can be facilitated
by social context, boredom, advertising, comfort seeking, or other social and psychological
variables (Levitsky, 2005; Speakman et al., 2011).
The terms that describe the termination of a meal and suppression of meal taking during
an intermeal interval have also been used in a confusing fashion. The condition that termi-
nates a meal should be correctly identified as satiation, whereas the voluntary reduction
of feeding between meals that also influences the quantity of food eaten during a meal
is usually called satiety. Because the conditions are controlled by different mechanisms,
making proper distinctions is important.

Mechanisms Controlling Psychophysical and Ingestive Aspects


of Eating
Hunger is controlled by different physiological, neural, and hormonal mechanisms than is
hedonic eating. The brain areas responsible for hunger (i.e., a motivation or drive to seek
and ingest food) are in the ventral tegmental area and the shell of the NAc, one of the basal
forebrain ganglia (Berridge, 2009; figure 6.5). Increased concentrations of either basal
insulin or leptin associated with obesity, or acute increases in the two hormones associated
with eating, suppress both the motivation to eat and the motivation for physical activity by
acting on these neural substrates (Figlewicz & Benoit, 2009; Fulton et al., 2000). These
neural substrates therefore represent the mechanism that energizes locomotor seeking of
food and food ingestion. The reduction in circulating leptin and insulin during acute negative
energy balance and longer-term energy deprivation leading to losses of body fat and body
mass removes the inhibitory influences of the two hormones from these neural substrates
underlying hunger drive and the motivation to locomote. Leptin administration blocks the
hyperactivity induced by semistarvation in the rat (Exner et al., 2000) or by anorexia ner-
vosa (Ehrlich et al., 2009). In addition, acyl ghrelin (the octanoylated version of ghrelin),
which is synthesized in the submucosa of the gastric fundus and released in higher amounts
during negative energy balance or after loss of body fat or mass, acts on these same brain
Hormonal Control of Energy Expenditure and Intake 135

Orbitofrontal Anterior
cingulate

Insular Thalamus

Ventromedial
Prefrontal
Nucleus Mesolimbic Parabrachial
accumbens dopamine VTA nucleus
Opioid “liking” and Ventral
“wanting” zones in pallidum
NAc shell

Hedonic hotspot Hedonic hotspots


“Liking”
Increase “Disliking”
Decrease

Hedonic coldspot
“Liking”
Decrease

“Wanting” hotspot Sagittal


Eating
Increase

Figure 6.5 Neural substrates mediating hunger and hedonic eating. Hunger is a response of mesolimbic DA
and its projection to the NAc shell to negative energy balance and loss of body mass and fat. Hormonal signals
E4517/Borer/fig.6.5/401645/TB/R2-alw
such as reduced leptin (and insulin) concentration act on NAc through the mesolimbic DA pathway and opiate
release to increase the incentive salience of food stimuli and motivation for spontaneous physical activity. Hedonic
appetite is a response to palatable food but not to declines in energy balance or body fat mass. It is mediated by a
somewhat broader but similar neural substrate and is triggered by release of opiates, DA, and endocannabinoids.
Reprinted from Physiology & Behavior, Vol. 97, K.C. Berridge, “Liking” and “wanting” food rewards: Brain substrates and roles in eating
disorders,” pgs. 537-550, copyright 2009, with permission from Elsevier.

substrates to increase the incentive salience of food (Chen et al., 2009; Fry & Ferguson,
2010). Both the decreased leptin and insulin concentrations and increased acylated ghrelin
concentration increase endogenous opiate release and activate mesolimbic dopamine (DA)
neurons to act on the nucleus accumbens (NAc) shell (Berridge, 2009; Quarta et al., 2009).
Thus, declines in basal leptin and insulin associated with weight loss increase in parallel the
hunger drive, motivation for action, and levels of spontaneous locomotor activity (Figlewicz
& Benoit, 2008), and higher ghrelin concentration increases GI motility (Chen et al., 2009).
Leptin also affects feeding through neuropeptidergic, monoaminergic, and endocannabi-
noid brain circuits in the ARC nucleus (figure 6.6). It inhibits orexigenic (appetite stimulat-
ing) neurons that release NPY and AgRP and stimulates neurons secreting anorexigenic
peptides (Gautron & Elmquist, 2011; Schwartz et al., 2000). The pro-opiomelanocortin
product α-MSH and CART are anorexigenic, and both activate neurons in PVN and inhibit
orexigenic circuits in the LHA. The LHA and the adjacent perifornical hypothalamus contain
orexigenic neurons that release orexin-A and melanin-concentrating hormone (MCH). The
PVN also suppresses feeding through the anorexigenic peptides TRH and CRH.
136 Advanced Exercise Endocrinology

PVN LHA/PFA

TRH + +
Orexins Second-order
* OXY – #
– MCH neurons
CRH

NPY/AgRP # POMC/CART * First-order


– + + neurons

ARC
Catabolic/Anabolic
response
Adipose Leptin
tissue Insulin
Ghrelin + NTS
(Adiposity signals)
GI tract CCK
(Satiety signals) GLP-1 Vagus
PYY * Anorexigenic peptides
# Orexigenic peptides

Figure 6.6 Brain neuronal circuits in the control of feeding. The adiposity signals leptin and ghrelin are pos-
tulated to reach the ARC throughBorer/E4517/Fig. 6.6/401646/Tim
circulation, where they B/R2-alw
inhibit orexigenic hormones and stimulate anorexigenic
hormones. GI fullness signals from gut peptide hormones reach the hindbrain NTS through the vagus nerve. These
two lines of information are integrated in the ARC and further relayed to the anorexigenic PVN and the orexigenic
LHA and PFA areas.
Reprinted, by permission, from Nutrition Metabolism and Cardiovascular Diseases, Vol. 18, E. Valassi, M. Scacchi, and F. Cavagnini,
“Neuroendocrine control of food intake,” pgs. 158-168, copyright 2008, with permission from Elsevier.

Hunger sensation also has a basis in rhythmic gastric contractions. This connection
was first demonstrated a century ago by Cannon and Washburn (1912) using polygraphic
recordings from inflated intragastric balloons swallowed by volunteers who underwent
episodes of food restriction and feeding. More recently the association between gastric
pyloric and antral contractile activity and the human propensity to eat was confirmed by
Seimon and colleagues (2010). Several variants of the gastric hormone ghrelin increase GI
motility when the stomach is empty, and their secretion and actions are reduced by meal
ingestion (Chen et al., 2009).
Obese individuals do not abstain from feeding despite having high basal plasma leptin
and insulin concentrations. This obvious failure of the putative negative feedback of the two
hormones over hunger and eating is attributed to the resistance of the brain and peripheral
tissues to action of the two hormones (Frederich et al., 1995; Guyenet & Schwartz, 2012;
Schwartz & Niswender, 2004). Instead, finicky feeding in obese individuals is strongly
stimulated by palatability and suppressed by negative sensory attributes of food. As shown
in figure 6.5, hedonic hotspots are found in the areas of the ventral pallidum that, when
stimulated, increase acceptance of palatable food and partially overlap the areas that activate
hunger. Neurochemicals that activate hedonic feeding include DA, opiates, and endocan-
nabinoids (Berridge, 2009). However, unlike structures that are responsible for the activation
of hunger and locomotion, hedonic hotspots are unresponsive to negative energy balance or
losses of body fat or body mass. The brain substrates that control hedonic responses to food
ensure the opportunistic and nonhomeostatic consumption of easily accessible, energy-rich,
palatable food even by well-fed animals and humans. This allows expansion of adipose
tissue mass and contributes to obesity.
Hormonal Control of Energy Expenditure and Intake 137

The GI hormones CCK, GLP-1, GIP, and PYY activate receptors on the PS vagus
nerve and may have limited direct access to the hypothalamus through the median emi-
nence that is devoid of blood–brain barrier (BBB) (Berthoud, 2008). Some of the actions
of the GI hormones contribute to postingestive satiation; this effect is partly the result of
GI hormonal suppression of gastric emptying. The vagus transmits the information about
GI nutrient processing to the NTS in the MO. This information then travels to the ARC
hypothalamic nucleus to be integrated with adiposity signals from leptin and insulin as
well as with hypothalamic and suprahypothalamic circuits that are involved in the control
of feeding (figure 6.6).

ROLE OF PHYSICAL ACTIVITY IN THE REGULATION OF


BODY MASS
According to the homeostatic view of energy regulation (figure 6.1), increases in body fat
level counteract adiposity through negative feedback exerted by high circulating leptin
concentrations. Researchers postulate that this occurs because of stimulation by leptin of
increased physical activity and metabolic means of energy expenditure. Conversely, decreases
in body fat levels are assumed to weaken this negative feedback through reduced physical
activity and metabolic rate (Guyenet & Schwartz, 2012; Schwartz et al., 2000). Injections
of leptin into leptin-deficient mice that are obese and profoundly hypoactive supported
this concept by stimulating higher levels of physical activity (Pelleymounter et al., 1995).
How physical activity contributes to the regulation of energy balance or the body mass
setpoint needs to be examined after manipulations that affect energy balance. Cross-sectional
studies show an inverse relationship between spontaneous physical activity and body fat
mass in both animals and humans (figure 6.7). When body fat mass in animals (Borer, 2010)
and humans (Rising et al., 1994; Schulz & Schoeller, 1994) increases, nonbasal energy
expenditure is shown to decline (figure 6.7) to the point where morbidly obese individuals
display almost complete inactivity (Vanhecke et al., 2009). This, however, does not reveal the
causality of the relationship. Reduced
physical activity could facilitate weight 70
gain and obesity, and larger body mass 60 Male
could discourage physical activity
Percent body fat

50 Female
because greater effort and energy cost
40
is needed to move it. Therefore, studies
in which activity levels are studied after 30
manipulation of body mass or body fat- 20
ness are important and instructive. They 10
strongly suggest that body fat content 0
exerts an influence over the motivation 0.00 0.04 0.08 0.12 0.16 0.20
to locomote. Profound hypoactivity Nonbasal energy expenditure of
develops in rodents when obesity is physical activity (MJ · kg−1 · day−1)
induced by high-fat or palatable diets
(Borer et al., 1989; Sclafani & Springer, Figure 6.7 The inverse relationship between spontaneous
1976), lesions of the medial basal hypo- physical activity and body fat.
E4517/Borer/Fig6.7/401589/alw/R1
thalamus (Kennedy & Mitra, 1963b) or Reprinted, by permission, from S.K. Powers and E.T. Howley, 2009,
Exercise physiology: Theory and application to fitness and performance,
the rostral septum and its connections 7th ed. (Boston: McGraw Hill), 401. © The McGraw-Hill Companies; Data
with the hippocampus (Borer et al., from Rising 1994: Schulz and Schoeller 1994.
138 Advanced Exercise Endocrinology

1977, 1979) that increase body fatness, or conditions in which leptin cannot be produced
or its receptors are inactivated (Pelleymounter et al., 1995). However, if hamsters that were
rendered hypoactive by septal lesions are forced to run on a treadmill under external motiva-
tion supplied by electric shock at its base, they display endurance and locomotor power that
is similar to that of their lean counterparts (Borer et al., 1983). Lesions of brain substrates
that result in obesity abolish the inverse correlation between body fat content and levels of
spontaneous locomotion (Borer et al., 1979; figure 6.8). In addition, chemical lesioning of
serotonin neurons in the same rostral septal area where neural damage produces obesity and
hypoactivity reverses the hypoactivity associated with a high-fat diet and restores levels of
spontaneous locomotion of neurologically intact obese hamsters to those of lean animals
(Borer et al., 1989). These data support the conclusions that obesity leads to hypoactivity
by affecting brain regions that are associated with the control of spontaneous running, that
hypoactivity in obese animals can be overcome by neurochemical manipulations of these
brain areas, and that hypoactivity is a motivational defect rather than a motor defect.
Decreases in body fat mass produce effects on spontaneous locomotion that are opposite
to those observed in obesity. Losses of body fat and body mass in rodents are accompanied
by proportional increases in the levels of voluntary locomotion. A striking demonstration
involves maintaining rats on restricted access to food that is insufficient to maintain their
body weight. Access to a running wheel leads to 300% to 500% increases in spontaneous
running as the body mass and body fat of the rats progressively decline. If the experiment
is not terminated, rats virtually run themselves to death from exercise-induced weight loss
and energy depletion (Routtenberg & Kuznesof, 1967). This animal model is paralleled by
human anorexia nervosa, a condition of suppressed food intake, sustained weight loss, and
high motivation for, and involvement in, physical activity (Casper, 2006). Thus, prevailing
evidence indicates that spontaneous physical activity levels are causally related to body
fat in an inverse and nonhomeostatic fashion as long as the brain regions involved in the
regulation of body fat remain intact (see figure 6.8).

100
Mean activity (RPD × 103)

50 50
30
20 40
10 30
5 20
3
2 10

0 0
0 5 10 15 20 25 30 35 100 125 150 175 200
Days after surgery Body weight (g)

Figure 6.8 The inverse relationship between spontaneous physical activity and body fat. Surgical disconnec-
E4517/Borer/Fig6.8/448746/alw/R1
tion between the septum and hippocampus causes obesity and lower voluntary running levels in lesioned (solid
symbols) golden hamsters than in nonlesioned (open symbols) golden hamsters (left). The inverse relationship
between running activity and body mass in nonlesioned animals is abolished in lesioned animals (right). RPD =
disc revolutions per day.
Reprinted, by permission, from K.T. Borer et al., 1979, “Contribution of growth, fatness, and activity to weight disturbance after
septohypothalamic cuts in adult hamsters,” Journal of Comparative Psychology 93: 907-913.
Hormonal Control of Energy Expenditure and Intake 139

REGULATION OF BODY MASS THROUGH


HOMEOSTATIC VERSUS NONHOMEOSTATIC
CONTROLS
Empirical evidence presented in previous sections precludes acceptance of the prevailing
homeostatic model of body mass regulation based on lipostatic properties of the hormone
leptin as a complete explanation of our feeding mechanism. This model does not explain
dietary obesity or the inability of leptin administration to suppress feeding and reduce body
fat mass in obese humans whose ability to secrete leptin is not compromised. In addition,
this model incorrectly interprets the role of physical activity in body mass regulation and
ignores the role of body components other than body fat in regulating body mass. Other
models of body mass regulation are similarly deductively constructed using features of vari-
ous engineering control systems (Speakman et al., 2011) and to a large extent do not account
for the empirical data. For instance, a settling point model attributes body mass stability to
passive misalignment between unregulated food intake or unregulated energy expenditure.
This model postulates that body mass passively settles at a plateau that represents a balance
between input and output. This model does not acknowledge that body mass losses, whether
deliberate or experimentally induced, lead to increased hunger and increased hormonal and
metabolic efficiency that favors weight regain (Anderson et al., 2001; Keys et al., 1950).
Yet another model, the deductive dual intervention point model, posits that upper and lower
body mass boundaries are physiologically defended whereas the intermediate range of body
mass is subject to environmental nonhomeostatic influences (see Speakman et al., 2011). As
discussed earlier, no strong evidence is available for behavioral and physiological defense
against the upward excursion of body mass.
The integrative view in this chapter is based on inductive conclusions that stem from
robust experimental data. The key data used in formulating this view show the following:
all components of body mass in adult mammals, not only fat mass, are strongly defended
against declines and are allowed to increase when excess energy is available through oppor-
tunistic overnutrition and increased metabolic efficiency; hormonal and neural mechanisms
account for vigorous defenses against body mass declines and ineffective defenses against
body mass accrual; and the blend of opportunistic feeding and homeostatic defense of
body mass against losses, which are now maladaptive in our contemporary environment
of sedentary lifestyle and overabundance of food, is likely to have evolved in response to
the unpredictable cycles of famine and feasting in human Paleolithic ancestors. Using this
somewhat speculative evolutionary framework for the explanation of body mass regulation,
we interpret why periodic losses of body mass through dietary restriction or restraint may be
necessary in order to establish levels of spontaneous activity required for the maintenance
of healthy lean body mass.

Evolutionary Insights Regarding Opportunistic Features of Food


Intake
Meal-to-meal eating in many animals is likely based on two principles: variable interest
in food depending on whether the digestive system is engaged in the processing of food,
and the opportunity for and access to suitable or attractive palatable food. Two examples
from poikilotherm animals illustrate the prevalence of opportunistic eating. An octopus is a
140 Advanced Exercise Endocrinology

mollusk that has an indeterminate ability to grow depending on environmental conditions.


A blowfly is a poikilotherm that has a finite body size in its last instar. The quantity of prey
an octopus captures is proportional to prey availability, although at all prey densities the
octopus does not capture and eat all available crabs. If metabolism and food processing are
accelerated in warmer water, the amount of prey an octopus captures and eats is increased
(Borer, 1971). A blowfly with a full crop ignores the sources of food, but when its crop is
empty it ingests sugar solutions in proportion to their concentration and sweetness. Thus,
rather than eating for calories by ingesting more of a less-concentrated solution, a blowfly,
like the octopus, opportunistically maximizes intake of the most stimulating food. Satia-
tion in the blowfly is signaled by a negative feedback transmitted by the recurrent nerve
from the crop to the brain. When this nerve is sectioned, the blowfly overeats. Given a
sufficiently concentrated sugar solution, it can rupture its crop (Dethier et al., 1965).
Human meal-to-meal eating in a nondeprived state does not substantially differ from
the feeding pattern of a blowfly. As discussed earlier, it is facilitated by the appealing taste,
odor, and presentation of food, social facilitation, easy access to food, and portion size.
Hormonal and mechanical signals associated with food ingestion and processing exert a
satiating influence (figure 6.9b and c) until food digestion is completed. As the GI tract
empties, the sensation of hunger increases (figure 6.4a) and satiation decreases (figure 6.4b)
until the opportunity arises to eat a volume of food that most likely corresponds to stomach
capacity (center panel of figure 6.4c; Kendall et al., 1991). This interpretation is supported
by the effectiveness of surgical restriction of stomach capacity in suppressing feeding and
inducing weight loss (Vetter et al., 2011). Thus, in contrast to the homeostatic interpretation
of human feeding (Guyenet & Schwartz, 2012; Schwartz et al., 2000), meal-to-meal eating
in humans appears to be opportunistic, nonhomeostatic, and dependent on GI processing of
food. Meal size in a nondeprived state is not affected by previous energy balance regardless

2,000 Meal Meal


Rest
1,800 Rest-TPN
Sed
1,600 EX
Plasma ghrelin (pg/mI)

EX-TPN
1,400

1,200

1,000

800

Infusion
600
Exercise
400
6:00 7:00 8:00 9:00 10:00 11:00 12:00 13:00 14:00 15:00 16:00 17:00
a Time of day
(continued)

Figure 6.9 Changes in (a) plasma total ghrelin and (b) leptin, and (c) GIP concentrations in response to differ-
ences in meal size, exercise energy expenditure, and TPN.
Reprinted, by permission, from K.T. Borer, E. Wuorinen, and K. Ku, 2009, “Appetite responds to changes in meal content while ghrelin, leptin,
E4517/Borer/Fig6.9a/448747/alw/R2
and insulin track changes in energy availability,” Journal of Clinical Endocrinology and Metabolism 94: 2290-2298.
Hormonal Control of Energy Expenditure and Intake 141

Rest
30.00
Meal Meal Rest-TPN
Sed
EX
25.00 EX-TPN
Infusion
Plasma leptin (ng/mI)

20.00

15.00

10.00
Exercise

5.00
6:00 7:00 8:00 9:00 10:00 11:00 12:00 13:00 14:00 15:00 16:00 17:00
b Time of day

600 E4517/Borer/Fig6.9b/448748/alw/R1
Infusion
Rest
500 Rest-TPN
Sed
Meal Meal EX-Saline
EX-TPN
Plasma GIP (pg/mI)

400
Exercise
300

200

100

0
6:00 7:00 8:00 9:00 10:00 11:00 12:00 13:00 14:00 15:00 16:00 17:00
c Time of day

Figure 6.9 (continued)

of whether this balance is manipulated by withholding or adding parenteral calories or by


E4517/Borer/Fig6.9c/448749/alw/R2
increasing or decreasing the amount of physical activity. Thus, meals eaten during variable
energy availability (figure 6.4c) or meals with varying caloric density (Lissner et al., 1987)
appear to be limited by stomach capacity rather than by the caloric imbalance. Manipulating
premeal energy availability through altering morning meal size, exercise energy expenditure,
or intravenous restoration of calories produces expected changes in plasma concentrations
of leptin (figure 6.9b) and ghrelin (figure 6.9a) (i.e., decreases in the former and increases
in the latter) when energy balance is reduced by small meals or exercise. However, in a non-
deprived state, these endocrine changes have no effect on sensations of hunger (figure 6.4a)
142 Advanced Exercise Endocrinology

or satiation (figure 6.4b), demonstrating a functional disconnect between the homeostatic


reactions of these hormones and psychophysical responses of food consumption (figure
6.4c). On the other hand, a clear correlation exists between the time course of satiation
(figure 6.4b) associated with food processing and the secretory pattern of the gut hormone
GIP released from the duodenum in response to nutrient entry and processing (figure 6.9c).

Evolution of Homeostatic and Nonhomeostatic Control of Food


Intake
An evolutionary perspective derived from the feeding pattern in hunter–gatherer societ-
ies provides a speculative explanation of why humans should have both opportunistic and
homeostatic mechanisms for the intake of food (Chakravarthy & Booth, 2004). Researchers
speculate that our mechanism for regulating body mass evolved during the Late Paleolithic
period (50,000-10,000 BC) when variable availability of foraged or hunted food led to cycles
of famine and weight loss on one hand and feast and overeating on the other. The 10,000 yr
since the development of agriculture, which increased food abundance, was not sufficiently
long to allow the mechanism that regulates body mass to evolve from one specific to a hunt-
ing and gathering lifestyle to one specific to a lifestyle that is predominantly sedentary and
allows easy access to abundant food (Cordain et al., 1998; O’Keefe et al., 2010).
As discussed earlier, evolutionary adaptations to periods of fast included hormonal
facilitation of hunger and motivation for physical hunting or foraging for food. Additional
autonomic adaptations to weight loss and negative energy balance included reduced S activa-
tion of metabolism to preserve energy stores and body structure and correspondingly lower
stimulation of metabolism by thyroid hormones. Parallel decreases in leptin concentration
and body fat mass also served to reduce activation by leptin of TRH and thyroid hormones
and thus further contributed to reduction of metabolic energy expenditure. Increased pul-
satile GH secretion protected oxidation of structural body proteins for energy and diverted
metabolism toward oxidation of lipids. Increased lipid utilization during exercise in an
energy-depleted state, particularly by Type I muscle fibers, served to preserve muscle gly-
cogen and produce greater ATP yield per unit mass of storage fuel. Other hormonal changes
during body mass loss and negative energy balance, such as increased insulin sensitivity,
contributed to what is construed to be part of the “thrifty genes” adaptation (Eaton & Eaton,
2003; Neel, 1962). An adaptive advantage in the affinity of muscle HK over hepatic GK
for glucose phosphorylation ensured that muscle glycogen was restored more rapidly than
was liver glycogen after a famine (Terjung et al., 1974). These adaptations minimized uti-
lization of stored fuels during exercise in an energy-depleted state and facilitated efficient
accumulation of storage fuels when the famine was broken with a feast.
The acquisition of food sources resulted in a feast (Chakravarthy & Booth, 2004; figure
6.10) during which opportunistic overeating allowed repletion of depleted energy stores.
Glycogen and fat depots were refilled, basal leptin and insulin concentrations were increased,
and the motivation for spontaneous physical activity decreased. An increase in plasma insulin
to concentrations that are too low to cause hypoglycemia has been experimentally shown
to decrease physical activity in rats (see Borer, 2010). The prevailing autonomic influence
involved the activation of PS control over digestion and secretion of insulin and gut hormones.
Our contemporary lifestyle causes a disruption in this Paleolithic pattern of alternating
between famine and feast. Adults in developed societies have unabated opportunity to feast
on abundant, energy-rich food and therefore experience few opportunities for energy deple-
tion and weight loss. This interrupts the transition from feast to famine that is necessary for
increased motivation to physically seek food and experience increased insulin sensitivity
and other beneficial enzymatic changes (Chakravarthy & Booth, 2004).
Hormonal Control of Energy Expenditure and Intake 143

Thrifty storage
(Replenish skeletal
muscle glucose and TG;

it y
tiv
more efficient storage of

ac
(More thrifty storage = more

in
excess glucose and TG
likely to survive through next

al
in adipose tissue)
ic
ys famine/activity phase until
ph
next feast)
ive
at
l
Re

Feast Famine and activity


(Intake glucose and fat) (Essentially simultaneous)
(Decrease glycogen and TG stores)
Successful
physical activity
(hunt or gather)
Contracting skeletal
muscle increases
GLUT-4 and AMPK

Figure 6.10 The feast–famineBorer/E4517/Fig.


and physical activity cycle postulated toB/R2
6.10/448750/Tim have evolved more than 10,000 yr ago.
During famine, neural and metabolic adaptations facilitated hunger, physical seeking of food, and autonomic and
enzymatic adaptations (including increased insulin sensitivity) for efficient nutrient storage upon the onset of feast.
Nutrient overconsumption during the feast induced inactivity and efficient nutrient store restoration.
Reprinted, by permission, from M.V. Chakravarthy and R.W. Booth, 2004, “Eating, exercise, and “thrifty” genotypes: connecting the dots
toward an evolutionary understanding of modern chronic diseases,” Journal of Applied Physiology 96: 3-10.

Mechanism Responsible for Regulation of All Three


Compartments of Body Mass
Reduction of body mass entails changes in the mass of all three body compartments: lean
tissue, fat depot, and bone. When 1 kg of body fat is gained or lost in humans, 16.5 g of bone
mineral is gained or lost (Jensen et al., 1994). Changes in body fat level are also accompanied
by changes in lean body mass. This insight is supported by findings that the hormone leptin
regulates bone and adipose tissue setpoints in adult mammals by inhibiting the bioactivity
of the bone hormone osteocalcin (OCN) through activation by leptin of S outflow (Karsenty
& Oury, 2010). Although this new concept of energy regulation has omitted measurements
and consideration of the role of physical activity, it allows an integrative reinterpretation of
the autonomic and hormonal roles in the regulation of adult body mass.
Researchers have recognized for more than 3 decades that regulation of body fat mass
involves an interaction between S and PS divisions of the ANS. Obesity-producing damage
to the VMH and ARC causes an acute period of insulin oversecretion. This hyperinsulinemia
is caused by hyperactivity of PS division of the ANS uncoupled from the restraining action
of the S division after damage to brain substrates initiating S outflow or containing neurons
of transit from these centers. Both the hyperinsulinemia (Berthoud & Jeanrenaud, 1979)
and hypoactivity (Eng et al., 1978) associated with VMH and ARC lesions are prevented
or abolished by subdiaphragmatic vagotomy, implicating the hyperactivity of PS in these
phenomena. These findings, along with historical use of sympathomimetic amphetamines
and E, highlight the probable involvement of S activation in acute exercise-induced sup-
pression of hunger. The role of the ANS in the regulation of the body mass setpoint and its
connection with physical activity is an area of inquiry that deserves additional attention.
144 Advanced Exercise Endocrinology

Connections between the role of leptin and the ANS in the regulation of the three com-
ponents of body mass were made possible by the development of powerful research tools in
the 21st century, including gene manipulation, selective neuronal chemical lesioning, and
anterograde and retrograde neuronal tracings. These technological advancements revealed
a connection between the hormone leptin, S nerve action, and regulation of bone mass.
Increased S tone suppresses the bioactivity of the bone hormone OCN by activating an Esp
gene that increases γ-carboxylation of the hormone (Ducy et al., 2000). Upregulation of this
gene reduces osteoblast numbers and blocks increases in bone mineralization and size. The
effect is dependent on beta AdRs on the osteoblasts. When beta AdRs on osteoblasts are
genetically ablated, a phenotype that is high in bone, body mass, and body fat is observed.
This phenotype is similar to the phenotypes of animals that have lesions of the medial basal
hypothalamus or the phenotypes of genetic models that lack the ability to produce leptin
(in ob/ob mice) or its receptors (in db/db mice). Most of these phenotypes are characterized
by obesity, inactivity, and low levels of metabolic energy expenditure. This research has
extended leptin actions from adipose tissue to the skeletal component of body mass.
This line of research also demonstrates that regulation of bone mass requires the integrity
of the VMH hypothalamus, whereas the regulation of body fat, blood glucose, and feed-
ing depends on the integrity of the ARC hypothalamic nucleus (Karsenty & Oury, 2010).
Because leptin was implicated in alterations in body fat in some obese phenotypes through
its activation of S outflow, the neural site of its inhibitory actions was of interest. The
hypothesis that circulating leptin directly acts on the medial hypothalamus by reaching the
ARC through median eminence (Guyenet & Schwartz, 2012; Schwartz et al., 2000) proved
incorrect because mice with deletions of LepRb leptin receptors on POMC neurons in the
medial basal hypothalamus were not obese yet they responded to intracerebroventricular
leptin administration (Karsenty & Oury, 2010). Instead, leptin was found to act on the
raphe nuclei in the MO that send separate serotonergic projections to the ARC and VMH
(figure 6.11). The separation of leptin actions on bone and body fat was achieved by chemi-
cal lesioning of ARC neurons with monosodium glutamate and of neurons in the VMH
with gold thioglucose. Selective chemical lesioning of the ARC nucleus caused obesity
and hypoactivity, whereas lesions confined to the VMH produced increases in bone mass.
Leptin achieves these effects by inhibiting serotonin neurons in the hindbrain raphe nuclei
that, in turn, inhibit S projections through the ARC and VMH to brain regions that control,
respectively, food intake via Htr1a and Htr2b receptors and bone osteoblast survival via
Htr2c receptors (figure 6.11).
Based on these findings, when body fat level and energy availability increase, plasma
leptin concentration increases and blocks serotonergic projections from the raphe nuclei to the
mediobasal hypothalamus from promoting body mass gain. Instead, leptin activates S outflow
in the ARC that leads to suppression of feeding and increases in lipolysis, thermogenesis,
and physical activity. Leptin suppresses bone accretion by facilitating S outflow through
the VMH. When the three components of body mass decline based on energy expenditure
from exercise or dietary restriction, plasma leptin concentration and S tone decrease. This
permits greater PS and serotonergic activation of anabolic processes, including stimula-
tion by bioactive OCN of secretion of adipokine adiponectin and insulin and of peripheral
tissue insulin sensitivity. Thus, leptin appears to be the link between body energy status
and neural substrates that are responsible for the activation of S outflow, thermogenesis,
and physical activity that restrains body mass gain on one hand and anabolic PS actions
that favor feeding and body mass accretion on the other.
Hormonal Control of Energy Expenditure and Intake 145

Brainstem

Serotonin

Hypothalamus

Raphe VMH
nuclei
Leptin Htr2c

ARC
Htr1a
Htr2b

Appetite Bone mass


accrual

Figure 6.11 Neural circuits through which leptin inhibits feeding and bone mass. Increased leptin concentration
E4517/Borer/fig.6.11/448751/TB/R3-kh
inhibits serotonergic projection from the raphe nuclei in the hindbrain to the VMH and ARC in the hypothalamus.
These projections ordinarily facilitate feeding and bone accrual by, respectively, preventing S action on feeding
circuits via Htr1a and Htr2b receptors in the ARC and by preventing S inhibition of osteoblast survival via Htr2c
receptors in the VMH. Leptin abrogates this serotonergic inhibition and facilitates inhibitory S effects.
Reprinted, by permission, from G. Karsenty and F. Oury, “The central regulation of bone mass, the first link between bone remodeling and
energy metabolism,” Journal of Clinical Endocrinology and Metabolism 95: 4795-4801.

BODY MASS REGULATORY MECHANISM VERSUS


OBESITY AND WEIGHT LOSS
The nonhomeostatic control of feeding and the operation of hedonic neural substrates that
facilitate food intake despite supranormal body fat accumulation instruct us that the prereq-
uisites for maintaining healthy body mass level are restraint in eating food, high levels of
physical activity, and periodic episodes of weight loss. The obstacles to using this strategy
for facilitating weight loss once obesity has developed include a powerful hunger drive,
reduced motivation to be physically active, and physiological adaptations that counteract
body mass loss and make weight regain highly efficient. A meta-analysis of the effective-
ness of different approaches to weight reduction in obese humans shows both a gradation
of effects depending on the method of energy restriction and clear evidence for a defense
of predeprivation weight (figure 6.12; Franz et al., 2007).
To optimally engage our presumed Paleolithic mechanism of regulating body mass, we
would need an exercise energy expenditure of about 500 kcal/d (Eaton & Eaton, 2003), higher
muscle mass relative to adipose tissue at sex-appropriate ratios of 50% muscle:10% fat in men
and 45% muscle:15% fat in women, and a BMI of about 21 kg/m2 (Eaton et al., 2009). This
would require a rigorous regime of endurance and strength training because such physique
is approached only in professional athletes. A change in muscle:fat ratio would increase the
proportion of IRs on muscle relative to adipose tissue and thus significantly increase glucose
tolerance and insulin action. We also may need to engage in cycles of fasting and feasting
and periodic episodes of body mass reduction to optimally activate insulin sensitivity and
the operation of our thrifty metabolic proteome (figure 6.10). A drastic change in our diet
would also be required (Eaton, 2006). The Paleolithic diet consisted of about 35% fat, 35%
146 Advanced Exercise Endocrinology

2
0
−2
−4
Weight loss (kg)

−6
−8
−10 Exercise alone
Diet and exercise
−12 Diet alone
−14 Meal replacements
Very-low-energy-diet
−16 Orlistat
Sibutramine
−18 Advice alone
−20
6 mo 12 mo 24 mo 36 mo 48 mo

Figure 6.12 A meta-analysis summary of the magnitude of weight loss as a function of different methods of
energy reduction.
E4517/Borer/Fig6.12/448752/alw/R3
Reprinted, by permission, from M.J. Franz et al., 2007, “Weight-loss outcomes: A systematic review and meta-analysis of weight-loss clinical
trials with a minimum 1-Year follow-up,” American Dietetic Association 107: 1755-1767.

carbohydrate, and 30% protein. Polyunsaturated fats predominated, 50% of carbohydrate


was derived from uncultivated plants and contained only 2 to 3% sugars from honey, and
almost no grains and dairy products were included. Fiber consumption was about 100 g/d,
and vitamin intake was between 2 to 8 times higher than vitamin intake in the contempo-
rary diet. Clearly, attainment of such a goal would require a drastic re-engineering of our
current lifestyle. The optimal way to approximate this goal would possibly be to prevent
the development of obesity by restraining food intake and significantly increasing levels
of physical activity. Absent this option, an obese person could reduce body mass through
exercise and dieting and then successfully maintain reduced body mass by restricting food
intake to about 1400 kcal/d and by spending about 1 h in physical activity each day (Epstein
& Wing, 1980; McGuire et al., 1998; Wing & Phelan, 2005). Success of this strategy has
been documented by the research by National Weight Control Registry, spearheaded by
Dr. Rena Wing. Considerable obstacles to the implementation of this strategy are the
disparity between the hedonic allure and ease of consuming palatable, energy-dense, and
easily accessible food and the amount of physical effort required to expend the calories so
obtained and the challenge of incorporating sufficient levels of physical activity into our
largely sedentary contemporary lifestyle.
147
148 Advanced Exercise Endocrinology

SUMMARY
This chapter examines the mechanisms responsible for the defense of body mass against
losses and for the accrual of excess body mass that leads to obesity. This chapter first
examines the prevailing concepts of energy equation and homeostatic regulation of body
fat through the adiposity hormone leptin. These concepts are inadequate to explain the
development of dietary obesity in the presence of high leptin concentrations and the failure
of leptin administration to suppress feeding and body fat mass in obese humans. Next, this
chapter presents evidence that meal-to-meal feeding in the nondeprived state in humans is
not guided by sensing of caloric balance or adiposity hormones but is instead opportunistic
and guided by the sensory properties of food and environmental factors and is acutely sup-
pressed by physical activity.
This chapter defines the terminology to accurately describe aspects of feeding and dis-
cusses the role of neural substrates and systemic and gut hormones in the control of hunger
and hedonic feeding. An erroneous interpretation of the role of physical activity in the
regulation of body mass is corrected to show that physical activity changes with body mass
in an inverse and nonhomeostatic fashion. This is explained in terms of active food seek-
ing which, with hunger, increases at times of negative energy balance and body mass loss.
The blend of nonhomeostatic feeding that facilitates obesity and increased motivation for
physical activity associated with hunger is interpreted as evolutionary adaptations based on
our genetic heritage from the food-seeking hunter–gatherer past. These adaptations include
more efficient utilization of metabolic fuels during fasting and opportunistic feeding when
food becomes available. This chapter also lists the required changes in contemporary human
lifestyle that would allow better utilization of our genetic heritage in order to achieve and
maintain a healthy body mass size and function.
H APTER

7
C

Exercise and
Reproductive
Hormones
R
eproductive hormones interact with exercise in at least three important ways. First,
hormones that control sexual maturation initiate and maintain sex differences in
musculoskeletal systems, body composition, and metabolic and circulatory processes
that affect physical performance. The first part of this chapter describes how the genes coding
for male and female hormones cause the development of a sexually dimorphic phenotype
and how the secretion and action of sex hormones change throughout the human life span.
Next, this chapter describes the effects of exercise on the secretion of sex hormones and
emphasizes the sex differences in sex hormone secretion during acute exercise and the
mechanism of reproductive hormone suppression during energy-costly endurance exercise.
This discussion complements the information on the secretion of sex hormones in response
to RE (discussed in chapter 5) and the influence of sex hormones on muscle hypertrophy
(discussed in chapter 8). Finally, this chapter examines the effects of sex hormones on physi-
cal activity and performance, in particular the role of fluctuations in female sex hormones
during the fertile period and pregnancy and after menopause.

DEVELOPMENT OF PHENOTYPIC SEXUAL


DIMORPHISM
Sex segregation in many competitive sports is based on the evidence that females gener-
ally differ from males in body size, body composition, musculoskeletal development, and
strength. Sex differences become a problem in sport when genetic or developmental errors
prevent unambiguous identification of an individual’s sex (see “The Challenge of Sex Iden-
tification in Sport”). An overview of the development of sex dimorphism provides useful
insight into the effects of different secretory patterns of sex hormones through the life span
and the ways the actions of sex hormones influence growth and differentiation of male and
female phenotypes.

149
150 Advanced Exercise Endocrinology

Secretion of Sex Hormones Across the Life Span


The Y sex chromosome in the male XY genotype is responsible for the development of
phenotypic dimorphism because in its absence a default female phenotype with immature
external genitalia develops. For normal ovarian development, both X chromosomes in a
genotypic female are required. Development of a male body build is determined by the Y
sex chromosome that is present, along with 22 somatic chromosomes, in paternal sperm
during the union with an X sex chromosome and 22 somatic chromosomes in a female egg.
The Y chromosome contains a male-specific region composing 95% of its length. Of its 27
distinct protein-coding genes identified so far, 12 are expressed ubiquitously throughout
many tissues and 11 are expressed exclusively or predominantly in the testes (Skaletsky
et al., 2003). Sex chromosomes direct the development of gonads, which are derived from
gonadal ridges formed from the coelomic epithelium and underlying mesenchyme during wk
4 of fetal development. By wk 5, they are joined by germ cells originating in the primordial
gut. By wk 6, androgen-producing interstitial cells differentiate to form Leydig cells in the
testes and theca cells in the ovaries. Concurrently, Sertoli cells in the testes and granulosa
cells in the ovaries differentiate and express ARs and the Cyp19 gene. The latter codes
for the P450 cytochrome aromatase enzyme, which is capable of converting androgens to
estrogens. At 8 wk, fetal Leydig cells begin to secrete T in response to activation by ste-
roidogenic acute regulatory protein (Ogilvie et al., 1999). T has anabolic properties that are
responsible for male-specific neural and musculoskeletal growth and androgenic properties
that control the development of male external genitalia. Formation of male external genitalia
includes the development of a penis and a scrotum between wk 9 and 12 of gestation and
an epididymis and vas deferens from the Wolffian duct, one of two primordial ducts serv-
ing kidney function, at about wk 10 of gestation. Seminal vesicles, ejaculatory ducts, and
the urethra are formed at about wk 13 of gestation. At this time, urogenital folds fuse over
the urethra to form the penis. This organ and the associated ducts increase in size during
the second and third trimesters of gestation, and the descent of the testes into the scrotum
is completed around the time of birth. In the female, the Wolffian ducts degenerate and
Müllerian ducts contribute to the formation of fallopian tubes, the uterus, and the vagina.
Vaginal development begins during wk 9 of gestation and ends at wk 20 of gestation. After
wk 10, urogenital folds close, except for the posterior ends, which form the labia and the
vaginal and urethral opening.
Secretion of sex hormones after birth is regulated by the hypothalamic GnRH neurons
that control the secretion of the pituitary gonadotropins, LH, and FSH through a cAMP
signaling pathway. Before birth, human chorionic gonadotropin (hGC) released by the
placenta appears to control fetal Leydig cell steroidogenesis until the end of gestation when
maternal LH becomes effective. GnRH neurons act as a pulse generator because their
secretory rhythm drives the pulses of pituitary gonadotropins. Before puberty (the stage
of human development that includes growth of the gonads, maturation of the genitalia and
reproductive capacity, and sex-specific patterns of statural growth), GnRH pulse frequency
is low and limited to the nocturnal period (figure 7.1). During puberty and when sexual
maturity is attained, the LH pattern changes to a circhoral pattern of approximately one
pulse every hour; this pattern is necessary for the initiation and maintenance of reproductive
function in both sexes. A menstrual rhythm of GnRH release orchestrates the secretion of
sex steroids and prepares the ovum and uterine lining for potential fertilization. A differ-
ence in half-lives of FSH (170 min) and LH (60 min) influences their plasma concentrations,
which vary at different stages of pubertal development and at different stages of the female
menstrual cycle. The circhoral frequency and low amplitude of LH pulses during the 14 d
Exercise and Reproductive Hormones 151

Menopause/Andropause
LH 150

Neonatal
FSH 100
50

Puberty
30
25
20
mlu/ml

15
10
5
0

Neonatal Prepubertal Reproductive Postmenopausal

Figure 7.1 The pattern of gonadotropin secretion throughout the life span. Top panel: GnRH secretion (not shown)
controls ovulatory gonadotropin surges during the female’s reproductive phase between puberty and menopause.
Lower panel: LH is pulsatile immediately after birth and quiescent during childhood and resumes a pulsatile pattern
during the prepubertal period, initially at night only and, upon sexual maturation, throughout the day. This pattern,
except for the preovulatory switch to positive feedback leading to gonadotropin surge, is under negative feedback
by sex steroids in males and E4517/Borer/Fig7.2/401590/alw/pulled-R1
females. Shading indicates the nocturnal period.
Adapted from S.S.C. Yen, 1999, The human menstrual cycle: Neuroendocrine regulation. In Reproductive endocrinology: Physiology,
pathophysiology, and clinical management, edited by S.S.C. Yen, R.B. Jaffe, and R.L. Barbieri (New York: W.B. Saunders), 194.

of the follicular menstrual phase drive follicular growth, synthesis of androgens in theca
cells, and a progressive increase in the follicular release of E2 that precipitates a surge in
LH concentration at ovulation (figure 7.2). Maintenance of fertility in both sexes depends
on the pulsatile pattern of pituitary gonadotropins and is extinguished by their continuous
delivery, a feature that some female contraceptives exploit.
FSH stimulates aromatization of androgens to E2. Before ovulation, E2 concentrations
exert negative feedback over LH secretion. Ovulation is made possible by a change from
negative to positive feedback in E2 action over LH secretion. After ovulation, E2 concentra-
tion decreases, but a corpus luteum (a remnant of the follicle) assumes the endocrine role of
secreting E2 and P4 during the 14 d of the luteal phase of the menstrual cycle. If fertilization
does not occur, the endometrial lining of the uterus, which has hypertrophied and become
vascularized, responds to a decline in both sex steroids by sloughing off in the form of
menstruation, and another round of follicular development is initiated. At about the aver-
age age of 46 yr, female menstrual cyclicity ceases and women enter menopause, which is
characterized by high plasma concentrations of gonadotropins and low ovarian production of
E2 (figure 7.1). The increase in the concentration of gonadotropins at the onset of menopause
represents a decline in negative feedback by ovarian E2 over the release of gonadotropins.
152 Advanced Exercise Endocrinology

In males, circhoral LH pulsatility 200


stimulates the synthesis of androgens in
LH
Leydig cells in the testes. Increased FSH 150
FSH
secretion stimulates spermatogenesis in

IU/L
testicular Sertoli cells. Sexually mature 100
males secrete about 5 to 10 mg of T daily
to produce a total plasma T concentra- 50
tion of between 300 (considered to be the
threshold of hypogonadism) and 1000 0
ng/dl, a concentration that is maintained 1,200
through T negative feedback over LH 900
secretion. The concentration of T is about Inhibin

U/L
600
10 times higher in males than in females,
given that the main source of this hormone 300
is the testis in the male and the adrenal 0
cortex in the female. In contrast to the 1,500 100
abrupt cessation of sex hormone secretion E2 75
in the female, an aging male undergoes a 1,000 P4
pmol/L

pmol/L
50
more gradual decline in the secretion of
500
T during the later years of his life span. 25
1,000 0
Actions of Sex Hormones −12 −8 −4 0 4 8 12 16
Days from LH peak
Sex hormones act predominantly on intra-
cellular receptors. The hormone–receptor E4517/Borer/Fig7.2/401647/alw/pulled-R2
Figure 7.2 Endocrine parameters in the control of a
complex binds to the specific HRE located female’s menstrual cycle. The arrow shows estimated
near a gene promoter (figures 1.19 and onset of ovulation, and the open box shows the time
7.5). Sex steroids upregulate a variety of period of implantation of the ovum.
proteins, including those involved in the Reprinted from S.S.C. Yen et al., 1974, Causal relationship between
the hormonal variables in the menstrual cycle. In Biorhythms and
development of secondary sexual charac- human reproduction, edited by M. Ferin et al. (New York: J Wiley &
teristics, phenotypic dimorphism during Sons), 220.
the pubertal growth spurt, and some sex
differences in metabolism. Sex hormones
can also act rapidly through nongenomic
signaling (see figure 7.5).

Genomic and Nongenomic Actions of Sex Hormones


T is a 19-carbon steroid, 95% of which is produced by the Leydig cells of the testes and, to
a lesser extent, by the zona reticularis of the adrenal cortex and by peripheral conversion
from the adrenocortical androgens androstenedione, DHEA, and DHEA-S (see figure 1.9).
The aromatase enzyme present in adipose tissue, the muscle, the brain, and bone marrow
can convert T to estrogen. Androgens also may be converted by 5α-reductase to a more
potent and nonaromatizable 5α-DHT (figure 7.3).
The AR is one of the proteins coded by the q12 gene on the X chromosome (figure 7.4).
After binding with T, AR is responsible for the biological actions of T and DHT, including
the regulation of growth of muscle and bone, spermatogenesis, and the development of
secondary sexual characteristics and additional sexually dimorphic traits. The first of its
8 gene exons defines the HVD, exons 2 and 3 specify the DBD, and remaining exons code
the short hinge region and LBD or HBD (Cary & La Spada, 2008). The DBD contain two
OH

ER

HO Estradiol

Aromatase
OH

O Testosterone

AR
5α-reductase

OH

O DHT
H

Figure 7.3 The male sex steroids T and its nonaromatizable form DHT act on AR. In some tissues, T is aroma-
tized to estrogen that acts on the ER.Borer/E4517/Fig. 7.3/401591/Tim B/R2-alw
Reprinted, by permission, from A.T. Kicman, 2008, “Pharmacology of anabolic steroids,” British Journal of Pharmacology 154: 502-521.

Chromosome Xq12

AR gene
1 2 3 4 5 6 7 8

• Hormone binding
• Transcriptional
transactivation
• Transcriptional transactivation • Hinge region • Co-regulator binding
• Co-regulator binding • Two ZN-fingers • Dimerization • Dimerization

1 HVD DBD H HBD 919

Figure 7.4 Location of the AR gene on the X chromosome (top) and distribution of the eight exons coding for
Borer/E4517/Fig. 7.4/401592/Tim B/R2-alw
the HVD, DBD, and HBD of the AR.
Adapted from Physical Medicine and Rehabilitation Clinics of North America, Vol. 19, G.A. Cary and A.R. La Spada, “Androgen receptor
function in motor neuron survival and degeneration,” pgs. 479-494, copyright 2008, with permission from Elsevier.

153
154 Advanced Exercise Endocrinology

zinc fingers (subdomains with cysteine residues) that are inserted between specific grooves
on the DNA helix.
The biological effects of estrogens are mediated through ER α and β that have structural
homology with AR and other nuclear steroid receptors (see figure 1.18). As described in
chapter 1, sex hormones bind to cytoplasmic receptors and the H–R complex functions as
a nuclear TF by binding to the steroid responsive element in the promoter region of sex
hormone-responsive genes or by triggering several rapid, nongenomic signaling pathways.
The rapid nongenomic action of sex hormones involves binding to sex hormone receptors on
the plasma membrane (figure 7.5). This activates PI3K–Akt and MAPK through interactions
with adaptor proteins and signaling molecules such as a tyrosine kinase c-Src (Meyer et al.,
2009). Alternatively, sex steroids also bind to G protein-coupled ER (GPER in figure 7.5)
in the endoplasmic reticulum that then activate adenyl C, cAMP, and c-Src. c-Src activates
metalloproteinases (MMPs) and epidermal growth factor (EGF). The latter, in turn, activates
PI3K–Akt and MAPK that may lead to NO release in endothelial cells (figure 7.5). Some
nongenomic signaling pathways can activate nuclear TF.
Sex steroids affect a number of hypothalamic GH-controlling neurons and their response
to sex steroids. Androgens control hypothalamic SRIF (GH release inhibiting hormone)
synthesis, and estrogens modulate GHRH synthesis. During the neonatal period, gonadal
steroids epigenetically alter GH gene expression through DNA methylation or deacetylation
of histones (proteins that organize chromosome structure) (Waxman & O’Connor, 2006).

Nongenomic

EGF EGF
F E2
EG
Genomic 1
E2 EGFR
mER
Adaptor

P
MM c-Src

c-Src cAMP

2 PI3K/Akt
MAPK
GPER
Rapid
ER effects
(•NO)

5 6 7
4
3

ER P
ER TF

ER
TF

E4517/Borer/fig.7.5/401593/TB/R4-kh
Figure 7.5 Nongenomic mechanisms of hormone action of sex steroids involve membrane estrogen receptors
(mERs) or androgen and several signal transduction pathways.
Reprinted from Molecular and Cellular Endocrinology, Vol. 308, M.R. Meyer et al., “Non-genomic regulation of vascular cell function and
growth by estrogen,” pgs. 9-16, copyright 2009, with permission from Elsevier.
Exercise and Reproductive Hormones 155

Neonatal exposure to T imposes a distinct pulsatile, male-specific GH secretion, and in the


female, GH exhibits an irregular secretory pattern that is continuously higher (Gabory et
al., 2009). The male-specific pulsatile pattern of GH secretion activates STAT5b, which is
responsible for sexual dimorphism of some hepatic enzymes. STAT5b upregulates several
hepatic cytochrome P450 genes, including Cyp8B1 for 12 α-hydoxylase, Cyp2d9 for steroid
16 α-hydoxylase, and Cyp2A2 for male-specific steroid hydroxylase (figure 7.6). Several
hepatocyte nuclear TFs (HNF), but not STAT5b, upregulate the gene for the female-specific
steroid hydroxylase Cyp2C12.

Sex Dimorphism During Sexual Maturation and the Pubertal


Growth Spurt
Puberty represents a progression of stages of sexual maturation and includes a statural
growth spurt (figure 7.7). It is initiated by increased frequency and circadian distribution
of discharges of GnRH and GHRH neurons and a corresponding increase in pulsatile GH
and gonadotropin secretion (figure 7.2). The stages of sexual maturation are adrenarche
(synthesis of the adrenocortical androgen DHEA-S) at about 8 yr of age, thelarche (female
breast development) at about 9 to 10 yr, pubarche (growth of pubic and axillary hair and of

Sex hormones
Neonatal
period Inputs
Hypothalamus controlling
GH synthesis
Puberty and secretion
Anterior pituitary

GH
Plasma GH

Plasma GH

~3.5h
Male Female

Time (hr) Time (hr)

?
++

pY-STAT5b HNF4α HNF6 + HNF3β


Adult patterns
of sexual HNF3γ
dimorphism
Cyp2A2
Cyp8B1, Cyp2C12
Cyp2d9 Female-specific
Sexually Male-specific
dimorphic
TTCCGGGC TTCm5CGGGC (TF binding site)
epigenetic marks

Figure 7.6 Epigenetic imprinting of the sex-specific pattern of GH secretion and of expression of hepatic genes.
Sex steroids affect hypothalamic neurons that control GH secretion such that a male-specific pulsatile pattern, but
Borer/E4517/Fig. 7.6/401594/Tim B/R2-alw
not a continuously elevated and irregular female pattern, activates male-specific hepatic genes for cytochrome
P450 enzymes. Several hepatic nuclear transcription factors (HNFs) control the expression of female-specific
hepatic P450 enzymes.
Reprinted from Molecular and Cellular Endocrinology, Vol. 304, A. Gabory, L. Attig, and C. Junien, “Sexual dimorphism in environmental
epigenetic programming,” pgs. 8-18, copyright 2009, with permission from Elsevier.
156 Advanced Exercise Endocrinology

11½–16½
Height spurt
9½ –14½

2 3 4 5
Pubarche 2 3 4 5
BUD
Thelarche 2 3 4 5
8 –12
Menarche 10 –16½

Gonadarche 2 3 4 5

7 8 9 10 11 12 13 14 15 16
Age (yr)

Borer/E4517/Fig. 7.7/401595/TimB/R2-alw
Figure 7.7 Timing of pubertal maturation in girls (♀) and boys (♂).
Reprinted from J.M. Tanner, 1974, Sequence and tempo in the somatic changes in puberty. In Control of the onset of puberty, edited by M.M.
Grumbach, G.D. Grave, and F.E. Mayer (Philadelphia: PA: Lippincott, Williams, and Wilkins), 460. By permission of M.M. Grumbach.

apocrine sweat glands) at about age 11 yr, gonadarche (growth of male genitalia) at between
11 and 12 yr, and menarche (initiation of menstrual cycles) at between 12 yr and the teen
years. Pubertal sexual maturation can be quantified as five Tanner stages on the basis of the
degree of development of pubarche, thelarche, and gonadarche (Marshall & Tanner, 1986;
Tanner, 1974). Researchers frequently stratify children by Tanner stages in order to study
physical activity and motor skills because this stratification better reflects stages of sexual
maturation and statural growth than does a classification based on age.
The pubertal statural growth spurt depends on the increased secretion of both sex hormones
and GH because gonadal dysgenesis or isolated GH deficiency will cause children to attain only
half the expected statural growth (Brook & Hindmarsh, 1992; Zachman, 1992). The pubertal
growth spurt begins at about 11 yr of age in girls and about 2 yr later in boys, but it may be
delayed along with the onset of menarche in girls due to malnutrition or be accelerated in
obese children (Ahmed et al., 2009). In case of undernutrition, the signal that blocks the onset
of menarche (primary amenorrhea) is a change in LH pulsatility that regresses from a mature
pattern to a predominantly nocturnal, prepubertal pattern (Boyar et al., 1974). The signals that
mediate information about nutritional status to the key brain circuits responsible for initiation
of puberty and the growth spurt may be hormonal because increases in both plasma IGF-I
(Mauras et al., 1996) and leptin concentrations are observed around the time of initiation of
puberty (Ahima, 2004; Confavreux et al., 2009; Cunningham et al., 1999) and their titers are
modulated by nutritional and statural growth status (Martos-Moreno et al., 2010).
After rapid linear growth of about 2.5 cm/wk during the second trimester of gestation,
linear growth velocity declines to about 10 mm/wk during the first decade of childhood.
Growth velocity accelerates again during the pubertal growth spurt when sex differences in
the rate of skeletal growth develop (Cameron et al., 1982; Tanner, 1981). Because boys have
about 2 extra yr of slow preadolescent growth, they begin their pubertal growth spurt about
9 cm taller than girls. Linear growth velocity is higher in boys than in the girls because
the pattern of GH secretion in boys is more pulsatile (figure 7.6). The linear growth of leg
bones reaches peak velocity 0.6 yr before trunk length does. It slows at puberty and then
ceases as high titers of sex hormones close the epiphyseal growth zone (EGZ), whereas the
Exercise and Reproductive Hormones 157

elongation of the spine accelerates and continues for a while longer. Therefore, girls who
enter the pubertal growth surge about 2 yr years earlier than boys tend to have shorter legs
relative to trunk length compared with boys or late maturers (Marshall & Tanner, 1969, 1970,
1986). The long bones and vertebrae grow wider in males than in females due to greater
periosteal appositional growth. E2 inhibits periosteal bone expansion in females at puberty
but increases endosteal growth. This results in greater cortical thickness but a narrower
medullary cavity in the female adolescent. E2 is also responsible for the fusion of the EGZ
toward the end of the pubertal period in both sexes (Chagin et al., 2009).

Actions of Sex Hormones on Bone


T and E2 affect bone growth and bone mineral content in adulthood. T acts on ARs along
with GH and IGF-I, causing greater periosteal bone growth in boys. E2 controls bone growth
in both sexes by stimulating GH and IGF-I secretion (Seeman, 2001) through its actions on
periosteal and endosteal apposition and by causing the closure of EGZs. Its importance is
seen in males who have a congenital deletion of the aromatase gene Cyp19 or a mutation
of ERα. Such individuals have markedly delayed bone age, unfused EGZs, and osteopenia
in young adulthood (Vandenput & Ohlsson, 2009). Sex hormones also produce sex differ-
ences in the growth of the bones of the shoulder and pelvic girdle. The former is greater
in males and the latter is greater in females. This further reflects sex differences in tissue
distribution of ARs and ERs.
Sex steroids also are important in bone mineral homeostasis in adolescence and adult-
hood. Although two-dimensional DXA scans suggest that in adolescent males have greater
bone density than do females, density is shown to be equal in the two sexes when values
are corrected for the difference in bone size (Seeman, 2001). Bone mineral is maintained
through the cooperative actions of bone-resorbing multinucleated osteoclasts and bone-
forming osteoblasts (see also chapter 2). Osteoblasts control both osteoclast differentiation
and apoptosis (figure 7.8). Although nonaromatizable androgens suppress osteoclastogenesis
by osteoblasts, and both bone cell types express both ARs and ERs, bone turnover is primar-
ily controlled by estrogenic action over bone resorption (Imai et al., 2009).
Osteoclastogenesis is initiated by three ligands expressed by the osteoblasts: RANKL (the
receptor activator of nuclear factor kappa-B [NF-κB] ligand), M-CSF (figure 7.8), and osteo-
protegerin (OPG). These bind with respective RANK and c-fms receptors on the osteoclast
precursor cell membranes to stimulate the conversion of preosteoclasts into multinucleated
bone-resorbing osteoclasts. Estrogen suppresses bone resorption by binding to the ERα in
osteoclast nuclei to express the apoptosis gene FasL. FasL binds to the membrane receptor
CD95/Fas and leads to the activation of a Fas-associated death domain, the assembly of
the death-inducing signaling complex, and of a series of proteolytic enzyme caspases that
lead to osteoclast apoptosis. In addition, estrogen controls osteoclastogenesis by reducing
c-jun activity, which blocks transcription of RANKL and M-CSF by upregulating in osteo-
blasts OPG, a decoy receptor for RANKL (Khosla, 2001), and by decreasing the release of
proinflammatory cytokines IL-1, IL-6, and TNFα from bone marrow monocytes (Imai et
al., 2009; Lewis-Wambi & Jordan, 2009). In this way E2 controls the osteoclast life span
by regulating osteoclast formation and inducing osteoclast apoptosis. In addition to these
direct actions of estrogens on bone mineral, estrogen indirectly promotes bone mineral by
facilitating intestinal Ca absorption independently of vitamin D (Christakos et al., 2011;
O’Laughlin & Morris, 1998). Estrogenic control of bone turnover and resorption is relevant
in the context of reduced estrogen secretion in young amenorrheic athletes and postmeno-
pausal women. With estrogen deficiency, increased bone resorption leads to osteopenia,
osteoporosis, and increased risk of bone breaks.
158 Advanced Exercise Endocrinology

E2
FasL

Osteoclast ERα FasL FAS


Preosteoclasts

Apoptosis

Osteocytes
Osteoblasts
RANKL

RANK
M-CSF
c-fms
Osteoclast
precursors
Hematopoietic
Macrophage stem cells
lineage cells

Figure 7.8 Osteoblasts facilitate osteoclast differentiation by expressing M-CSF (macrophage colony-stimulating
factor) and RANKL ligands. Estrogen controls osteoclast apoptosis by binding to ERα and initiating expression of
the apoptosis gene FasL.
E4517/Borer/fig.7.8/401596/TB/R3-kh
Reprinted, by permission, from Y. Ima et al., 2009, “Molecular mechanisms underlying the effects of sex steroids on bone and mineral
metabolism,” Journal of Bone and Mineral Metabolism 27: 127-130.

At puberty, T stimulates erythrocyte proliferation from proerythroblasts in red bone


marrow. As a result, the hemoglobin level in men is about 12% greater than that in women
(16 vs. 14 g/dl of blood) and the oxygen-carrying capacity of blood is higher in men than
in women. Because venous oxygen content at maximal exercise is independent of sex, the
maximal arteriovenous oxygen difference is typically about 20% greater in adult men and
contributes to sex differences in maximal aerobic capacity.

Actions of Sex Hormones on Muscle


Sexual dimorphism in skeletal muscle growth is caused primarily by the anabolic actions of
T on cell proliferation during fetal development, protection of developing and mature moto-
neurons from apoptosis, sexually dimorphic determination of muscle fiber types during the
pubertal growth spurt, and stimulation of protein synthesis in contractile myofibrils during
skeletal muscle hypertrophy in adulthood. Dimorphism in muscle growth is expressed mostly
in the muscles of the upper extremities, generally after 8 yr of age and several years before
peak growth velocity (figure 7.9). The total muscle mass of young adult females is usually
about 20% smaller than that of young adult males of equal body mass. Analysis of muscle
fiber type on the lower-limb muscles, which are less sexually dimorphic than upper-limb
muscles, reveals a larger CSA of all three major fiber types in the vastus lateralis in men
compared with women. Type IIA fibers are largest in men and Type I fibers are largest in
women. The sex distribution of specific fiber types is IIA > I > IIB in men and I > IIA >
IIB in women (Staron et al., 2000).
Exercise and Reproductive Hormones 159

Researchers do not fully understand 8


the mechanism of androgen action in
muscle growth during fetal development. 6
Androgens stimulate the differentiation

Gain (mm/yr)
of pluripotent mesenchymal cells to the
4
myogenic lineage and the proliferation
of myoblasts and satellite cells during
both developmental growth and muscle 2
hypertrophy in adulthood. T stimulation
of skeletal muscle growth is tightly con- a 0
nected to actin-associated proteins that act 8
as AR coregulators (Ting & Chang, 2008).
During skeletal muscle myogenesis, myo-
6
blasts proliferate, differentiate, and fuse

Gain (mm/yr)
into multinucleated myotubes. This requires
elongation and reorganization of actin fila- 4
ments into myofibrils. Several actin filament
structural proteins act as AR coregulators 2
(figure 7.10). Among those, a scaffold
protein, paxillin, participates in myoblast
0
proliferation as well as in overload-induced −3 −2 −1 0 1 2 3
hypertrophy. Archvillin contains a domain b Time from PHV (yr)
that is required for interacting with AR
during myogenesis. Filamin C plays a role Figure 7.9 Velocities of muscle growth in girls and
boys during the pubertal growth spurt. Mean velocities
in myotube differentiation.
of (a) arm muscle width and (b) leg muscle width in girls
Androgens are neuroprotective of moto- (open circles) and boys (solid circles), normalized by the
neurons both during fetal development E4517/Borer/Fig7.9/401648/alw/pulled-R1
time of peak height velocity (PHV).
and in adulthood. They promote dendritic Reprinted, by permission, from J.M. Tanner, P.C.R. Hughes, and
arborization, increased motoneuron soma R.H. Whitehouse, 1981, “Radiographically determined widths of
bone, muscle and fat in the upper arm and calf from age 3-18
size, and axonal growth by stimulating years,” Annals of Human Biology 8: 505. By permission from Taylor
the expression of neurotrophic factors & Francis Ltd., http://www.tandf.co.uk/journals.
(NTFs)—their receptors (Cary & La Spada,
2008)—and prosurvival genes. GDNF
(glial cell line-derived NTF), BDNF (brain-derived NTF), and CNTF (ciliary NTF) are
particularly protective of motoneurons. They can be upregulated by either the genomic or
nongenomic activation of signaling by Ca, PI3K–Akt, or ERK–MAPK (Cary & La Spada,
2008; figure 7.5).
In adulthood, sex difference in the size of skeletal muscle is maintained by the capac-
ity of T to cause muscle hypertrophy. T increases net protein synthesis and reutilization
of intracellular amino acids without affecting inward amino acid transport or protein
degradation (Ferrando et al., 1998). In a multinucleate skeletal muscle fiber, each nucleus
controls protein synthesis in a finite cytoplasmic nuclear domain. T and high-resistance
exercise trigger increased synthesis of contractile proteins that requires addition of myo-
nuclei. These are added by the incorporation of satellite cells as myonuclei into myofibers
after T causes satellite cell proliferation (see figure 8.3). The myonuclei move from their
initial central location to their usual peripheral location on the surface of muscle fibers. T
induces hypertrophy in both Type I and Type II muscle fibers, but the effect is greater in
the more hormone-sensitive Type I fibers (Kadi, 2008). A type of positive feedback exists
160 Advanced Exercise Endocrinology

Myoblast Actin filament


Testosterone
Calpains AR
Gelsolin
Rac Filamin
Archvillin/Supervillin
Growth ARA55
Caspase Differentiation
Proliferation Paxillin
Apoptosis Myosin filament
P

Pyk2 FAK Fusion


+
Ca2 src
Focal Growth Neuron
ECM adhesion factors
Integrins
region
Anabolic Exercise
effect Insulin Costameres

+ Skeletal muscle
Ca2 Pyk2
Akt fiber
Calpains
PKC
P
Contraction

Nuclear

E4517/Borer/fig.7.10/401597/TB/R2-alw
Figure 7.10 The role of actin-associated proteins as coregulators of AR actions in muscle cell proliferation,
growth, differentiation, and apoptosis.
Reprinted from Journal of Steroid Biochemistry and Molecular Biology, Vol. 111, H.J. Ting and C. Chang, “Actin associated proteins function
as androgen receptor coregulators: an implication of androgen receptor’s roles in skeletal muscle,” pgs. 157-163, copyright 2008, with
permission from Elsevier.

between androgen action and muscle hypertrophy in that high-resistance training increases
the concentration of ARs (Bamman et al., 2001; Deschenes et al., 1994).
Administration of physiological doses of T to healthy males generally does not cause an
additional anabolic or androgenic effect, presumably because the ARs are saturated. This
dose that is physiological in males is anabolic in women (whose endogenous T concentra-
tions are between 20 and 80 ng/dl) and in hypogonadal men. Supraphysiological doses of
T are needed to induce anabolic effects in healthy young adults (Bhasin et al., 1996, 2001).
This issue is discussed in more detail in chapter 8.
Sex hormones also affect the structure and function of tendons and ligaments. The Achilles
tendon and the ACL of the knee are the two best-known tendons and ligaments. The CSA
of the Achilles tendons is smaller in women than it is in men. This gives women less elastic
recoil and less ability to release stored energy during locomotion, which may account for
why the rate of rupture of the tendon at about 5 cm above its insertion into the calcaneus,
an area that receives poor blood supply, is about twofold to sixfold lower in women than in
men. However, women are more prone to ACL injuries than are men, an effect that has been
attributed to the action of female sex hormones on E2 and P4 receptors in this ligament. The
higher incidence of knee-joint laxity at the phases of the menstrual cycle when E2 levels
are high (Shultz et al., 2005) and a correlation with knee-joint loads at those times (Park et
al., 2009) is not supported by some studies (Belanger et al., 2004; Van Lunen et al., 2003).
Exercise and Reproductive Hormones 161

Actions of Sex Hormones on 0.3

Adipose Tissue 0.2


Sex steroids induce dimorphism in body

Gain (mm/yr)
fat depot size and distribution. During the 0.1
pubertal growth spurt, women develop
significantly larger body fat depots (figure 0
7.11) and proportionally greater distribu-
−0.1
tion of SC fat in the thighs and buttocks
than do men. a −0.2
Both T and E2 suppress adipose tissue
0.3
mass through several mechanisms. T
stimulates lipolysis by regulating cat- 0.2
echolamine signal transduction in fat cells

Gain (mm/yr)
(Arner, 2005), inhibits lipid accumulation 0.1
by suppressing the activity of LPL, and
blocks differentiation of adipocyte pre- 0
cursor cells into fat cells (Blouin et al.,
2009). ERα inhibits fat cell hyperplasia −0.1
and hypertrophy, increases oxidative
metabolism of fatty acids by stimulating −0.2
−3 −2 −1 0 1 2 3
the expression of mitochondrial respira- b Time from PHV (yr)
tory chain enzymes and the enzymes for
beta oxidation, suppresses food intake, FigureE4517/Borer/Fig7.11/401598/alw/pulled-R1
7.11 Velocities of SC fat growth during the
and increases the expression of GLUT-4 pubertal growth spurt. Mean velocity of SC fat growth in
glucose transporters. On the other hand, the (a) arms and (b) legs of girls (open circles) and boys
E2 produces the opposite effects on adi- (solid circles), normalized by the time of peak height
velocity (PHV).
pose tissue growth when it acts on its
Reprinted, by permission, from J.M. Tanner, P.C.R. Hughes, and R.H.
ERβ that facilitate increases in fat cell Whitehouse, 1981, “Radiographically determined widths of bone,
and fat pad size and may be responsible muscle and fat in the upper arm and calf from age 3-18 years,” Annals
for sex-specific differences in body fat of Human Biology 8: 505. By permission from Taylor & Francis Ltd,
http://www.tandf.co.uk/journals.
distribution (Chen et al., 2009).
Aldo-keto reductase 1C (AKR1C) in
SC adipose tissue and cortisol in visceral adipose tissue have antiandrogenic actions. The
enzyme 11βHSD-1 increases its activity and converts cortisone to cortisol as T concentra-
tion decreases in muscles of aging males and visceral adipose tissue mass increases in
postmenopausal women. This may contribute to the accumulation of fat in skeletal muscle
and visceral adipose tissue in aging men and women.

EFFECTS OF EXERCISE ON SEX HORMONE SECRETION


How exercise affects sex hormone secretion is of interest for at least three reasons.
First, an acute exercise-induced increase in androgen secretion could improve physical
performance (see the section on “Effects of Sex Hormones on Physical Performance”).
Second, an increase in androgen secretion in females could increase muscle strength
and hypertrophy; this topic is discussed in chapter 8. Finally, an understanding of the
reproductive hormone response to energy depletion in endurance sports is of interest
in the context of human fertility. The additional issues regarding T secretion in RE
and its effect on muscle strength and hypertrophy in men are described in chapters 5
and 8, respectively.
162 Advanced Exercise Endocrinology

Sex Hormone Secretion in Acute Exercise


Exercise elicits different patterns of androgen secretion in response to exercise intensity,
duration, and type (i.e., aerobic vs. high resistance) and subject sex and age (Kraemer &
Ratamess, 2005). As discussed in chapter 5, high-intensity acute exercise, either resistance
or endurance type, is an effective stimulus for T release in the male—more so in adolescent
and young adult males than in older males (Kraemer & Ratamess, 2005). Researchers do not
fully understand precisely what causes increased T secretion under such circumstances. The
hypothesis that T release during acute exercise is triggered by the hypothalamic–pituitary–
gonadal (HPG) axis is not strongly supported because poor concordance exists between LH
and T secretion and poor correlation exists between the timing of the secretion of the two
hormones during acute exercise.
The alternative hypothesis is that T originates from the innermost reticular layer of the
adrenal cortex. Exercise-associated T secretion appears to be synchronous with the release
of adrenal androstenedione, cortisol, and DHEA. While LH secretion coincides with the
secretion of these steroids, it is unlikely to be the cause of T release for which about 45 min
are required (figure 7.12; Cumming et al., 1986). In addition, T concentration increases in men
at 40 to 50 min of running at 75% of maximal effort to exhaustion without a corresponding
change in LH concentration (Galbo et al., 1977; Tremblay et al., 2005). Such studies suggest
that during acute exercise stress, the increase in T along with ACTH concentration may be
of adrenocortical origin. On the other hand, researchers should not exclude the possibil-
ity of a mechanism of rapid testicular T release because some experiments suggest that S
stimulation may trigger such T release during acute exercise (Jezova & Vigas, 1981; Jezova
et al., 1985) and others document rapid T increases in response to viewing erotic, humor-
ous, and aggressive content (Cook & Crewther, 2012). Researchers have recently mapped
a direct neural pathway for the suppression of testicular T release that bypasses the HPG
axis (James et al., 2008), but the identity of a possible direct neural trigger of testicular T
release during exercise is currently unknown.
As is the case with the GH secretory pattern, a conspicuous sex dimorphism exists in
absolute levels and types of androgens in circulation and in their secretory response to
exercise. The concentration of T is approximately 10 times higher in male plasma than in
female plasma (figure 7.13a), as is expected given that the main source of this hormone is
the testis in the male and adrenal cortex in the female. On the other hand RE elicits higher
plasma concentrations of androstenedione, an androgen of adrenocortical origin, in the
female during the early follicular (EF) phase than in the male (figure 7.13b; Weiss et al.,
1983). Brief, intense RE with both upper and lower limbs at 80% of 1RM, and of the type
used in strength training, is a highly effective stimulus for T secretion in males but not in
females in the EF menstrual stage (figure 7.13a). Such exercise elicits comparable increases
in the adrenocortical androgen androstenedione in both sexes (figure 7.13b). On the other
hand, RE of moderate intensity (10RM) with both upper and lower limbs that is carried out
to failure can elicit T secretion in women in the EF phase. As in males, this type of acute
exercise in females elicits T and cortisol responses that differ in pattern and timing from
that of LH response (figure 7.14; Cumming et al., 1987). The stage of menstrual cycle can
affect sex hormone secretion during RE in women. A low volume (3 sets) of moderate-
intensity RE (10RM) elicits greater E2, androstenedione, and GH response during the
luteal menstrual phase than during the follicular menstrual phase. This exercise elicited no
change in T or P4 concentration but did stimulate an E2 response during the follicular phase
(Kraemer et al., 1995).
Exercise and Reproductive Hormones 163

Men
15 10,000
Exercise Testosterone
LH 7,500
mlU/ml

pg/ml
10
5,000

0 0
−30 0 30 60 90 120 −30 0 30 60 90 120
Time (min) Time (min)

160E4517/Borer/Fig7.12a/401599/alw/pulled-R1 15.0E4517/Borer/Fig7.12b/401600/alw/pulled-R2
Cortisol DHEA
140 12.5
ng/ml

µg/ml
120 10.0

80 7.5

0 0
−30 0 30 60 90 120 −30 0 30 60 90 120
Time (min) Time (min)

2,000
E4517/Borer/Fig7.12c/401601/alw/pulled-R1 E4517/Borer/Fig7.12d/401649/alw/pulled-R1
Androstenedione
pg/ml

1,500

1,000
−30 0 30 60 90 120
Time (min)

Figure 7.12 Timing of T, adrenocortical hormone, and LH secretion during acute exercise in men. Cycle ergometer exercise
E4517/Borer/Fig7.12e/401602/alw/pulled-R1
progressively increased to maximum intensity. Secretory timing of T was similar to that of adrenocortical hormones and was
too rapid to be influenced by LH secretion.
Adapted, by permission, from D.C. Cumming, et al., 1986, “Reproductive hormone increases in response to acute exercise in men,” Medicine and Science
in Sports and Exercise 18: 369-373.

The causes of exercise-associated decreases in T secretion are better understood than


are the causes of exercise-associated increases. Six hours after two brief but intense 300
m sprints that elicit strong T, androstenedione, GH, and LH responses but modest cortisol
responses in men, T concentration declines whereas LH concentration does not (Kuoppas-
almi et al., 1976). Acute moderate-intensity aerobic exercise at 75% of maximal effort and
carried to exhaustion is accompanied by significant decreases in plasma T but no associated
change in LH titers (Galbo et al., 1977). Similarly, men running at 55% of maximal effort
for between 40 and 180 min experience increases in concentrations of free and total T after
1 h and increases in DHEA and cortisol after 120 min, but T concentration starts declining
after about 80 min of running (Tremblay et al., 2005). Similar decreases in T but increases
in androstenedione and cortisol have been reported in males after noncompetitive marathon
runs (Dessypris et al., 1976).
164 Advanced Exercise Endocrinology

4.50 1.48
Female Female
4.25 1.40
Male 1.32 Male
4.00

Androstenedione (ng/ml)
3.75 1.24
Testosterone (ng/ml)

3.50 1.16
3.25 1.08
3.00 1.00
.92
1.00 .84
.75 .76
.50 .68
.25 .08
Weightlifting Weightlifting
.00
Control 0.0 0.5 1.0 1.5 2.0 Control 0.0 0.5 1.0 1.5 2.0
a Time (h) b Time (h)

Figure 7.13 Sex differences in basal plasma androgen concentrationsE4517/Borer/Fig7.13b/401650/alw/R1


and responses to high RE. (a) T; (b) androstenedione.
Springer E4517/Borer/Fig7.13a/401603/alw/R1
and European Journal of Applied Physiology and Occupational, Vol. 50, 1983, pgs. 415, “Comparison of serum testosterone and androstenedione
responses to weight lifting in men and women,” L.W. Weiss, K.J. Cureton, and F.N. Thompson, figure 1, with kind permission of Springer Science
+Business Media B.V.

The rapid onset of decreases in plasma T during acute intense and exhausting exercise,
and even several hours after discontinuation of such exercise, is most likely mediated by a
stress hormone neural circuit. Significant and large increases in plasma E and NE concen-
trations are a prominent feature of graded treadmill exercise that is performed to maximal
effort and are a marker of both central and peripheral catecholamine release (Galbo et
al., 1977). Activation of CRF in the PVN by beta adrenergic stimulation rapidly inhibits
T secretion by suppressing the activity of the steroidogenic acute regulatory protein in
Leydig cells without affecting testicular blood flow (James et al., 2008; Selvage & Rivier,
2003; Selvage et al., 2004). This neural pathway operates independently of the endocrine
HPG axis and accounts for the inhibitory effects of alcohol on T secretion. Although this
multisynaptic inhibitory circuit connecting the brain with the testes involves the nuclei in
the ventral noradrenergic pathway and limbic forebrain, PVN is its essential controlling
part (James et al., 2008; Selvage & Rivier, 2003).

Sex Hormone Secretion in Habitual Endurance Exercise


The suppression of the activity of the HPG axis in both sexes is a consistent response
to habitual endurance exercise that involves large volumes of energy-depleting physical
activity. Functioning of the HPG axis is suppressed in trained male marathon runners who
run between 125 and 200 km/wk (MacConnie et al., 1986). Although the baseline plasma
concentrations of gonadotropins and T and the endocrine response to acute exercise appear
normal in these males, a closer examination reveals that activity of the pituitary portion of
the HPG axis is suppressed. Thus, a 2 h run at about 75% of maximal effort (figure 7.15)
elicits an expected 62% increase in plasma T 40 min after the start of running. T concen-
tration declines within 2 h postexercise, whereas plasma LH concentration and its pulsatile
pattern remain unchanged. However, an examination of baseline gonadotropin pulsatility
reveals that LH pulse frequency is 39% lower and pulse amplitude is 44% lower in trained
males than in untrained males despite a similar mean LH concentration. In addition,
intravenous administration of exogenous GnRH at doses of between 0.25 and 2.5 μg/kg
elicits 25% and 32% lower plasma LH responses in trained compared to untrained males,
Exercise and Reproductive Hormones 165

200
LH Exercise

180

Percent change 160

140

120

100

Baseline = 7.9 ± 1.7 mlµ/ml


80
a

E4517/Borer/Fig7.14a/401604/alw/R1
75

Lactate (mg%)
3,000 2,000 50
Testosterone pmol/L

Cortisol (nmoI/L)
2,500 1,500 25

2,000 1,000 0

1,500 500

1,000 A B C E D F 0
−30 −15 Pre Resistance Post 15 30
b (min) exercises (min)

Figure 7.14 (a) Effects of a large volume of moderate-intensity RE on plasma concentrations of LH, steroid
hormones, and LA in women. (b) Isokinetic moderate-intensity (10RM) RE performed with the arms (A-C) and legs
to the point of failure elicits secretion of LH, T, and cortisol and release of LA.
E4517/Borer/Fig7.14b/401605/alw/R2
Reprinted, by permission, from D.C. Cumming et al., 1987, “Reproductive hormone responses to resistance exercise,” Medicine and Science
in Sports and Exercise 19: 234-238.

whereas intramuscular administration of an LH analog produces similar T responses in both


groups. Thus, the exposure to large volumes of endurance training suppresses the pituitary
gonadotropin response to GnRH in males, possibly due to increased negative feedback by
T during prolonged daily runs.
How these changes in the male HPG axis develop is seen in a 22-mo-long aerobic-training
study consisting of 60 wk of training at either a moderate (60% of maximal effort) or high
(80% of maximal effort) intensity. The reversibility of these changes is demonstrated during
a recovery period that consisted of 36 wk of low-intensity exercise (Safarinejad et al., 2009).
Serum total and free T, gonadotropins, and sperm concentration and motility decreased
throughout the 24 wk of training, more so during intense exercise than during moderate-
intensity exercise (figure 7.16). Training intensity also blunted, in an intensity-dependent
fashion, LH response to a 100 μg intravenous GnRH challenge (figure 7.17, top) and T
response to intramuscular injection of an LH analog (figure 7.17, bottom). All parameters
of HPG axis function returned to normal during the recovery phase, which consisted of 36
166 Advanced Exercise Endocrinology

12

10

Testosterone (ng/mI)
8

6
LH (mIU/mI)

0
40
Cortisol (µg/dI)

30

20

10

0
0 60 120 180 240 300 360
Time (min)

Figure 7.15 Endocrine responses toE4517/Borer/Fig7.15/401606/alw/R2


a 2 h run at 72% of maximal effort in endurance-trained male athletes
with altered basal gonadotropin action.
New England Journal of Medicine, S.E. MacConnie et al., “Decreased hypothalamic gonadotropin-releasing hormone secretion in male
marathon runners,” Vol. 315: 411-417. Copyright © 1986 Massachusetts Medical Society. Adapted with permission from Massachusetts
Medical Society.

wk of low-intensity exercise. These two studies strongly implicate the magnitude of stress
and energy drain associated with prolonged exposure to exercise of different intensity in
the suppression of pituitary and gonadal components of the HPG and fertility in the male
(MacConnie et al., 1986; Safarinejad et al., 2009). They also show that, in the male, HPG
suppression by this form of exercise is reversible.
In the female, high volume and intensity of endurance exercise or protracted dietary
restriction suppress reproductive function and the rate of statural growth. Energy restric-
tion interferes with reproductive function in the female more severely than that in the male
most likely because the growth and development of a fetus during pregnancy requires a
considerably greater investment of energy than does the production of sperm. About 4
decades ago, suppression of female LH pulsatility was found to occur with severe energy
restriction. Boyar and colleagues (1974) noted that anorexic female adolescents with delayed
onset of menarche or primary amenorrhea display a predominantly nocturnal prepubertal
pattern of LH pulsatility but that the pattern reverts to a mature circhoral pattern upon
realimentation. Around the same time, researchers established an association between
Exercise and Reproductive Hormones 167

8
LH
7
6
5

IU/L
4
3
2
1
0
0 12 24 36 48 60 12 24 36
16
14 Testosterone
12
10
nmoI/L

8
6
4
2
0
0 12 24 36 48 60 12 24 36

80
Sperm concentration
70
60
Millions/mI

50
40
30
20
10
0
0 12 24 36 48 60 12 24 36

70
Sperm motility
60
50
40
%

30
20
10
0
0 12 24 36 48 60 12 24 36
Running phase Weeks Recovery phase

Figure 7.16 Changes in LH and T plasma concentrations, sperm concentration, and sperm motility during 60
wk of aerobic training at either 80% or 60% of maximal effort and 36 wk of low-intensity exercise that allowed
recovery from previous training. HighE4517/Borer/Fig7.16/401607/alw/R1
training intensity (solid symbols) suppressed LH and T concentrations and
sperm count and motility more than moderate-intensity training did (open symbols). Both the hormone and sperm
parameters returned to normal during 36 wk of recovery at low-intensity exercise.
Adapted, by permission, from M.R. Safarinejad, K. Azma, A.A. Kolahi, 2009, “The effects of intensive, long-term treadmill running on
reproductive hormones, hypothalamus-pituitary-testis axis, and semen quality: A randomized controlled study,” Journal of Endocrinology 200:
259-271.

the incidence of primary and secondary (cessation of menstrual periods after puberty)
amenorrhea and sports such as ballet, gymnastics, and distance running in which athletes
restrict their diets in order to meet esthetic or performance demands (Frisch & Revelle,
1971; Malina et al., 1973; Warren, 1980). In such sports, energy expenditure substantially
exceeds energy intake, is associated with reduced body fat, and is often combined with a
high degree of stress.
168 Advanced Exercise Endocrinology

Increase from basal value (%)


600
LH
500
400
300
200
100
0
0 20 40 60 120 0 20 40 60 120
−100 Running phase Time (min) Recovery phase
Increase from basal value (%)

500
Testosterone
400
300
200
100
0
0 20 40 60 120 0 20 40 60 120
−100 Running phase Time (min) Recovery phase

Figure 7.17 Reduced pituitary LH response to a GnRH challenge and T response to an LH analog challenge
after 60 wk of high-intensity endurance training and 36 wk of low-intensity training. High-training intensity (solid
symbols) suppressed LH and T responses more than moderate-intensity training did (open symbols). Both responses
E4517/Borer/Fig7.17/401608/alw/R1
returned to normal during 36 wk of recovery at low-intensity exercise.
Adapted, by permission, from M.R. Safarinejad, K. Azma, and A.A. Kolahi, 2003, “The effects of intensive, long-term treadmill running on
reproductive hormones, hypothalamus-pituitary-testis axis, and semen quality: A randomized controlled study,” Journal of Endocrinology
200(3): 259-71.

Acute exercise of increasing intensity alone is not sufficient to induce LH secretory


abnormalities, as was shown in the study of Williams and colleagues (1994) in which 3
consecutive 20 min runs at intensities increasing from 50% to 70% of maximal effort had
no effect on LH pulse frequency and slightly increased only LH amplitude. Exercise has
to be aerobic, habitual, and sufficiently demanding in terms of energy expenditure in order
to produce a substantial energy drain. Thus, even in the absence of clinical evidence of
menstrual disturbances, female cyclists and participants in competitive sports display an
incidence in abnormalities of sex hormone production that is 10 times higher than in non-
exercising women only if they are studied closely (De Souza et al., 2010). The important
role of limited energy availability (the difference between nutrient energy provided and
the energy cost of exercise) in the development of menstrual abnormalities and exercise-
associated amenorrhea was demonstrated in an acute 5 d study in which energy availability
by exercise or dietary restriction was reduced from 30 kcal·kg of fat-free mass-1·d-1 to 10
kcal·kg-1·d-1 (Loucks & Thuma, 2003; Loucks et al., 1998). As energy availability declined
from 30 to 10 kcal·kg-1·d-1, proportional changes occurred in metabolic markers of energy
deprivation and hormone responses. The decline in plasma glucose was accompanied by
increased production of hepatic ketone bodies (figure 7.18a). Plasma cortisol concentration
increased (figure 7.18b) along with GH pulsatility, whereas plasma leptin, T3, insulin, and
IGF-I concentrations decreased (figure 7.18, b-d). The HPG axis responded to this shortfall
75 75
50 ßHOB/50 50
Cortisol
25 25
Effects (%)

Effects (%)
0 0
−25 −25
2 ∙ Glucose
−50 −50
Insulin
−75 −75
0 10 20 30 40 50 0 10 20 30 40 50
a Energy availability (kcal ∙ kgFFM−1 ∙ day−1) b Energy availability (kcal ∙ kgFFM−1 ∙ day−1)

75 75
E4517/Borer/Fig7.18a/401609/alw/R2 E4517/Borer/Fig7.18b/401610/alw/R2
50 50
GH
25 25
Effects (%)

Effects (%)
0 0
IGF-I/BP3
−25 −25
IGF-I 2 · T3
−50 −50
IGF-I/BP Leptin
−75 −75
0 10 20 30 40 50 0 10 20 30 40 50
c Energy availability (kcal ∙ kgFFM−1 ∙ day−1) d Energy availability (kcal ∙ kgFFM−1 ∙ day−1)

75 75
E4517/Borer/Fig7.18c/401611/alw/R3-kh A11/3
E4517/Borer/Fig7.18d/401658/alw/R2
50 50
A/3
25 25
A>11/3
Percent

Percent

0 0
F>11
−25 −25
F
−50 −50
F11
−75 −75
0 10 20 30 40 50 0 10 20 30 40 50
e Energy availability (kcal ∙ kgFFM−1 ∙ day−1) f Energy availability (kcal ∙ kgFFM−1 ∙ day−1)

Figure 7.18 Hormonal and metabolic changes to reductions in energy availability from 45 kcal/kg of lean body mass to 10
during 5 d of exercise at 70% of maximal effort. (a) ketone body βHOB
kcal/kg E4517/Borer/Fig7.18e/448757/alw/R2 (βhydroxybutyrate) and
E4517/Borer/Fig7.18f glucose concentra-
/448758/alw/R2
tions. (b) Plasma cortisol and insulin. (c) Plasma concentrations of GH, IGF-1, and IGF BP 1 and 3. (d) Plasma concentrations
of T3 and leptin. (e-f) LH pulse amplitude (A/3) and frequency were different for women with luteal phase length of greater
or less than 11 d.
Reprinted, by permission, from A.B. Loucks and J.R. Thuma 2003, “Luteinizing hormone pulsatility is disrupted at a threshold of energy availability in
regularly menstruating women,” Journal of Clinical Endocrinology and Metabolism 88: 297-311.

169
170 Advanced Exercise Endocrinology

in energy availability by reducing basal LH pulse frequency and increasing LH amplitude


(figure 7.18, e-f). Hormonal and metabolic changes blocked ovulation and produced a sig-
nificant decline in follicular and luteal production of E2 (Loucks et al., 1989). Other studies
have induced shortening of the luteal phase of the menstrual cycle and reduction in luteal
P4 production in adolescent females subjected to increased energy drain through exercise
training and dietary restriction (Beitins et al., 1991). A severe decline in E2 production in
response to energy-costly endurance sports leads to premature bone mineral loss, osteope-
nia, and osteoporosis due to reduced estrogenic actions on dietary Ca absorption and the
maintenance of bone mineral (Borer, 2005; Rencken et al., 1996).
Whereas reduction in training intensity in men rendered hypogonadal by protracted
intense endurance training (Safarinejad et al., 2009) and realimentation of amenorrheic
females restores the mature pulsatile pattern of LH secretion (Boyar et al., 1974), restora-
tion of menstrual cycles was observed in only about 18% of 373 collegiate athletes who
displayed oligomenorrhea and amenorrhea (Arends et al., 2012). Weight regain and an
increase in BMI were prerequisites for restoration of normal menstrual function. On the
other hand, once substantial losses of bone mineral content take place, E2 administration
in the form of contraceptive therapy does not correct the bone deficits, and reductions in
training volume or increases in energy intake often are unacceptable to competitive female
athletes (Ducher et al., 2011).
The demonstration of the critical importance of low energy availability in inducing
menstrual disturbances in female athletes does not exclude the possibility that prolonged
exposure to stress of such exercise or substantial losses of body fat contribute individually or
collectively to the development or maintenance of amenorrhea. The stress hormones CRF,
ACTH, cortisol, β-endorphin, E, and NE are all released during intense endurance exercise
and are known to inhibit GnRH secretion (Berga et al., 1988; Ellingboe et al., 1982; Rivier &
Rivest, 1991). Chronic elevation of cortisol in amenorrheic athletes is a result of attenuated
negative feedback in the hypothalamo–pituitary–adrenal axis (De Souza et al., 1994). In a
primate model, a combination of environmental stress and dietary restriction increases more
than threefold the incidence of abnormal menstrual cycles compared with the individual
impacts of these stressors (Williams et al., 2007). In addition, the association of amenorrhea
and delayed onset of menarche with reduced body fat level has been recognized for at least
40 yr (Frisch, 1972; Frisch & McArthur, 1974; Frisch & Revelle, 1971). Close associations
between body fat and plasma leptin concentration on one hand and between plasma leptin
concentration and the onset of menarche on the other (Donoso et al., 2010), and induction
by leptin of menstrual cycles in circumstances of leptin deficiency (von Schnurbein et al.,
2012), attest to the strong possibility that body fatness plays an important role in the etiology
and maintenance of exercise-induced amenorrhea.

EFFECTS OF SEX HORMONES ON PHYSICAL


PERFORMANCE
The possibility that sex hormones may have an effect on exercise performance has been of
interest to athletes of both sexes and to women who desire to continue participating in sport
during pregnancy. Of particular interest is whether short-term increases in plasma T may
have an effect on strength and power in men and whether sex hormone differences across
menstrual (circamensal) phases may affect exercise performance in women.
Exercise and Reproductive Hormones 171

Effects of T on Acute Performance


T concentration can rapidly increase when men view erotic or humorous material, aggressive
behavior, or training procedures, and cortisol is elicited when men view aggressive behavior.
These changes in plasma concentrations of T and cortisol were accompanied by significant
increases in 3RM squat performance. A significant within-individual correlation also existed
between the relative changes in plasma T and the 3RM squats with the visual experience
of all types (Cook & Crewther, 2012). This raises the possibility that exercise-associated
increases in plasma T could have a positive effect on physical performance. T actions on
electrophysiological and contractile properties of skeletal muscle support this hypothesis
(Crewther et al., 2011). T stimulates the excitability of muscle fiber membranes and neuro-
muscular transmission, facilitates rapid influx of Ca into skeletal muscle cells, upregulates
intracellular Ca channels, and stimulates Ca release from the sarcoplasmic reticulum.

Effects of Menstrual and Contraceptive Sex Hormone Changes


on Performance
The sex hormone milieu is significantly different during the EF, periovulatory (PO), and
midluteal (ML) phases of the female menstrual cycle (figure 7.1). Concentrations of both
E2 and P4 are low during the EF phase. E2 reaches its highest concentration and P4 remains
low during the PO phase, and concentrations of both hormones (particularly P4) are high
during the ML phase. Contraceptives block ovulation by maintaining circulating concen-
trations of both female sex steroids that are often supraphysiologically high. So, besides
the developmental sex differences in body composition, allometric difference in the size
of key organs like the heart, biomechanical features of the skeleton, and oxygen carrying
capacity of blood, researchers are interested in finding out whether physical performance
is enhanced or compromised in very different hormonal environments across the menstrual
cycle. Of special interest are sex hormone-dependent differences in circulation, ventilation,
thermoregulation, muscle function, and substrate utilization that may contribute to a dif-
ference in exercise performance.

Female Sex Hormone Effects on Circulation in Exercise


Premenopausal females experience considerably less cardiovascular disease than do men,
possibly due in part to cyclical elevations of E2. In females, blood pressure is lowest and vas-
cular reactivity to flow-mediated dilatation is greatest during the ML phase of the menstrual
cycle, when E2 concentration is highest. This effect is attributed, to a large extent, to the
release of NO (Adkisson et al., 2010). A decline in blood pressure during the ML phase is in
part caused by reduced vascular responsiveness to the vasoconstrictive effects of α2, but not
a1, agonists(Limberg et al., 2010). In response to orthostatic challenges such as head-down
tilt, increases in mean arterial blood pressure and in vascular resistance in the arms, but not
legs, are greater during the ML phase than during the EF phase, although reflex activation
of S activity to the muscle is unaffected by circamensal hormones (Lawrence et al., 2010).
Circamensal fluctuations in female sex hormones do not affect systemic and femoral increases
in vascular conductance and postexercise decreases in blood pressure after 1 h of exercise
at 60% of maximal effort (Lynn et al., 2007) or steady-state dynamic exercise at 15% and
30% of maximal voluntary contraction (Limberg et al., 2010). Thus, the modest effects of
female sex hormones at rest do not appear to influence vascular responses during exercise.
172 Advanced Exercise Endocrinology

Female Sex Hormone Effects on Ventilation During Exercise


Ventilatory function is affected by menstrual variation in sex hormones at rest. During the
ML phase, arterial pCO2 is lower and arterial acidity, basal VE, and VE to hypercapnia and
hypoxia are higher than during the EF phase (Dombovy et al., 1987; Macnutt et al., 2012).
Increased response to hypercapnia, but not hypoxia, is attributable to P4 (Bonekat et al.,
1987). However, despite increased VE drive, lower VE recruitment threshold, and lower VE
threshold during incremental exercise during the ML phase, sex hormone variations do not
affect performance at maximal exercise intensity (Bemben et al., 1995; Dombovy et al.,
1987; Smekal et al., 2007).

Female Sex Hormone Effects on Thermoregulation During


Exercise
A 0.3 °C increase in Tc marks the transition from the PO to the luteal phase of the menstrual
cycle. This adds a concern about the risk of heat illness in women, who have a heat-dissipation
handicap in that they have a lower body surface:body mass ratio and higher body fat content
compared with men. The ML phase is associated with a 0.5 °C increase in sweating and
vasodilatatory thresholds at rest and during moderate- to high-intensity exercise through a
range of ambient temperatures and humidities (Hessemer & Brück, 1985; Hirata et al., 1986;
Kolka & Stephenson, 1989, 1997). At a given exercise intensity, no change occurred in the
sensitivity of the thermoregulatory response, and forearm blood flow stabilized at a higher
level than it did during the EF phase (Kolka & Stephenson, 1989), but the thermoregulatory
slope was reduced with increases in exercise intensity (Hirata et al., 1986). Similarly, during
the ML phase, sweating rate increased by 11% during 1 h of exercise at 60% of maximal
effort in a hot and very humid environment (32 °C, 80% relative humidity) compared with
the EF phase (Garcia et al., 2006). This apparent upward change in the thermoregulatory
setpoint during the ML phase is related to reduced neuronal firing in the thermosensitive
hypothalamic preoptic area (Nakayama et al., 1975). The consequence of the thermoregu-
latory changes during the ML menstrual phase or during use of oral contraceptives is a
decreased tolerance for physical work in the heat, regardless of whether it entails exercise
in humid environment (Pivarnik et al., 1992) or light exercise in an environment of uncom-
pensable heat (Tenaglia et al., 1999). By contrast, circamensal variation in sex hormones
has no effect on thermoregulation in the cold (Glickman-Weiss et al., 2000).

Female Sex Hormone Effects on Muscle Function in Exercise


Circamensal variations in sex hormones do not affect muscle strength (Elliott et al., 2003;
Fridén et al., 2003; Janse de Jonge et al., 2001), endurance (Fridén et al., 2003), or increases
in myofibrillar protein and collagen synthesis 24 h after 1 h of leg exercise at 67% of maximal
contraction (Miller et al., 2006). However, users of oral contraceptives experienced lower
magnitude and duration of delayed onset muscle soreness and concentrations of creatine
kinase after 30 min of downhill running at 60% of maximal effort (Carter et al., 2001) and
50 min of stair stepping (Thompson et al., 1997). Before and after E2 administration, females
upregulate more skeletal muscle genes involved in fat metabolism (e.g., genes for PGC-1α,
PPARα,γ, CPT 1, and SREBP-1c) than do males (Fu et al., 2009).

Female Sex Hormone Effects on Substrate Utilization in Exercise


Researchers have recognized for some time that women rely more on fat metabolism during
some forms of exercise than do men. The effect is seen during endurance exercise at intensi-
Exercise and Reproductive Hormones 173

ties between 40% and 80% of maximal effort (Friedman & Kinderman, 1989; Horton et al.,
1998; Mittendorfer et al., 2002; Tarnopolsky et al., 1990) but not at maximal effort (Froberg
& Pedersen, 1984). This effect is attributed to the greater sensitivity of female SC abdominal
adipose tissue to beta AR activation, higher plasma concentrations of GH, and lower activa-
tion of adipose tissue LPL by E2 (Hardman, 1998; Hellstrom et al., 1996; Jensen et al., 1996).
A difference in substrate utilization also is observed as a function of circamensal changes
in sex hormone concentrations. During the ML phase (Hackney et al., 1994; Wenz et al.,
1997; Zderic et al., 2001) or during the use of oral contraceptives (Casazza et al., 2004),
lipid oxidation is higher relative to carbohydrate utilization at exercise intensities of up to
65% of maximal effort. In addition, lipid clearance was increased during the ML phase of
the menstrual cycle in response to a high-fat meal (Gill et al., 2005). Higher lipid utilization
during exercise in the ML phase is enhanced when carbohydrate availability is low (Wenz
et al., 1997) and is reduced when carbohydrate is supplemented (Campbell et al., 2001). No
influence of circamensal hormones is observed at exercise intensities greater than 70% of
maximal effort (Hackney et al., 1994; Vaiksaar et al., 2011; Wenz et al., 1997). According
to two recent reviews (Isacco et al., 2012; Oosthuyse & Bosch, 2010), these results suggest
that high titers of E2 at low intensities of aerobic exercise increase FFA availability (in part
through activation of AMPK), increase lipid oxidative capacity, and, along with P4, suppress
gluconeogenesis. P4, on the other hand, opposes E2 action on lipid oxidation and promotes
protein catabolism. Use of oral contraceptive facilitates lipid utilization in endurance activi-
ties, and a concomitant increase occurs to lipolytic hormones and GH. With the onset of
menopause and a decline in E2 titers, the utilization of lipids during exercise declines and
becomes similar to that observed in men.
From the evidence presented, one could justifiably conclude that, with the exception of
a thermoregulatory handicap for exercise in the heat during the luteal phase, the effects of
variations in sex hormones during the menstrual cycle do not appear to significantly impair
performance in females. This conclusion has been repeatedly supported in studies that assess
peak aerobic performance (Casazza et al., 2002) and endurance (Burrows & Bird, 2005;
Vaiksaar et al., 2011) and anaerobic (Bushman et al., 2006) and peak and mean (Sunderland
et al., 2011) power performance.

Effects of Exercise During Pregnancy


Exercise (Oscai et al., 1972) and dietary restriction (Oscai et al., 1974) have an apparent
epigenetic capacity during the neonatal period to permanently reduce body mass and fat
content in sampled fat depots in rats. More recently, dietary restriction during human
pregnancy was shown to reduce body size at birth and in adulthood and to cause increased
visceral obesity and plasma cholesterol, higher incidence of T2D and coronary heart disease,
higher preference for fatty foods, and increased tendency for reduced physical activity in
adulthood (Lussana et al., 2008). Because exercise can reduce energy availability and cause
endocrine and metabolic adaptations, researchers must better understand how exercise may
affect placental and fetal growth as well as the health and well-being of a child. One such
possible epigenetic effect of exercise is its capacity to reduce the strong 60% genetic influence
over expression of obesity in adulthood (Andreasen et al., 2008; McCaffery et al., 2009).
During pregnancy, increases in maternal and fetal body mass are accompanied by increases
in total and resting metabolic rate. Concurrently, spontaneous physical activity declines and
preference shifts to lower-intensity physical activities. This is the most likely reason why
researchers frequently report that occupational and recreational physical activities have either
little or mostly moderate beneficial effects on pregnancy outcomes (Olson et al., 2009).
174 Advanced Exercise Endocrinology

With the advent of ultrasound technology that permits one to take measurements of
placental and fetal growth (Thomas et al., 2008) and better examine the effects of specific
exercise modalities, intensities, durations, and timing with respect to stage of pregnancy,
new evidence suggests that exercise modality is one of the more significant variables in
the outcome of pregnancy (Clapp, 2006). RE of moderate intensity during early pregnancy
has a positive stimulatory effect on placental growth and vascularization. The particularly
beneficial effect of this exercise modality on the proliferation of terminal villi ensures a
more abundant supply of oxygen and nutrients to support fetal growth and an approxi-
mately 250 g increase in birth weight, 1 cm increase in birth length, and balanced effect
on body composition. In contrast, prolonged and intense endurance activities, especially
in women who have engaged in this type of exercise before pregnancy and enter it with
reduced body fat mass, produce only moderate placental growth, particularly of stem and
intermediate rather than terminal villi, and result in slower fetal growth that may result in
a 300 g decrease in birth weight, little change in birth length, and lower body fat content
of the newborn (Clapp, 2006).
The mechanism through which aerobic exercise or RE produce different effects on
pregnancy outcome includes a differential effect on stimulation of TNFα and HIF-1 and
suppression of factors that inhibit the production of IGF-I and VEGF. The specific stimuli
that trigger the differential effect are not fully understood. These may include the inter-
mittent exercise-associated reduction (~50%) in blood flow to the placenta and in oxygen
and nutrient supply to the fetus over a 24 h period. RE also is likely to produce less of an
oxygen and nutrient shortfall, whereas prolonged endurance exercise may produce a more
pronounced shortfall. The essential feature of a positive adaptive response to RE is the inter-
mittent pattern of circulatory supply to the placenta during exercise. Prolonged suppression
of placental blood supply followed by delayed increase or sustained prolonged oversupply
does not produce the same adaptive responses.
The second important epigenetic effect on pregnancy outcome is the availability of
dietary carbohydrate and maternal plasma glucose that interacts with exercise to affect
placental growth (Clapp, 2002). Thus, pregnant women who consumed a mixed diet in
which about 60% of calories had low glycemic index displayed no changes in postprandial
plasma glucose, and they delivered infants of normal size (25th-75th percentile). When
women consumed diets with a high glycemic index, postprandial plasma glucose increased
190% with advancing gestation. These women delivered infants of normal weight but of lean
body mass greater than the 90th percentile without an excessive increase in infant fat mass.
Reduced carbohydrate delivery to the fetus reduces placental and fetal growth. Because
early influences of exercise and diet may have a profound effect on growth of lean body
mass, adipose tissue size, and metabolic risk factors in adulthood, focused research in this
area may have potential clinical value and deserves more attention.

SUMMARY
This chapter discusses three themes that relate sex hormone secretion and action to physi-
cal activity: how sex chromosomes, through coding for sex hormones and their receptors,
guide phenotypic dimorphism during puberty and establish reproductive competence in
adulthood; how different types of exercise affect sex hormone secretion in the two sexes;
and how male and female sex hormones affect physical activity and performance. The first
section describes the development of gonads, the guidance by sex hormones of the pubertal
growth spurt and of sexually dimorphic body build, and the pattern of sex hormone secre-
175
176 Advanced Exercise Endocrinology

tion necessary for the maintenance of fertility. This section also describes genomic and
nongenomic actions of sex hormones. The second section describes sex differences in the
secretion of sex hormones in response to different types and modalities of exercise. It also
discusses putative and established mechanisms for the stimulation and suppression of sex
hormone secretion during acute and prolonged energy-depleting exercise. The final section
outlines the effects of male sex hormones and variations in the female sex hormones on
physiological functions that have an effect on physical performance.
H APTER

8
C

Hormonal Mediation
in Training
Adaptations
C
hemical messengers play an important role in adaptations to both endurance and
resistance training. They respond either to physical stress, mechanical strain, and
neural signals or to metabolic, bioenergetic, hypoxic, or temperature cues generated
through aerobic exercise or RE. The same stimuli that elicit endocrine responses in the
tissues usually can initiate cellular signaling responses in the absence of hormonal action
(see chapter 2). These signaling cascades may be shared by the hormones at the receptor
end and by cellular stimuli at more distal parts, or they may differ in signaling details but
produce similar changes in enzyme activity or gene expression. The resulting structural and
functional adaptations allow more efficient and effective execution of aerobic and resistance
activities. Most training adaptations, be they new enzymes or new tissue, require adequate
supply of nutrient energy and amino acid building blocks. The sensing of changes in energy
availability occurs in musculoskeletal tissues engaged in physical activity, in fuel-storing
organs that are affected by the energy costs of exercise, and in the brain. All of these areas
initiate additional signaling pathways, either through activation of hormone secretion or
autonomic nerves or directly in the affected organs.
Several complications may arise when linking the actions of circulating hormones to
specific training adaptations. One complication may include direct hormonal activation of
heart and skeletal muscle contractions (by catecholamines, for instance) that in themselves
can cause structural and functional adaptations. Another complication is that the same hor-
mone (e.g., IGF-I) can be produced in the liver in circulating form and can be upregulated
by physical activity in tissues such as the muscle or the bone, where it exerts local influ-
ence. Most hormones have pleiotropic effects and can act under different circumstances
as metabolic, anabolic, or catabolic agents. For example, GH affects tissue growth as well
as lipolysis, sweating, thermoregulation, extracellular and plasma fluid volumes, and heart
structure and function. In addition, hormones may exert completely different actions during

177
178 Advanced Exercise Endocrinology

exercise than they do at rest. Hydromineral hormones reabsorb water (ADH) and Na (aldo-
sterone) or affect PV (ADH, ANP, and aldosterone) at rest, but during exercise (see chapter
4) they primarily increase blood pressure (angiotension II and ADH) and stimulate lipolysis
(ANP) because their hydromineral actions are curtailed by reduced renal blood flow. Further,
hormone actions can be synergistic or antagonistic. Actions of insulin and cortisol, or T and
cortisol, are antagonistic with respect to protein synthesis and tissue growth. Cortisol, GH,
and glucagon synergistically increase blood glucose concentration, whereas cortisol and
insulin favor fat synthesis at rest but cortisol and GH support lipolysis during exercise. To
establish a causative role in training adaptations, systemic hormones and training adaptations
have to demonstrate associations that are consistent, biologically and temporally plausible,
and related to the adaptations in a dose-dependent fashion. Hormonal and local effects can
be dissociated with a unilateral limb-training paradigm in which hormonal contribution to
functional and structural adaptations can be observed in both trained and untrained limbs,
whereas local adaptations are limited to the trained limb.
This chapter examines how hormones elicited by aerobic exercise and RE augment and
complement nonhormonal signaling initiated in affected tissues by physical activity to
produce adaptations to exercise training. Several questions are of particular interest: Do cir-
culating hormones induce training adaptations directly or indirectly by changing conditions
in tissues affected by exercise? Are training adaptations produced and maintained through
chronic changes in concentrations of systemic hormone? Do systemic hormones exert the
same or different actions during exercise and at rest? Which of the many actions of systemic
hormones influence training adaptations? Research on the role of the endocrine system in
training adaptations is also driven by our desire and ability to manipulate hormones with
the goal of improving physical performance.

SYSTEMIC HORMONES IN ADAPTATIONS


TO ENDURANCE TRAINING
Endurance training improves aerobic fitness by increasing cardiorespiratory capacity and
increasing the ability of muscles to oxidize lipid and carbohydrate substrates. In order to
develop and maintain increased aerobic fitness, exercise training must be carried out at high
intensities for prolonged periods of time or intermittently at near-maximal intensities. This
type of training promotes adaptations by creating tissue hypoxia when exercise intensity
exceeds cardiorespiratory and circulatory capacity to deliver oxygen to the muscle and by
causing increases in LA and hydrogen ion concentrations in the tissues and the blood as
a result of a mismatch between the availability of metabolic fuels and the ability of the
mitochondria to oxidize them. Elicitation of hormonal and tissue responses to these stimuli
improves the delivery of oxygen to the muscle by increasing cardiac output, overall PV,
capillary density, and respiratory capacity. Fuel delivery is improved by increased sensitiv-
ity of the liver, muscle, and adipose tissue to fuel-mobilizing actions of counterregulatory
hormones. Improvements in muscle oxidative capacity are achieved through an increase in
mitochondrial biogenesis and conversion of anaerobic glycolytic muscle fibers to fibers that
have higher oxidative capacity. Conversion to more oxidative muscle fibers is accompanied
by changes in the pattern of motoneuron firing and intracellular Ca release. The follow-
ing section outlines how systemic hormone secretion complements intracellular signaling
pathways (discussed in chapter 2) to produce adaptations to endurance training.
The interest in the role of systemic hormones in adaptation to endurance training is
a natural outgrowth of the observation that acute aerobic exercise triggers secretion of
Hormonal Mediation in Training Adaptations 179

catecholamines, sex hormones, GH, and IGF-I, usually in a dose-dependent fashion (see
chapters 3 and 5), and that many of these hormones control metabolism and tissue growth
(see chapter 7) in a way that could explain adaptive changes to exercise training. Some con-
comitants of aerobic exercise, such as hypoxia and increased LA production, elicit GH and
cortisol secretion (Sutton, 1977; VanHelder et al., 1987). When marked changes in hormone
secretion are not observed during aerobic exercise, as is the case with thyroid hormones,
the involvement of T3 and T4 is assumed because of the known influence of these hormones
on tissue adaptations that are seen after endurance training.

Role of GH
Among systemic hormones, temporally and biologically plausible evidence most consistently
shows that GH is a mediator of two important adaptations to endurance exercise: increased
oxygen and substrate delivery to, and lipid utilization by, exercising muscle (Birzniece et al.,
2011; Møller & Jorgensen, 2009; Widdowson et al., 2009). Biological plausibility for the role
of GH in these adaptations stems from a positive correlation between GH concentrations and
aerobic fitness (Eliakim et al., 1996; Weltman et al., 1994), improved oxygen and substrate
delivery to exercising muscle in GH deficient (GHD) subjects receiving supplemental GH,
and changes in these functions in acromegalics while they oversecrete GH and after they
are treated with inhibitors of GH action. Further, dose-dependent increases in GH secretion
are observed during acute aerobic exercise in an untrained state (Pritzlaff et al., 1999; figure
5.10) and are enhanced in an endurance-trained state (Weltman et al., 1992). Rather than
directly affecting pulmonary function, GH increases oxygen delivery to exercising muscles
by stimulating erythropoiesis, increasing cardiac output through expansion of PV due to
its antinatriuretic activity (Hoffman et al., 1996), and by stimulating cardiac contractility
and the thickness of the left ventricular posterior wall (Widdowson et al., 2009). Because
increases in PV do not lead to increases in blood pressure, GH appears to reduce total
peripheral resistance by stimulating NO production.
Improved substrate delivery is the result of GH effects on lipid and carbohydrate metabo-
lism. GH contributes to training adaptations by diverting metabolism from oxidation of
carbohydrate and protein to oxidation of FFA (see figure 5.9; Gibney et al., 2007; Møller &
Jorgensen, 2009; Widdowson et al., 2009). It does so mainly by increasing lipolysis threefold
during and after exercise. GH stimulates some of the lipolysis in skeletal muscle (Fryburg et
al., 1991; Rabinowitz & Zierler, 1963) into which the hormone is taken up during exercise
(Brahm et al., 1997) but most of the lipolysis in the adipose tissue. Peak lipolytic action
occurs about 2 h after the onset of acute exercise or GH infusions, reflecting its genomic
action (figure 8.1; Ceseña et al., 2007), and is enhanced by endurance training. GH also
augments the lipolytic action of catecholamines in adipose tissue.
GH secretion is not essential for exercise-associated increase in lipolysis because when
GH secretion is blocked during 30 min of exercise at 70% of maximal effort, no reduction
in FFA mobilization occurs during exercise or 90 min postexercise (Chalmers et al., 1979).
Similarly, administration of recombinant GH does not affect lipid oxidation during or after
exercise (Healy et al., 2006; Lange et al., 2002; Møller & Jorgensen, 2009). Despite its dispens-
ability, enhanced basal and exercise-associated GH secretion in aerobically trained individuals
diverts metabolism from carbohydrate and protein oxidation to oxidation of lipids (Gibney et
al., 2007; Widdowson et al., 2009). GH-induced lipolysis during exercise and after training
has two metabolic consequences. First, it blocks carbohydrate oxidation, in part by causing
insulin resistance in the muscle. GH-induced insulin resistance is achieved without an effect
on PI3K–Akt signaling in the muscle (Jessen et al., 2005). This increases circulating glucose
180 Advanced Exercise Endocrinology

GHR

JAK-2

STAT P13K MAPK

Nucleus

STAT1,3 STAT5 C/EBPβ Elk SRF


c-fos c-fos

sie igf-i c/ebp sre


spi 2.1
cyp2/3

Figure 8.1 Genomic actions of GH. GH can also stimulate c-fos transcription by mediating the binding of CREB to cre.
E4517/Borer/fig.8.1/401613/TB/R3-kh
Adapted from Molecular Genetics and Metabolism, Vol. 90, T.I. Cesena et al., “Multiple mechanisms of growth hormone-regulated gene
transcription,” pgs. 126-133, copyright 2007, with permission from Elsevier.

availability and complements the gluconeogenic action of GH in the liver. The second effect
of lipolysis is the protection of body proteins from being used for oxidation (Nørrelund et
al., 2001). In circumstances in which GH secretion is blocked and lipolysis is significantly
reduced, either acutely through administration of SRIF or chronically in GHD subjects, pro-
tein degradation and urea formation increase by 40% to 50% and body adiposity increases
(Møller & Jorgensen, 2009). Conversely, GH administration to GHD subjects reduces body
fat and increases body protein (Rudman et al., 1990). Increases in whole-body and muscle
protein synthesis seen after both acute endurance exercise (Sheffield-Moore et al., 2004) and
4 mo of endurance training (Short et al., 2004) and decreases in body fatness are very likely
at least partly attributable to GH secretion.
Additional aerobic training adaptations attributed to GH include improved thermoregula-
tion through facilitation by GH of sweat production, conversion of glycolytic to oxidative
muscle fiber types (Klover et al., 2009; Lange et al., 2002; Schuenke et al., 2008; Vescovo
et al., 2005), and enhancement of mitochondrial function by upregulation of oxidative genes
(Short et al., 2008).

Role of Catecholamines
Catecholamines are secreted during acute exercise in a dose-dependent fashion (see figure
3.6) and play a central role in coordinating cardiovascular adjustments and metabolic fuel
delivery that supply exercising muscle with oxygen and metabolites. It is difficult to distin-
guish the direct role of catecholamines in training adaptations from their indirect effects on
tissues during exercise. Training adaptations are not maintained by changes in circulating
catecholamines (figure 8.2) because their response to an absolute exercise workload declines
within 3 wk of high-intensity aerobic training (Winder et al., 1978) but remains unchanged
at the same relative workloads (Deuster et al., 1989).
Catecholamines most likely affect structural and functional adaptations during endur-
ance training by influencing protein turnover. Adrenal E and NE inhibit Ca2+-dependent
protein degradation in oxidative muscles by increasing calpastatin levels. Calpastatin is an
Hormonal Mediation in Training Adaptations 181

endogenous inhibitor of the Ca-dependent Absolute effort Relative effort


proteolytic enzyme calpain, which is found 6
in the muscle and some other tissues. As sug-
gested by its name, which is an amalgamation 4

NE (ng/mI)
of the words calmodulin, papain, and statin,
calpain is very sensitive to small changes in
intracellular Ca levels. Calpastatin activation 2
appears to be mediated by β2 and β3 AdR
activation of cAMP and PKA phosphoryla- 0
tion (Navegantes et al., 2002). Suppression Untrained Trained Untrained Trained
by catecholamines of protein degradation is
most effective under the conditions of nutri-
tional deficiency (Navegantes et al., 2009). E
0.8
infusion also affects expression of the genes
involved in glycogen and lipid metabolism
0.6
and strongly downregulates the genes for
ubiquitin degradation enzymes but does not E (ng/mI) 0.4
affect genes that mediate protein synthesis
(Viguerie et al., 2004). Catecholamines also 0.2
may influence hypertrophy of the left heart
ventricle in response to endurance training 0.0
(Scheuer, 1999) and a shift toward the faster Untrained Trained Untrained Trained
MHC isoform in the ventricles (Gupta, 2007).
Figure 8.2 Changes in plasma catecholamine
concentrations in response to endurance training,
Role of Thyroid Hormones expressed in terms of absolute and relative exercise
E4517/Borer/Fig8.2/401612/alw/R2
The thyroid hormones T4, T3, and rT3, and intensity. Plasma NE (top) and E (bottom) decline after
their pituitary secretagogue TSH, are often 3 wk of high-intensity training when expressed as a
omitted from discussion of adaptations to function of absolute effort in a trained state (solid bars,
left) compared with an untrained state (open bars
endurance exercise because their circulating left). When expressed in terms of relative effort, no
concentrations generally show low-magnitude difference exists in plasma catecholamine responses
or inconsistent responses to aerobic exercise to exercise in highly trained (solid bars, right) and
at intensities between 60% and 80% of maxi- untrained (open bars, right) subjects.
mal effort whether such exercise is of short or Absolute data; adapted, by permission, from W.W. Winder et al.,
long duration (Caralis et al., 1977; Kraemer 1978, “Time course of sympathoadrenal adaptation to endurance
exercise training in man,” Journal of Applied Physiology 45:
et al., 2003; O’Connell et al., 1979; Terjung 370-374; Relative catecholamine data: adapted from Molecular
& Tipton, 1971). When endurance exercise Genetics and Metabolism, Vol. 38, P.A. Deuster et al., “Hormonal
and metabolic responses of untrained, moderately trained, and
is prolonged and leads to significant energy highly trained men to three exercise intensities,” pgs. 141-148,
depletion, T3 concentrations decline and rT3 copyright 1989, with permission from Elsevier.
titers increase because thyroid hormones are
highly responsive to negative energy balance (O’Connell et al., 1979). Despite the apparent
absence of an association between changes in circulating thyroid hormones and endurance
exercise, one can expect that they play a role in adaptations to this type of training for three
reasons: the key role thyroid hormones play in cardiac contractile muscle remodeling toward
increased force production, their contribution to increased arterial compliance, and their
strong stimulatory influence on mitochondrial biogenesis.
The speed and force of heart muscle contraction depends on the properties of its com-
ponent MHC isoforms: the fast αMHC and slow βMHC. Researchers have recognized for
some time that thyroid hormones control the shift toward βMHC isoform during adaptations
182 Advanced Exercise Endocrinology

to pathological hypertrophy that results from chronic hypertension. Now evidence shows
that thyroid hormone is involved in the shift toward αMHC isoform during physiological
left ventricular hypertrophy in response to habitual endurance exercise (Gupta, 2007).
Thyroid hormones bind to both nuclear and mitochondrial TR. In their nuclear action, the
thyroid hormone–TR complex acts as a TF by binding to TRE in the 5' promoter region of
the nuclear MHC genes located on chromosome 14 in humans. After binding to TRα1, a
product of the c-erbA proto-oncogene, thyroid hormones induce the expression of αMHC
genes and repress the expression of βMHC genes in the ventricles, but not the atria, of the
heart. When they bind to TRβ1, they repress the βMHC promoter. By promoting slow-to-
fast MHC conversion in the heart muscle, thyroid hormones antagonize the reverse action
of Cn, which controls the fast-to-slow MHC isoform conversion (Bigard et al., 2008; figure
2.4). In addition to their ability to alter the expression of MHC isoforms, thyroid hormones
can stimulate heart muscle hypertrophy by activating the PKB–Akt–mTOR and Ca−CaM
signaling pathways (Ojmaa, 2010). Thyroid hormones lower peripheral vascular resistance
by attenuating vasoconstriction to catecholamines (Pappas et al., 2009). When T4 is admin-
istered together with a β2 agonist, a reduction in α1 AdR-induced vasoconstriction is not
synergistic, suggesting that thyroid hormones reduce vascular resistance by enhancing β2
adrenergic responsiveness.
The best-known function of thyroid hormones is their ability to regulate metabolism by
acting on the mitochondria in the heart, muscle, liver, and WAT and BAT. In these tissues,
thyroid hormones can both stimulate mitochondrial biogenesis and modulate transcriptional
activation of genes for enzymes involved in OXPHOS, two distinguishing features of periph-
eral aerobic adaptation (Marín-Garcia, 2010). Thyroid hormones also are responsible for
the upregulation of UCP in response to cold exposure, during recovery from exercise, and
during postprandial stimulation of metabolism.
Stimulation of mitochondrial biogenesis is a response to relative cellular hypoxia, mito-
chondrial generation of ROS, and inadequate capacity of the mitochondria for OXPHOS
(Marín-Garcia, 2010). The interdependence and joint actions of systemic hormones and
subcellular factors in mitochondrial biogenesis is evident from the interaction between the
master transcriptional regulator PGC-1α and thyroid hormone activation of this process.
Activation of PGC-1α by muscle contraction induces deiodinase 2, which locally converts
T4 to biologically active T3.
In addition to their actions in the nucleus, T3 can act on mtDNA because the mitochon-
dria possess a truncated version of TRα1. Along with stimulating mitochondrial biogenesis,
thyroid hormones facilitate fatty acid beta oxidation by upregulating OXPHOS enzymes
(Marín-Garcia, 2010). This action is mediated through regulatory factors PPARα and
retinoic X receptor-α, both of which respond to thyroid hormones as well as PGC-1α.
Finally, as discussed in chapter 7, prolonged aerobic exercise training leads to depletion
of glycogen and general energy and triggers a reduction in thyroid hormone secretion and
actions. Global neuroendocrine response to negative energy balance, whether produced by
endurance training or dietary restriction, includes reduced TSH, thyroid hormone and S
stimulation of metabolism, and changes in the secretion of leptin, insulin, and IL-6. These
signals of energy deficit then reduce metabolic energy expenditure (Kelly, 2000; Steinacker
et al., 2005).

Intracellular Signaling Pathways in Endurance Training


The cellular stimuli that elicit cardiorespiratory and mitochondrial adaptations to endurance
training may independently trigger signaling cascades in affected tissues or converge and
synergize with the ones initiated by systemic hormones. Among the tissue changes most
Hormonal Mediation in Training Adaptations 183

likely to contribute to aerobic adaptations are cellular hypoxia, ROS, low cellular energy
charge, and the pattern of Ca release that is characteristic of oxidative slow-twitch skeletal
muscle fibers.
Prolonged aerobic exercise may produce inadequate circulatory delivery of oxygen to
the cardiac and skeletal muscles in case of inadequate myoglobin concentration and thus
produce cellular hypoxia and low creatine phosphate concentrations. LA and hydrogen ions
accumulate through anaerobic metabolism when inadequate capillary supply of oxygen and
mitochondrial volume fail to support oxidative metabolism. Increased ROS release results
from a mismatch between available mitochondria metabolic fuels, and exercise requirement
for OXPHOS. Severe and repetitive hypoxic conditions that can arise from sustained high-
intensity exercise result in the transcription of the hypoxia inducible factor-1α (hif-1α) gene.
The HIF-1α protein upregulates glycolysis, angiogenesis, and mitochondrial biogenesis in
response to low levels of tissue oxygenation (Mason & Johnson, 2007). Cellular hypoxia and
ROS stress also upregulate nuclear respiratory factors (NRF1 and NRF2) that in turn act as
important TF in translating signals of cellular energy demands and stress into mitochondrial
DNA synthesis (see figure 2.4). NRF1 controls the transcription of genes for cytochrome
oxidase (COX) enzymes and heme biosynthesis, whereas NRF2 upregulates genes involved in
ROS detoxification. These TF therefore augment mitochondrial proliferation and respiration,
protein synthesis, and cytochrome content. The complementary actions of thyroid hormone
in heart tissue on mitochondrial biogenesis and expansion of ETC capacity, and of thyroid
hormone on the shift in cardiac MHC isoforms toward faster and more powerful contractions,
result in a sharp increase in myocardial oxygen consumption. In the skeletal muscles, these
exercise conditions stimulate the transcription of genes that code for proteins necessary for
mitochondrial biogenesis, mitochondrial OXPHOS enzymes, angiogenesis, and conversion
of glycolytic muscles to oxidative type (Bassel-Duby & Olson, 2006; Wu et al., 2000).
PGC-1α is the master cellular regulator of gene expression for mitochondrial biogenesis,
Type I muscle transformation, and oxidative metabolism in response to aerobic exercise.
Whereas some systemic hormones such as GH are implicated in the conversion of skeletal
muscle fibers from glycolytic to oxidative types through the elicitation of PGC-1α (Lange
et al., 2002; Schuenke et al., 2008; Vescovo et al., 2005), locally induced PGC-1α during
exercise can induce the conversion in their absence (see figure 2.4). In cooperation with
myocyte enhancer factor 2 (MEF2), PGC-1α controls the transcription of genes for slow
oxidative contractile skeletal muscle proteins, myoglobin, COX, and UCP1. Insulin too can
increase adaptations to endurance training by stimulating MEF2 binding to DNA (Bassel-
Duby & Olson, 2006).
In the muscle, PGC-1α stimulates mitochondrial biogenesis by coactivating NRF1 (see
figure 2.4) and the mitochondrial TF Tfam (not shown) that controls mtDNA replication
and upregulates mitochondrial genes (Joseph et al., 2006). In WAT, PGC-1α upregulates
UCP1 by way of the hormone irisin and mitochondrial cytochromes, conferring to the tissue
thermogenic characteristics of BAT (Lin et al., 2002). Activation by Ras of MAPK can also
mediate the switching of muscle fibers to the slow oxidative type (see figure 5.1; Murgia
et al., 2000). The particular pattern of muscle contractions that characterizes endurance
activities selectively activates AMPK and PGC-1α and suppresses downstream regulators
of protein translation initiation and elongation that lead to muscle hypertrophy (Atherton et
al., 2005). By antagonizing the signaling for hypertrophic growth (see figure 2.8), PGC-1α
and AMPK form the cellular basis for incompatibility of simultaneous aerobic and hyper-
trophic muscle adaptations when both endurance and resistance training modalities are
applied concurrently (Hickson, 1980).
In the muscle, PGC-1α stimulates an increase in expression of fndc5, a membrane protein
that is cleaved and secreted as a newly identified hormone, irisin. Irisin acts on white adipose
184 Advanced Exercise Endocrinology

cells in culture and in vivo to stimulate UCP1 expression and a broad program of brown-
fat-like development. Irisin is induced with exercise in mice and humans, and increases in
irisin levels in the blood cause an increase in thermogenesis in mice with no changes in
locomotion or food intake. This contributes to training-associated declines in body fat and
improvements in insulin sensitivity (Boström et al., 2012).
Muscle and liver glycogen depletion and negative energy balance usually result from
prolonged and habitual endurance activities. The chief enzyme that senses cellular energy
depletion is AMPK (see figure 2.11). It responds to energy cost of exercise by switching off
ATP-consuming biosynthetic pathways and switching on ATP-producing processes (see
chapter 2). In addition to these metabolic actions during acute exercise, AMPK, like PGC-
1α, promotes TF MEF2 DNA binding and thus stimulates transcription of oxidative genes
(see figure 2.4). Other transcriptional regulatory actions of AMPK include upregulation of
COX, UCP3, heme oxygenase-1, HK II, and GLUT-4 genes in the skeletal muscle. AMPK
also increases mitochondrial biogenesis and DNA-binding activity of NRF1 that is neces-
sary for the synthesis of respiratory proteins (Sakamoto & Goodyear, 2002).
Whereas systemic hormones such as thyroid hormone and catecholamines favor the
conversion of slow MHC to the fast isoform in the heart, endurance training enhances the
reverse adaptation from fast to slow MHC in the skeletal muscle. The pattern of motoneuron
discharges through which oxidative fatigue-resistant skeletal muscles are recruited drives
low-amplitude, high-frequency fluctuations in cellular Ca concentration, which in turn trigger
Ca-dependent signaling pathways and conversion of more muscle fibers into an oxidative
phenotype (see figures 2.4 and 2.7). Among the Ca-sensing enzymes are CaM, Cn, and
CaMK. Activation of the genetic program of oxidative muscle fiber is initiated by these
contraction-induced, low-amplitude, high-frequency Ca fluctuations and by Ca binding to
the CaM. This activates the protein phosphatase Cn, dephosphorylation by Cn of Nfat, and
translocation of Nfat to the nucleus. Nfat acts as a TF for transcription of oxidative genes
(see figure 2.4). Thus, CaMK, Cn, AMPK, MEF2, and TF CREB all are affected by Ca that
is released during aerobic exercise and, in turn, activate PGC-1α. Within the 6 h period of
recovery from a bout of aerobic exercise, CaMK activates MEF2 that is necessary, along with
Nfat, for transcription of slow MHC, slow troponin, and myoglobin genes (see figure 2.4).

ROLE OF SYSTEMIC ANABOLIC HORMONES


IN ADAPTATIONS TO RESISTANCE TRAINING
Several systemic hormones—GH, IGF-I, T and other androgens (including anabolic steroids),
catecholamines, PTH, and cortisol—have received attention as potential controllers of muscle
hypertrophy and strength in response to RE. Researchers are interested in GH, IGF-I, and
T and other androgens because they have the known ability to stimulate protein synthesis
and muscle growth, in the catecholamines because they contribute to heart hypertrophy, in
PTH because it can stimulate bone formation as well as resorption, and in cortisol because
of its role in promoting muscle atrophy and bone loss.

Role of GH
In a set of early classical experiments with rats, Arthur Goldberg demonstrated that GH was
not necessary for muscle hypertrophy because ablation of some leg muscles produced the
same percentage of hypertrophy in remaining overloaded muscles in hypophysectomized
rats as it did in rats with intact pituitaries (Goldberg, 1967). A quarter of a century later, the
hypertrophy in hypophysectomized rats was shown to result from direct upregulation of the
Hormonal Mediation in Training Adaptations 185

IGF-I in similarly overloaded skeletal muscles (see figure 2.5; DeVol et al., 1990). Despite
these findings, the question whether GH can contribute to muscle hypertrophy lingers and
fuels the abuse of GH in sport (see “Abusing Hormones for Superpower”). The most likely
reasons for GH doping are the robust GH response during a hypertrophic resistance-training
protocol (see figure 5.19), the well-known role GH plays in musculoskeletal growth during
puberty, and stimulation by GH of muscle protein synthesis after acute administration. GH
infusion into the arm in a postabsorptive state suppresses the release of branched-chain
and essential neutral amino acids from the arm and thus stimulates amino acid retention
and protein synthesis (Fryburg et al., 1991; Rabinowitz & Zierler, 1963). Whether GH can
upregulate the Igf1 gene in the muscle in response to RE remains unresolved. More recent
studies have examined the potential anabolic actions of three GH signaling pathways: the
JAK2 kinase–STAT, the PI3K–Akt, and the Ras–MAPK (Herrington & Carter-Su, 2001;
Lanning & Carter-Su, 2006; see figure 8.1). JAK2–STAT5 upregulates hepatic genes for the
circulating form of IGF-I (Waxman & O’Connor, 2006); the operation of this pathway in the
muscle would support the role of GH in muscle hypertrophy. However, local GH infusion
increased GH anabolic signaling in the muscle and the adipose tissue within 30 min of infu-
sion, but IGF-I expression was observed only in the adipose tissue (Jorgensen et al., 2006).
Although the GH–IGF-I hormonal axis has thus far not been shown to play a major role
in muscle hypertrophy, its actions should be outlined in some detail because the expectation
that it does play a major role keeps fueling exogenous GH administration as a means of
enhancing muscle hypertrophy. All three anabolic signaling pathways mediate facilitation
by GH of protein synthesis, but each also activates transcription of additional genes (figure
8.1). Of the STAT TFs that are recruited and phosphorylated by JAK2, STAT1 dimers are
involved in the upregulation of immune cells, STAT3 dimers are involved in the regulation
of cellular growth and its dysregulation in cancer (both by way of sie, the SIS-inducible
gene element), and STAT5 dimers are involved in the expression of the hepatic circulat-
ing form of IGF-I and hepatic Cyp 2/3 genes involved in the transcription of sex-specific,
steroid-metabolizing cytochrome P450 enzymes (see chapter 7; Waxman & O’Connor, 2006).
Genomic actions of GH involve the regulation of several TF involved in the expression of
the proto-oncogene c-fos. These include Elk, C/EBP β, and CREBP, which act by binding
to their respective HRE. c-fos dimerizes with c-jun to form AP-1, which in turn acts as a
TF to regulate the genes involved in cell proliferation and differentiation. The Ras–MAPK
GH signaling pathway also activates the c-fos (figure 8.1) and c-myc genes (not shown) and
through AP-1 promotes cell proliferation and differentiation. The MAPK pathway involves
sequential phosphorylations that activate Raf, MEK1/2, and ERK1/2 kinases (figure 2.13).
ERKs translocate to the nucleus and phosphorylate the TF c-myc and Elk-1 (Force & Bon-
ventre, 1998). ERKs stimulate protein translation through activation of P90RSK and eIF4E
(see figures 2.1, 2.7, and 2.8). The PI3K–Akt cascade activated by GH involves recruitment
by JAK2 of the IRS and its downstream kinases. It activates mTOR and is likely to be active
during the late postprandial period when both insulin and GH are relatively high, when
GH has transient, insulin-like actions, and when both hormones may synergize in nutri-
ent uptake and protein synthesis (Rabinowitz & Zierler, 1963). Stimulation of lipolysis by
GH involves the upregulation of lipolytic enzymes (figure 5.9) rather than PKA, PKG, or
PKC (see figure 5.7). GH may reduce abdominal and whole body fat through inhibition of
adipogenic enzyme 11βHSD type 1 and PPARγ gene expression in adipose tissue, thereby
preventing adipocyte differentiation and lipid accumulation, particularly in visceral WAT
(Herrington & Carter-Su, 2001; Zhao et al., 2011).
The temporal dissociation between GH and IGF-I secretion during RE and training
further supports the conclusion that GH does not play a direct role in muscle hypertrophy.
186 Advanced Exercise Endocrinology

GH responds strongly to both acute aerobic exercise (see figure 5.10) and a high volume of
moderate-intensity (10RM) RE (see figure 5.19; Kraemer et al., 1990, Kraemer & Ratamess,
2005). However, these acute-exercise parameters produce little change in IGF-I or its
GH-stimulated BP, IGFBP-3 (Manetta et al., 2002). Moreover, IGF-I concentration often
declines during the early weeks of exercise training (Rarick et al., 2007), does not respond
to acute weightlifting, and becomes elevated only after longer exposure to training (Nindl
& Pierce, 2010). For instance, increases in IGF-I levels become detectable only after 2 wk
of training and stay elevated for at least 6 mo (Roelen et al., 1997). Large increases in IGF-I
and IGFBP-3 were also seen after 4 mo of training (Koziris et al., 1999). However, no clear
evidence exists that increased secretion of GH during a hypertrophic resistance-training
protocol leads to skeletal muscle hypertrophy. GH does stimulate collagen synthesis in con-
nective tissues and tendons, contributing to their resistance to damage after training. This
partly justifies the popularity of GH as an anabolic agent in sport (Doessing & Kjaer, 2005).

Role of IGF-I
The hypertrophic responses of skeletal muscle to RE more likely reflect the expression in
skeletal muscle of IGF-I and some other growth factors in response to high-resistance muscle
contractions rather than anabolic effects of circulating IGF-I. Evidence in support of local
anabolic adaptations includes increases in IGF-I and IGF-II mRNA in the mechanically
loaded muscles of hypophysectomized rats (DeVol et al., 1990; figure 2.5), increases in IGF-I
mRNA expression in human muscle in response to exercise training unaccompanied by an
increase in circulating IGF-I (Hambrecht et al., 2005), and normal muscle hypertrophy in
resistance-trained mice in which the hepatic IGF-I gene has been knocked out (Matheny
et al., 2009).
Hepatic IGF-I differs from the IGF-I expressed in skeletal muscle because mRNAs
of the two isoforms are spliced differently from 6 exons of the Igf1 gene (Matheny et al.,
2010). The hepatic IGF-I and other IGF-I variants contain protein segments coded by exons
3 and 4, whereas the variant IGF-IEa in addition expresses exon 6 and the human IGF-IEb
variant expresses exon 5. Mechano growth factor (MGF), the variant that is upregulated in
rodents by muscle overload (Goldspink, 2005), resembles IGF-IEb but also expresses exon
6. The IGF-I variants are thought to have distinct roles in the hypertrophic proliferation and
differentiation of satellite cells (discussed in the section “Role of Satellite Cells in Skeletal
Muscle Hypertrophy”).

Role of Androgens
RE that varies in volume and intensity and engages a large muscle mass appears to uniformly
produce robust increases in plasma T concentrations in men (see figure 5.20; Kraemer et al.,
1990) but small or inconsistent increases in women (Kraemer & Ratamess, 2005; Vingren
et al., 2009). Chapter 7 briefly outlines the capacity of T to significantly stimulate muscle
protein synthesis and hypertrophy. Anabolic actions of T are mediated by T binding to ARs
and appear to be additive to signaling that is triggered directly by mechanical loading during
RE (Bhasin et al., 1996). This suggests that different cellular mechanisms mediate protein
synthesis from the two stimuli. Resistance training does not alter resting plasma T concen-
trations. Instead, research has focused on possible training-induced changes in AR number.
ARs are unevenly distributed among skeletal muscles in males, where they are more
prevalent than they are in females. In men, the number of ARs is significantly higher in
the trapezius muscle that supports the arms than in the leg muscle vastus lateralis, which is
typically sampled in most exercise studies. The number of ARs in the trapezius is increased
Hormonal Mediation in Training Adaptations 187

in proportion to the level of resistance training and exposure to anabolic steroids, whereas
no such effect is seen in the vastus lateralis (Kadi et al., 2000). The volume of mechanical
loading and its catabolic consequences influence T and AR response to RE. A single set of
lifting at 10RM produces an inconsistent increase in plasma T or change in AR. One hour
after 6 sets of 10 weightlifting repetitions, a 20% increase in T response, an approximately
40% increase in plasma cortisol, and a 46% decrease in AR number occur (Ratamess et
al., 2005). After a series of exercise bouts, T response progressively increases, and 48 h
postexercise a delayed increase in AR expression occurs (Willoughby & Taylor, 2004). Pro-
tein–carbohydrate supplementation before and after training blunts the magnitude of initial
AR decline (Kraemer & Ratamess, 2005). Thus, a threshold volume of RE that produces
a metabolic demand is necessary to elicit a T response and an early decline in AR 1 to 3 h
postexercise followed by an increase in AR 2 d after RE.
Suppression of T secretion by 12 wk of GnRH antagonist administration does not blunt
mRNA expression of myogenic factors MyoD, myogenin, IGF-IE isoforms, or ARs or inhibit
the expression of myostatin in response to 8 wk of strength training. However, presence
of T along with resistance training was necessary for stimulating hypertrophic growth,
increases in strength, and reduction in percentage of body fat (Kvorning et al., 2006, 2007).
The data suggest that the T–AR complex requires the upregulation of IGF-IEb–MGF during
skeletal muscle loading to produce hypertrophy, although after supraphysiological doses of
T, hypertrophic mechanisms of exercise and the hormone appeared to be separate (Bhasin
et al., 1996). Hypertrophic action of T is mainly mediated by skeletal muscle satellite
cells. However, dual actions of androgens in facilitating skeletal muscle hypertrophy and
reduction in body fat suggest that androgenic stimulation of hypertrophic muscle growth
may also occur by directing the pluripotent MSC toward myogenesis and inhibiting MSC
differentiation into the adipogenic lineage (Bhasin et al., 2003).

Role of Satellite Cells in Skeletal Muscle Hypertrophy


Satellite cells are undifferentiated embryonic myoblasts with a large nucleus:cytoplasm
ratio that are located between the muscle cell membrane and basal lamina (the layer of
extracellular matrix secreted by cells that line the sarcolemma; Hawke & Garry, 2001;
Mauro 1961). The surface molecule integrin α7β1 binds the satellite cell cytoskeleton to the
laminin in the basal lamina. The satellite cells attach to the underlying myofiber through
M-cadherin. Satellite cells actively participate in intrauterine and postnatal muscle growth
but are quiescent during adulthood as long as they are anchored to the basal lamina and
sarcolemma, and proteins in the extracellular matrix such as TGF-β restrain the activity of
some activating growth factors and hypertrophic growth.
A variety of stimuli activate satellite cells to proliferate and differentiate into myoblasts
that then become incorporated into mature myofibers as nuclei during hypertrophic growth
(figure 8.3). Activating stimuli include muscle overloading during RE, tissue events associ-
ated with muscle damage, systemic hormones such as T, and autocrine and paracrine growth
factors such as IGF-I variants, MGF, EGF, and HGF as well as NO (Matheny et al., 2010).
T activates satellite cells by binding to ARs that are abundantly expressed in these cells
(Sinha-Hikim et al., 2004) and induces satellite cell proliferation and muscle hypertrophy
in older men in a dose-dependent fashion (Sinha-Hikim et al., 2006). Following Igf1 gene
transcription in the skeletal muscle, the IGF-I variants are transported in interstitial fluid to
the cell surface to bind to IGF-I receptors in an autocrine fashion (figure 1.3; Adams, 2002).
IGF-I receptors, like IR (see figures 1.11b, 5.1), have an intrinsic tyrosine kinase that gets
autophosphorylated after binding with the hormone and then attracts and phosphorylates
188 Advanced Exercise Endocrinology

Activation Proliferation
MB
MB MB
Differentiation
Basal lamina
SC

MB
N
Incorporation

Sarcolemma

Figure 8.3 The role of satellite cells in muscle hypertrophy. Muscle overload, T, and a number of paracrine
messengers activate satellite cells to proliferate and differentiate into myoblasts (MB) that then become incorporated
into myofibers as nuclei, allowing additional contractile protein synthesis.
E4517/Borer/fig.8.3/401614/TB/R2-alw

docking proteins IRS and Src. IRS initiates PI3K–Akt signaling and Src activates the MAPK
pathway to increase satellite cell proliferation.
The presence of multiple IGF-I variants helps explain opposing actions of IGF-I in the
hypertrophic process (Clemmons, 2009; Hill & Goldspink, 2003). An early prolifera-
tive effect of IGF-I on satellite cells is a component of satellite cell activation and is seen
through an increase in M-cadherin mRNA while differentiation into myoblast and associ-
ated myogenin expression are blocked. This proliferative phase is attributed to IGF-IEb
and MGF variants and is characterized by MAPK signaling. Satellite cell activation also
results from microfiber microdamage and rupture of basal lamina during RE. This allows
an influx of macrophages into the myofiber that release proinflammatory cytokines. The
macrophages phagocytize necrotic proteins; the TNFα, IL-1β, and granulocyte monocyte
colony-stimulating factor that they release are necessary for hypertrophic adaptation (Ten
Broek et al., 2010). HGF, FGF, and PDGF released from the extracellular matrix by MMPs
contribute to muscle remodeling, whereas VEGF stimulates angiogenesis. These growth
factors activate satellite cells through upregulation of the TF Pax7 and the MRF MyoD
and Myf5. Hypertrophic activation of satellite cells also occurs through the release from
the endings of motoneurons of neurotransmitters and NTFs such as calcitonin gene-related
peptide, cilitary NTF, and neuroregulin (Sakamoto & Goodyear, 2002).
Whereas a blockade of PI3K signaling facilitates satellite cell proliferation and inhibits
satellite cell differentiation, the blockade of the MAPK pathway prevents the IGF-I prolif-
erative effect and initiates satellite cell differentiation and myogenin expression (Adi et al.,
2002; Coolican et al., 1997). Satellite cell differentiation into myoblasts and hypertrophic
synthesis of contractile proteins is mediated by the PI3K–mTOR signaling pathway and is
attributed to delayed action of IGF-IEa and anti-inflammatory cytokines such as IL-10. It
also involves upregulation of Mrf4 and myogenin and downregulation of Pax7. The process of
satellite cell fusion into myofibers is less understood. The addition of myonuclei derived from
satellite cells expands the capacity of muscle to synthesize additional contractile proteins.
Muscle hypertrophy can be hampered by members of TGFα and TGFβ and BMP and
by the negative cell-growth mediator myostatin (figures 2.6 and 8.5). TGFβ interferes
with hypertrophy by way of a Smad protein that binds to response elements that affect its
gene transcription. It also can lead to scar formation by stimulating fibroblasts to produce
collagen and fibronectin. Myostatin is expressed in satellite cells and myoblasts. It blocks
Hormonal Mediation in Training Adaptations 189

muscle hypertrophy by downregulating MRF Pax3 and Myf5 and by blocking expression
of MyoD (Ten Broek et al., 2010).

Role of Catecholamines
Catecholamines stimulate ultradian rhythms and frequency and force of heart contraction
(see chapter 3), the primary hemodynamic stimuli of exercise-induced cardiac remodeling
and physiological heart hypertrophy (Weiner et al., 2012). NE acting on α1 AdRs (figure
1.14), angiotensin II acting on AT-1 Rs, and endothelin acting on ET receptors all contribute
to pathological cardiac hypertrophy in response to sustained hypertension. Activation of
α1 AdRs and the Gq signaling causes pathological cardiac hypertrophy and heart failure in
several ways, by activating gamma isoform of PI3K and inactivating its alpha isoform that
reacts to growth factors, activating protein kinase C alpha and beta isoforms, and activat-
ing Cn/NFAT signaling (Heineke & Molkentin, 2006). Physiological hypertrophy requires
sensing biomechanical stress signals through an integrin and Z-disc sensory apparatus (see
chapter 2), the activation by IGF-I and some other growth factors of protein synthesis via
IP3–Akt–mTOR signaling pathway (figures 2.1 and 2.6), and suppression of GSK3β actions
(Dorn & Force, 2005; figures 2.1, 2.7, and 2.8).
In addition to the rapid and acute enzymatic action of catecholamines in promoting
fuel mobilization and suppressing protein degradation, E infusion affects expression of a
number of skeletal muscle genes that could contribute to the metabolic and antiproteolytic
effects of endurance training in humans. Of the 30 responsive genes, those promoting gly-
cogen synthesis were downregulated, and the gene for the GSK3 enzyme was upregulated.
E infusion induced genes for glycogen phosphorylase, for several enzymes facilitating
glycolysis, and for some apolipoproteins, whereas it downregulated several lipid transport
genes (Viguerie et al., 2004).

Role of PTH in Bone Formation and Resorption


Activation by PGE2 of an osteogenic response to mechanical loading of the bone is described
in chapter 2. The same osteogenic response is triggered by intermittent secretion of PTH
(Jilka, 2007) elicited by bouts of exercise. Both mechanical loading and intermittent PTH
pulses increase osteoblast numbers in cancellous bone and suppress their apoptosis, thereby
increasing bone formation and mineralization. PTH actions can be mediated through both Gs
and Gq signaling, but only the activation of cAMP by Gs produces an osteogenic effect. The
effect is the result of the transient increase in the expression of runt-related TF-2 (Runx2) and
the transient suppression of cyclin D1, a protein that controls the cell cycle but also targets
Runx2 for proteasomal degradation (figure 8.4). Runx2 activates short bursts of survival sig-
naling through the expression of survival genes such as Bcl-2 (see figure 2.16). By decreasing
the expression of cyclin D and increasing the expression of CDK inhibitors, pulsatile PTH
facilitates osteogenic differentiation of osteoblasts, assisted by locally produced growth factors
and cytokines, including IGF-I, FGF-2, and Wnt. Concurrently, this action of PTH reduces
the adipogenic differentiation of mesenchymal osteoblast progenitors in bone marrow through
attenuation of the adipogenic actions of PPARγ. With continuous PTH elevation, or as PTH
disappears from circulation, Runx2 returns to basal levels below the threshold that is needed
for survival gene signaling and is subject to proteasomal degradation. The antiapoptotic
effect of a PTH pulse lasts only about 6 h due to proteasomal degradation of Runx2 via the
ubiquitin protein ligase Smurf1. A decline in Runx2 allows the expression of apoptotic genes
such as Bad, bax and the apoptosis of osteoblasts. The insight into the details of osteogenic
responses to pulsatile PTH and intermittent bone loading helps explain why increases in
rodent bone mass are twice as large when such loading is separated by 8 h as it is when it
190 Advanced Exercise Endocrinology

PTH is delivered in a single bout (Robling et al.,


2001) and why daily injections of PTH analog
teriparatide stimulate bone mineral deposition
Proteasomal (Han & Wan, 2012) whereas chronic PTH
degradation elevation (secondary hyperparathyroidism) in
cAMP some postmenopausal women produces bone
Bad mineral losses.
PKA
SMURF1 CREB 14
Bad P 3
Role of Cortisol in Muscle
Runx2
3 Atrophy and Bone Resorption
P CREB Survival Bcl-2 Bcl-2 Given that cortisol leads to muscle atrophy,
genes researchers are interested in its secretory
pattern during RE and its relationship to hor-
p21 mones that promote skeletal muscle hypertro-
phy such as T and IGF-I. A hypertrophic RE
protocol increases plasma concentrations of
Antiapoptosis the pituitary corticotropin ACTH, more so
in untrained young men than in older men,
and increases concentrations of cortisol to the
PTH PTH PTH PTH
Antiapoptotic

same extent in both age groups. After 10 wk


signals

of periodized strength training, resting cor-


tisol and acute increases in its concentration
during heavy RE declined, more so in young
men than in older men (Kraemer et al., 1990).
Runx2

Determining the possible role of cortisol


in training adaptations is complicated by the
pleiotropic nature of this hormone. Cortisol is
Figure 8.4 Stimulation of bone formation by the final hormone in the endocrine secretory
pulsatile PTH secretion and of bone resorption by response to physiological, environmental, and
E4517/Borer/fig.8.4/401615/TB/R3-alw
unchanging PTH concentration. Pulsatile secretion psychological stress. Hypothalamic CRH
of PTH (bottom left) activates cAMP–PKA signal- triggers the release of ACTH, which then
ing and increases the concentration of Runx2 TF stimulates the synthesis and release of corti-
(top left). This activates expression of antiapoptotic sol from the adrenal cortical zona fascicularis
genes in the osteoblasts, including Bcl-2, for about
(figure 3.18). Lipid-soluble cortisol diffuses
6 h (top left). Absence of subsequent change in PTH
concentration (bottom right) leads to proteasomal from circulation into the cells, where it binds
degradation of Runx2 and expression of proapoptotic to dimerized GR to orchestrate changes in
genes, including Bad. protein, carbohydrate, and lipid metabolism;
Reprinted, by permission, from R.L. Jilka, 2007, “Molecular and the immune system; vascular tone; bone min-
cellular mechanisms of the anabolic effect of intermittent PTH,” eralization; and brain function. The GR gene
Bone 40: 1434-1446.
is located on chromosome 5 and is spliced as
variants GRα and GRβ. Cortisol-glucocorticoid receptor alpha complex (CGRα) exerts the
known hormone effects, whereas the CGRβ complex interferes with GRα binding to the GRE
responsive to cortisol (Kino et al., 2009). After binding to a GRE, CGRα recruits coactiva-
tor proteins to initiate gene transcription. Its gluconeogenic actions lead to transcription of
the enzymes PEPCK and tyrosine and alanine aminotransferases (Revollo & Cidlowski,
2009). Gene repression or silencing in the liver occurs when CGRα interacts with negative
glucocorticoid-responsive element on a gene (nGRE). GRα gene silencing can occur by way
of protein–protein interactions independent of CGR binding to DNA. To that end, GRα has
Hormonal Mediation in Training Adaptations 191

two transactivation domains: AF-1 in the HVD and AF-2 in the hormone or ligand-binding
domain (HBD). Through AF, the GR interacts with various coactivator proteins.

Effects of Cortisol on Skeletal Muscle


Glucocorticoids contribute to muscle atrophy, particularly of Type IIx and Type IIb muscle
fibers, by both inhibiting protein synthesis and increasing protein degradation (figure 8.5;
Schakman et al., 2009). Synthesis is indirectly inhibited by CGRα suppression of enzymes
that are responsible for amino acid transport into the muscle (alanine and tyrosine ami-
notransferases). Cortisol directly inhibits protein synthesis in the muscle and other tissues
in several ways. First, CGRα blocks action of the TF STAT5 downstream of GH binding
to its receptor (figure 8.1). Thus CGRα can repress GH action that includes expression of
many protein genes via the TF AP-1. Second, CGRα blocks stimulatory actions of insulin,
IGF-I, and amino acids (particularly leucine) on protein translation via phosphorylation of
eIF4E-BP and p70S6K (figures 2.1, 2.7, and 2.8). By suppressing PI3K signaling of insulin
and IGF-I, p70S6K phosphorylation is reduced. This represses mTOR via expression of
REDD1 and reduces TF ATF-4, both of which are necessary for uptake of essential amino

IGF-1

C Mstn AA

PIP2 PIP3

P85 P110
ATF-4
PI3K

CGRα
REDD1 Akt/PKB

mTOR FOXO GSK3β

Protein Proteolysis
synthesis

Nucleus

CGRα
CGRα IL-1β
IGF-1

STAT5
CGRα OCN CGRα Mstn
CGRα ATF-4

CGRα REDD1

Figure 8.5 Genomic actions of cortisol leading to protein degradation. Cortisol bound to GRα attaches to GRE
in the promoter region of cortisol-sensitive proteins to block their transcription.
E4517/Borer/fig.8.5/401616/TB/R3-kh
192 Advanced Exercise Endocrinology

acid and synthesis of nonessential amino acids (figure 8.5). By inhibiting Akt, CGRα also
increases GSK3β activity, which promotes ubiquitinization of the transcriptional activator
β-catenin and increases the expression of the FOXO gene that regulates apoptotic genes
atrogin-1, MuRF-1, cathepsin-L, PDK4, and 4E-BP-1 (Schakman et al., 2009). Third,
CGRα interaction with NF-κB produces an immunosuppressive action by inhibiting protein
synthesis. As illustrated in figure 2.14, NF-κB is activated when inflammatory cytokines
cause its phosphorylation by IκK kinase. NF-κB then translocates to the nucleus, where
GRα blocks its transcription of genes for proinflammatory cytokines and their receptors
(e.g., TNFα, IL-1β, and granulocyte monocyte colony-stimulating factor) by binding to its
p65 component protein. Fourth, similar to its action on NF-κB, CGRα also inhibits MAPK
signaling for mitosis by promoting the expression of the gene for MAPK phosphatase.
Fifth, myogenesis is also inhibited when CGRα downregulate myogenin and MyoD TF
that are involved in the differentiation and incorporation of satellite cells into muscle fibers
and upregulate the negative TF myostatin (figure 8.5). Protein degradation is initiated by
activation by CGRα of several cellular proteolytic systems, including the ubiquitin protea-
some system (see figure 2.14), cathepsin lysosomal system, caspase-3 (see figure 2.16), and
Ca-dependent calpains. Calpains are thought to dissociate actin and myosin to make them
accessible for ubiquitination.

Effects of Cortisol on Bone


Cortisol can dramatically reduce bone density and increase the risk of bone fracture in
circumstances in which the hormone is oversecreted (e.g., Cushing’s disease). It contributes
to the decline in bone density with aging by converting cortisone to cortisol by 11βHSD1
in the osteoblasts (Hardy & Cooper, 2010). Glucocorticoids block bone formation and
mineralization by similar mechanisms through which they cause skeletal muscle atrophy
(Canalis & Delany, 2002). The suppression by cortisol of bone mineralization involves GRα
binding to the nGRE in the osteocalcin gene promoter (figure 8.5). GRα binding to the
nGRE in the osteocalcin gene promoter reduces osteoblast synthesis and affects receptor
binding and BPs of anabolic growth factors. In particular, this action decreases the number
of osteoblasts by impairing osteoclast proliferation and differentiation and promoting
their apoptosis. Two anabolic growth factors that are particularly blocked by cortisol are
IGF-I (figure 8.5) and TGFβ. TGFβ stimulates the synthesis of collagen and bone matrix
proteins whereas the glucocorticoids activate collagenases (MMP enzymes that cleave
collagen, which is the major component of bone matrix). The cortisol block of osteoblast
proliferation by IGF action is mediated by the C/EBP TF and by suppression of IGFBP-5.
The growth factors FGF-2 and PDGF (which accelerate fracture repair) and HGF (which
promotes tissue regeneration and repair) are also suppressed by cortisol. Glucocorticoids
also increase osteoclastogenesis by increasing the expression of RANKL and decreasing
the expression of its decoy receptor, OPG. As described in chapter 2, RANKL is produced
by osteoblasts to stimulate, along with colony-stimulating factor, osteoclastogenesis (figure
7.8). By suppressing the synthesis of OPG and its binding with RANKL, cortisol facilitates
the formation of bone-resorbing osteoclasts.

NUTRITIONAL MODULATION OF HORMONAL


ADAPTATIONS TO EXERCISE TRAINING
Most training adaptations require tissue remodeling, and resistance training causes substan-
tial increases in skeletal muscle mass. These adaptations therefore require adequate supply of
nutrient energy and, in particular, the availability of essential amino acids in the diet to pro-
duce protein synthesis. Because exercise acutely depletes cellular energy and energy stores,
Hormonal Mediation in Training Adaptations 193

researchers are interested in examining 300 600


how it interacts with ingested nutrients
to produce training adaptations. After

Cortisol (nM)
Insulin (pM)
200 400
an overnight fast, protein degradation
exceeds protein synthesis in the rest-
ing and exercised muscle, although 100 200
prior exercise increases the rate of
both processes. If an amino acid meal
0 0
is provided, protein synthesis exceeds 6 a.m. 10 2 6 p.m. 10 2 6 a.m.
protein degradation; the magnitude a Clock time
of the response is greater 3 h after
exercise. Finally, adding glucose to an 9
E4517/Borer/Fig8.6a/401651/alw/R2
amino acid meal increases the protein

Cortisol/Insulin (× 1,000)
synthetic rate above that of protein
degradation in nonexercise condition 6
and even more so postexercise, likely
because of greater insulin response to
3
the combination of the two nutrients
(Rasmussen & Phillips, 2003).
Muscle hypertrophy is clearly
0
associated with the anabolic actions 6 a.m. 10 2 6 p.m. 10 2 6 a.m.
of insulin and IGF-I, both of which are b Clock time
secreted in response to nutrient intake.
Figure 8.6 Postprandial periods when cortisol:insulin
Whereas exercise acutely suppresses
ratios favorE4517/Borer/Fig8.6b/401617/alw/R3-kh
protein synthesis. Postprandial increases in
secretion of insulin but not of IGF-I, insulin and meal-associated and circadian changes in cortisol
postexercise meals increase secretion secretion (a) interact to produce 3 to 7 h postmeal periods
of both hormones and facilitate their when the cortisol:insulin ratio is less than 4 (horizontal line
anabolic effects, which are mediated in b) and favorable for protein synthesis.
through the already-discussed insulin Reprinted, by permission, from A.B. Loucks, 2006, The endocrine system:
and IGF-I–PI3K–Akt signaling path- Integrated influences on metabolism, growth, and reproduction. In ACSM’s
advanced exercise physiology, edited by C.M. Tipton et al. (Philadelphia:
ways (see figures 2.1 and 5.1). Lippincott, Williams & Wilkins), 464; Data modified from E. Van Cauter,
The possible negative role of cortisol et al., 1992, “Circadian modulation of glucose and insulin responses to
meals: relationship to cortisol rhythm,” American Journal of Physiology
in adaptations to both endurance and 262: E467-E475.
resistance training has attracted a great
deal of interest among scientists and sport coaches. T:cortisol ratio is often a focus of interest
in both endurance and resistance training (Loucks, 2006), but it has not turned out to be a
useful predictor of hypertrophic response (Kraemer & Ratamess, 2005). Insulin stimulates
protein synthesis and prevents protein degradation at concentrations greater than 25 µU or
150 pM; in the presence of cortisol, cortisol:insulin ratio has to be less than 4 to produce the
same outcome. Using meal-associated values for concentrations of insulin and cortisol in
nonexercising humans, the optimal times for stimulating muscle protein synthesis are 3 to
7 h-long postprandial periods when cortisol:insulin ratio is less than 4 (figure 8.6; Loucks,
2006). This explains the efficacy of postexercise amino acid and glucose supplementation
in increasing protein synthesis (Rasmussen & Phillips, 2003). Other studies show that IGF-I
(Schakman et al., 2009) or amino acid (Paddon-Jones et al., 2005) supplementation inhibit
cortisol-induced protein degradation.

Tissue Sensing of Energy Depletion


Many tissues, including skeletal muscle, the liver, the pancreas, and the brain, can detect
the decline in cellular energy availability caused by exercise. Short-term signals of energy
194 Advanced Exercise Endocrinology

shortage elicit the secretion of the counterregulatory hormones glucagon, the catecholamines,
cortisol, GH, and ANP, the function of which is to mobilize metabolic fuels for use by the
exercising muscles and the brain. Concurrently, the catecholamines suppress secretion of
insulin and its antagonism of metabolic fuel mobilization. (These endocrine responses are
discussed in chapter 5.) Endurance training often entails large energy drain that is carried
out over prolonged periods of time and stimulates oversecretion of some counterregulatory
hormones. This type of exercise training is frequently associated with reduced adipose tissue
lipid stores. An acute decline in energy balance triggers declines in plasma leptin within
several hours of exercise energy expenditure (Borer et al., 2009), and pulsatile GH secre-
tion increases after several days of caloric restriction (Ho et al., 1992). Declines in basal
insulin, leptin, IGF-I, and T3, the active form of thyroid hormone, are proportional to a 4 d
reduction in energy availability below 35 kcal/kg of fat-free mass, whether this reduction
is produced by exercise or dietary restriction (figure 7.18, b through d; Loucks & Thuma,
2003). Reduction in body fat mass by exercise or dietary restriction leads to proportional
reduction in basal insulin, leptin, IGF-I, and T3 (Baylor & Hackney, 2003; Considine, 2005;
Hickey et al., 1996; Nindl & Pierce, 2010). Collectively and individually, changes in the
concentrations or actions of these systemic hormones reflect energy store depletion and
interfere with protein synthesis and energy-costly processes such as growth and reproduc-
tive function (Ahima et al., 1999; Loucks & Thuma, 2003).
Many tissues can sense energy shortage and trigger transduction pathways that achieve
energy mobilization and utilization and decrease energy-costly biosynthetic processes.
A decrease in plasma glucose concentration elicits glucagon release from pancreatic islet
cells in vitro. Chapter 2 discusses the activation of the energy-sensing enzyme AMPK by
increased concentration of AMP and a decline in ATP in the skeletal muscle and other
tissues. Activated AMPK stimulates glucose uptake, lipid oxidation, and mitochondrial
biogenesis and upregulates metabolic genes (Long & Zierath, 2008). Besides activating
energy-producing processes, AMPK blocks protein synthesis by stimulating the activity of
TSC in the PI3K–Akt–mTOR pathway. This prevents mTOR activation through the inhibi-
tion of Rheb (see figure 2.8).

Tissue Sensing of Energy Abundance


Adequate nutritional support is necessary to mitigate the protein degradation that occurs
during acute exercise (Rennie et al., 2006). Thus, how nutrient ingestion during or after
training makes protein synthesis and hypertrophic growth possible is of relevance to train-
ing adaptations. The systemic hormones again play an important role in that food intake
stimulates the release of the master metabolic regulator, insulin (see chapter 5). Insulin
secretion increases in response to absorbed glucose and amino acids in circulation and
declines as its job of promoting tissue uptake of these nutrients concludes. Concentrations of
the gut peptides GIP, GLP-1, and PYY also increase in the course of nutrient digestion and
decline when it ends, whereas ghrelin increases before a meal and declines when nutrients
are ingested. RE upregulates the expression of IGF-I in the muscle, and both insulin and
IGF-I initiate protein synthesis through the PI3K–Akt–mTOR pathway. The key downstream
component of this pathway is the TORC1 (Akt-stimulated mammalian target of rapamy-
cin complex), which consists of an mTOR enzyme, a scaffolding protein raptor, and Akt
substrate PRAS40 (figure 8.7). TORC1 integrates signals from insulin, IGF-I, mechanical
stress, and energy-sensing enzymes in muscle. It can respond to mechanical loading even
in the absence of IGF-I receptors (Spangenburg et al., 2006). Its principal role is to initiate
protein translation.
Hormonal Mediation in Training Adaptations 195

Systemic nutrient sensing IGF/Insulin

InR

TORC2 PIP3 PIP3 PIP2


Akt PDK-1 P13K IRS

AMP/ATP Glycogen

TSC1 TSC2 AMPK LKB1

Energy sensing

Rheb

RagA/B TORC1
MAP4K3
RagC/D
Growth

Amino acid sensing

Figure 8.7 Cellular signaling pathways sensitive to systemic hormones, intracellular nutrients, and amino acids.
Insulin and IGF-I stimulate TORC1-mediatedE4517/Borer/fig.8.7/401618/TB/R2-alw
growth through the activation of the PI3K–Akt pathway, and branched-
chain amino acids do so through the binding of RAG GTPases to TORC1. AMPK senses declines in cellular energy
charge and degree of glycogen repletion and can inhibit TORC1 directly or via TSC1/2.
Reprinted, by permission, from V. Hietakangas and S.M. Cohen, 2009, “Regulation of tissue growth through nutrient sensing,” Annual Review
of Genetics 43: 389-410.

Activation of the Akt–mTORC1 pathway is achieved in two ways. Akt directly removes
the inhibitory action of PRAS40 over mTOR. It also blocks the inhibitory action of TSC1
and TSC2 over a GTP BP, Rheb. Freed of inhibitory regulation by TSC1/2, Rheb activates
mTOR (Gibbons et al., 2009). The main downstream targets of mTOR are components of
protein translation machinery, including 4E-BP and 40S ribosomal protein S6 kinases, both
of which are necessary to initiate protein translation (Ma & Blenis, 2009). Phosphoryla-
tion by mTOR of 4E-BP frees up eIF4E, eIF4B, and eIF4G to associate with the initiation
ribosomal components. Akt also phosphorylates the p70S6K pathway, which contributes
both to translation initiation by activating eIF4B and protein chain elongation by activating
eEF2 (see figures 2.1, 2.7, and 2.8). In addition to participating in protein translation, the
Akt–mTOR pathway regulates RNA polymerase I-mediated synthesis of ribosomal RNA.
TORC1 stimulation of protein synthesis and growth is blocked by AMPK when this enzyme
senses energy shortage, which in turn stimulates TSC1/2 to restrain mTORC1 activity
(Hietakangas & Cohen, 2009; figure 2.8). On the other hand, muscle glycogen repletion
blocks the inhibitory influence of AMPK over the Akt–mTOR pathway. In addition to the
PI3K signaling pathway, MAPK also contributes to nutrient-induced response to exercise
training. After phosphorylation, the IRS proteins bind to proteins that contain SH2 domains
and include PI3K and the adaptor molecule Grb, which activates the Ras–MAPK pathway.
Of the two isoforms, IRS-1 more closely regulates glucose uptake and IRS-2 is more closely
196
Removal of Esterification confers
Attachment of methyl group the angular depot activity for
methyl group intramuscular administration
Introduction of double bond
OH
Attachment of various Attachment of 17 α-alkyl
17
groups at C-2 C D group confers oral activity
1
2
Attachment of pyrazole A B
3 4 7
ring to the A-ring O Attachment of 7 α-methyl group

Attachment of chlorine
or hydroxyl group

Figure 8.8 Modifications of the T molecule in anabolic steroids. Type A, B, and C modifications
Borer/E4517/Fig.
of the T molecule entail esterification of 8.8/401619/Tim
the 17β-hydroxyl B/R2-alw
group, alkylation at the 17α position, and
modification of the steroid ring structure, respectively.
Reprinted, by permission, from A.T. Kicman, 2008, “Pharmacology of anabolic steroids,” British Journal of Pharmacology 154:
502-521.

(continued)

197
(continued)

198
Hormonal Mediation in Training Adaptations 199

linked to MAPK activation. Activation of Ras results in phosphorylation and activation of


MEK1 (MAPK and ERK1/2) kinases (see figure 5.1). Activated ERKs control cell growth
and cellular differentiation through phosphorylation of p90RSK and the nuclear TF Elk (see
figure 2.13; Taniguchi et al., 2006).
Because amino acid building blocks are necessary for skeletal muscle hypertrophy, it is
of interest that TORC1 activity is sensitive to, and dependent on, the availability of amino
acids. Administration of amino acids after acute exercise increases translation of contractile
proteins through stimulation of mTORC1, mTOR, and p70S6k and inhibition of TSC1/2
complex (figure 8.7). Amino acid action on mTORC1 occurs downstream of TSC1/2 but
requires activation by Rheb. Thus, administration of both glucose and amino acids after
exercise augments the hypertrophic response by different mechanisms: Glucose stimulates
insulin and its PI3K signaling pathway, and amino acids directly stimulate downstream
molecules in the same pathway.

SUMMARY
The first part of this chapter outlines the processes through which hormones contribute to
endurance-training adaptations. GH, catecholamines, and thyroid hormones are shown to
interact with nonhormonal processes in the tissues that are affected by exercise to produce
cardiorespiratory, thermoregulatory, and mitochondrial enhancements. Catecholamines and
thyroid hormones contribute to the shift in cardiac MHC toward the faster and more forcefully
contracting isoform, whereas GH and Ca BP with skeletal muscle mediate mitochondrial
biogenesis and the shift in MHC toward slower and more oxidative isoforms.
The second part of the chapter focuses on the role of GH, IGF-I, and T on muscle protein
synthesis and skeletal muscle hypertrophy through activation of the quiescent satellite cells.
The chapter describes the role of cortisol in suppressing protein synthesis and bone accre-
tion, the role of pulsatile PTH secretion in bone mineralization, and the role of sustained
elevation of PTH in bone resorption. Finally, this chapter addresses the role of nutrition
as a necessary enhancer of hypertrophic response. Both systemic hormones and enzymes
in skeletal muscle can sense variations in cellular energy availability to make appropriate
adjustments in protein synthesis, tissue remodeling, and other energy-requiring processes.
This page intentionally left blank.
H APTER

9
C

Exercise and
Endocrine Rhythms
H
ormones are secreted cyclically over periods of a few hours (circhoral or ultradian
pulsatility), a day (circadian or nycthemeral rhythm), several days (rodent estrous
cycle), a week (septadian rhythm), a month (circatrigintan, circamensal or menstrual
rhythms), a year (circannual rhythm), or a lifetime (periods of human statural growth and
female fertility). Circhoral or ultradian periodicity plays an important role in hormone action
when brief, intermittent hormone pulses are necessary to produce a specific physiological
effect. Some hormones follow a predominantly circadian or nycthemeral pattern; these
secretory variations must be taken into account when making measurements or studying
endocrine action (see “How to Avoid the Pitfalls of Hormone Measurement” in chapter 10).
Circatrigintan or menstrual cycles in hormone secretion introduce variability to studies of
female physiology and require that measurements be synchronized to specific phases of the
hormone cycle. Circannual rhythms in hormone secretion and physiology are more relevant
to processes such as variations in body fat storage, rates of tissue repair or growth, and
timing of fertility than to short-term, episodic behavior such as exercise. Lifetime rhythms
such as the prenatal and prepubertal cycles of statural growth (see figure 7.17) or the cycle
of female menstrual oscillations between menarche and menopause (see figure 7.1) often
are studied in the context of age-associated limitations on the capacity to exercise or adapt
to exercise training. Chronobiology—the study of biological rhythms in human, animal,
and plant physiology—has flourished during the past half century. This chapter reviews
the basics of biological rhythm physiology and terminology and the relevance of hormonal
and physiological rhythmicity to human performance, adaptations to exercise, and health.

BASICS OF BIOLOGICAL RHYTHM PHYSIOLOGY


AND TERMINOLOGY
Rhythmicity of physiological, endocrine, and behavioral processes is a fundamental feature
of human biology. The geophysical time based on the 24 h rotation of the Earth around the
sun allowed organisms on Earth to develop the ability to anticipate the onset of light and
darkness by adapting to either a diurnal or nocturnal pattern of activity and consigning the

201
202 Advanced Exercise Endocrinology

other part of the cycle to sleep and inactivity. This was recognized in the 1960s when Franz
Halberg first introduced the term circadian (Halberg, 1963). A circadian diurnal phase in
humans is characterized by the secretion of corticosteroids, insulin, and catecholamines,
whereas the circadian nocturnal phase is dominated by secretion of GH, PRL, TSH, AVP,
and melatonin (Boden et al., 1996; Rao et al., 1995). By having an endogenous clock with a
cycle period that approximates 24 h and that can be reset by zeitgebers (entraining stimuli,
such as light, physical activity, metabolism, food consumption, timed meals, and some
nutrients), animals and humans have developed the ability to capitalize on predictable
opportunities in their environment during the active phase and to conserve energy and avoid
predators during the inactive phase of the cycle. A circadian rhythm is endogenous because
it continues to free run with a period that approximates 24 h in an environment that lacks
entraining 24-h clues and can persist in free running with a different phase (onset point)
after being reset by a zeitgeber. External variables can only transiently mask the expression
of an endogenous rhythm. Exogenous rhythms, on the other hand, are entirely controlled
by an external variable. Thus, human exposure to light or exercise during the nocturnal
period will suppress or mask the secretion of melatonin without affecting the period of its
endogenous rhythm (Monteleone et al., 1990). Usual tests of an endogenous rhythm involve
maintenance of a constant routine without light, physical activity, or dietary entraining cues
or imposing a desynchrony routine that involves a period (e.g., 28 h) to which the endogenous
clock cannot easily entrain.

Parameters of Biological Rhythms


Sinusoidal ultradian and circadian fluctuations characterize all energy-dependent, life-
sustaining endogenous periodicities (e.g., heart and respiratory rates, metabolism and tem-
perature regulation). This important feature provides for bioenergetically efficient alternation
between energy-expending and waste-producing functions of the body on one hand and
energy-conserving and repair functions on the other. Each repeating unit in a sinusoidal
function represents a cycle in a time frame called a period or tau. The central value is called
a MESOR (midline estimating statistic of rhythm), and the change in value to either side
of the MESOR is the amplitude. The highest value in the amplitude above the MESOR is
the acrophase, whereas the lowest value below the MESOR is the nadir. Leptin (Ahima et
al., 1998), cortisol (Brandenberger & Follenius, 1975), melatonin (Monteleone et al., 1990),
and the catecholamines (Froberg et al., 1972) are hormones that have a sinusoidal circadian
pattern of secretion. The acrophases of their respective circadian rhythms occur at about
0400 h, 0900 h, 1600 h, and midnight. Most other endocrine rhythms display ultradian
secretory bursts above a relatively steady baseline concentration.
By convention, circadian time is measured with reference to the onset of a solar day,
which is designated as circadian time 0, or to the onset of spontaneous physical activity
in nocturnal mammals, which is designated as circadian time 12. Circadian rhythms are
sometimes represented by a circle with a periodicity of 360° and arrows or lines provid-
ing vectorial representation of the phase and amplitude of the acrophase. The phase of a
rhythm refers to any particular point in the period, whether it is presented as a sinusoidal
curve or a circle. Phase advances or delays in response to zeitgebers refer to changes in the
time of appearance of a phase in a periodicity. Zeitgebers are effective during the phases
of the cycle when the periods of the endogenous and exogenous entraining rhythm are not
aligned. For instance, light-entrainable endogenous clocks are photosensitive during the
transitions between light and darkness. Thus, in diurnal humans and nocturnal mammals,
the light-entrainable clock is phase advanced when light is presented during darkness at the
dark–light transition in the rhythm. It is phase delayed by light during the onset of dark and is
Exercise and Endocrine Rhythms 203

unresponsive to darkness and light during Constant


the midpoint of the light or dark periods, Entrained routine
respectively. Melatonin or naps act as Wake
Sleep
Wake Wake
Wake
Wake

pulses of darkness and thus phase delay


37.5
the light-entrainable clock when presented Core body
in early morning and phase advance the temperature 36.75
(°C)
clock when presented at the end of the 36.0
light period. Some biological effects 3
depend on changes in the phase relation- Urine volume
ships between two or more hormonal or (ml/min) 1.75
physiological rhythms. 0
Sleep and spontaneous physical activ- 200
ity display circadian periodicity but Melatonin
(pmol/L) 100
are discrete rather than continuous and
0
sinusoidal. They have the capacity to
entrain other rhythms. For instance, the 200
Cortisol
secretory patterns of GH and PRL are pre- (μg/100ml) 100
dominantly entrained to sleep and are not
0
governed by a circadian clock (Van Cauter
et al., 1992). Their nocturnal amplitudes Growth 28
are diminished during sleep deprivation hormone 14
(ng/ml)
(figure 9.1, right) and can be shifted when 0
the timing of sleep is altered (Maywood 18
Prolactin
et al., 2007). (μg/L) 9
0
Genetic Mechanism of the
Time (h) Time (h)
Circadian Clocks
Individual cells in most tissues (e.g., Borer/E4517/Fig.
Figure 9.1 Circadian 9.1/428900/Tim
and sleep-dependent B/R2-alw
control
endocrine glands, liver, intestine, heart, of daily physiological and endocrine rhythms. Some
adipose tissue, muscle, and retina) have physiological processes (e.g., temperature regulation,
the capacity to keep circadian time (Mel- urine volume) and hormone secretion (e.g., melatonin,
lani et al., 2011; Reppert & Weaver, 2002; cortisol) are predominantly circadian, some (e.g., TSH)
Young, 2006). In individual cells, basic are under the dual influence of the circadian clock and
circadian oscillation resides in coexpres- sleep (Gronfier & Brandenberger, 1998), and others (e.g.,
GH, PRL) depend on the sleep–wake rhythm.
sion of several core clock genes and their
Reprinted, by permission, from E.S. Maywood et al., 2007,
translation products, which exert positive “Minireview: The circadian clockwork of the suprachiasmatic nuclei-A
and negative feedback over the clock gene analysis of a cellular oscillator that drives endocrine rhythms,”
Endocrinology 148: 5624-5634; Adapted from Czeisler and Klerman
expression in a time-dependent fashion. 1999.
Additional posttranslational modifications
may affect the amplitude and timing of
circadian oscillations. The core genes of the molecular clock are Clock (circadian locomotor
output cycles kaput), Bmal1 (brain and muscle-Arnt-like protein1), Per (period) 1,2,3, and
Cry (cryptochrome) 1,2 (figure 9.2; Froy, 2010). The protein products of these genes act as
either positive (CLOCK and BMAL1) or negative (PER1,2,3 and CRY) TFs, all of which,
except for CLOCK, express circadian periodicities. The positive TF BMAL1 and CLOCK
accumulate in the cytoplasm, and when BMAL1 concentration threshold is exceeded it
dimerizes with CLOCK TF. A BMAL1–CLOCK heterodimer enters the nucleus to bind to
an E-box in the promoter region of the Per and Cry genes (figure 9.2 center top) as well as
204 Advanced Exercise Endocrinology

CLOCK BMAL1

CLOCK BMAL1
PERs E-box Per1-3
CRYs PERs
CLOCK BMAL1
CRYs
E-box Cry1-2

PERs
CRYs CLOCK BMAL1

PPRE E-box Rev-erbα CLOCK BMAL1


CLOCK BMAL1 E-box Pparα
E-box Rorα

Nucleus

Clock Cytoplasm

PPRE RORE Bmal1

PPARα RORs
REV-ERBs

Figure 9.2 The cellular oscillator is composed of positive feedback TFs (CLOCK and BMAL1) and negative
E4517/Borer/fig.9.2/428901/TB/R2-alw
feedback TFs (CRY and PER). Upon their translocation to the nucleus, CLOCK and BMAL1 initiate transcription of
other clock genes, including Per and Cry, whereas the translation products CRY and PER act as TF to inactivate
Clock and Bmal1 gene transcription. Auxiliary genes that contribute to stability and modulation of the core circa-
dian gene clock are Pparα, Rev-erbα, and Rorα. Their gene products act as TF that either inhibit (REV-ERBα) or
facilitate (PPARα, RORα) Per1 gene transcription.
Reprinted, by permission, from O. Froy, 2010, “Metabolism and circadian rhythms—Implications for obesity,” Endocrinology Review 31: 1-24.

other clock genes (Per1, Per2, Cry1, Cry2, Cry3, Pparα, Rev-erbα, and Rorα). This activates
the transcription of Per and Cry genes and auxiliary clock genes and leads to accumulation
4 to 6 h later of their products in the cytoplasm. PER and CRY proteins dimerize and, in
conjunction with a casein kinase 1ϵ (CK1 ϵ), act as a negative-feedback loop over the genetic
circadian clock. These TFs enter the nucleus and prevent further transcriptional gene activa-
tion by CLOCK–BMAL1 TFs. In addition to the core autoregulatory clock gene feedback
loop mediated by CLOCK–BMAL1 and PER–CRY, additional feedback loops contribute
to the stability and modulation of the transcriptional–translational clock mechanism. An
example of such auxiliary feedback loops is increased transcription by CLOCK–BMAL1
of Rev-erbα, Rorα, and Pparα genes. REV-ERBα TF inhibits the expression of the Bmal1
gene, whereas PPARα and RORα TF facilitate its transcription (figure 9.2).
Degradation of TF in both the positive and negative feedback pathways is initiated by
their phosphorylation by serine–threonine kinases CK1ϵ, GSK3β, and MAPK (Mellani et
al., 2011). Subsequent to phosphorylation, these TFs are degraded by ubiquitinization and
processing in the 26S proteasomal pathway (see chapter 2). This allows the reinstatement
of transcriptional activity of Clock and Bmal1 genes. Thus, the negative-feedback loop
between clock gene products and core clock gene expression results in robust, self-sustained
circadian rhythmicity of function in individual cells in many tissues.
Exercise and Endocrine Rhythms 205

Circadian Rhythms in Peripheral Tissues


Individual core clock genes are present in all peripheral tissues examined to date. Although
fewer than 4% of genes examined in these tissues display circadian rhythmicity, they con-
trol the expression of a number of genes of enzymes that are involved in lipid, amino acid,
glucose, and glycogen metabolism and in the citric acid cycle, especially during the active
portion of the nycthemeral cycle. Interestingly, the types of circadian clock-driven genes
expressed in different tissues possess a distinct specificity. Fewer than 10% of circadian
genes are shared between any two organs such as the liver, heart, kidney, and skeletal muscle
(McCarthy et al., 2007).
To date, researchers have most studied the circadian pattern of gene expression in the
heart, skeletal muscle, and the liver. In the heart, a circadian pattern of gene expression is
strongly expressed in 13% of genes studied (Martino et al., 2004) and is in part synchro-
nized by the strong circadian influence of sympathoadrenal motor output (see chapter 3
and figure 3.8). Some genes (e.g., Per, Bmal, MMP [metalloproteinase] genes associated
with myocardial protein transcription and translation, and genes for collagen remodeling,
mitochondrial respiration, and signaling pathways) display distinct circadian oscillations
driven by either the central clock or alternation between day and night. Other genes show
abrupt and sustained changes at dark–light transitions. Expression of the Per1 gene in
the heart was shown to phase shift within 1 d to a change in the timing of light exposure
(Sakamoto & Ishida, 2000).
Gene expression and timing of activities of some metabolic enzymes and processes can
also be synchronized to times of feeding, physical activity, and metabolic demand. This is
particularly clear in skeletal muscle, where, in addition to core clock genes such as Clock,
enzymes that reflect metabolic energy state (e.g., AMPK, PGC-1α, PGC-1β, and sirtuin 1),
all display circadian periodicity (Lamia et al., 2009; Liu et al., 2007; Nakahata et al., 2008)
and have the ability to influence the core clock genes (Liu et al., 2007, Mellani et al., 2011,
Nakahata et al., 2008). Among the core clock-driven genes that display circadian periodicity
is MyoD, the protein product of which acts as a TF that commits undifferentiated stem cells to
muscle lineage (Lefta et al., 2011; McCarthy et al., 2007). Expression of this gene is driven by
CLOCK–BMAL TF and it regulates expression of specific structural and metabolic skeletal
muscle genes. In addition to the role of tissue-specific core clock genes and the master brain
clock (discussed in the next section), circadian expression of heart and brain genes can be
mediated by the circadian influence of the hormone melatonin (Anisimov et al., 2002, 2004)
and circadian expression of skeletal muscle genes can be mediated by circadian influence of
the hormone cortisol (Almon et al., 2008). Melatonin influences circadian expression of genes
related to the cell cycle, apoptosis, biosynthesis, processing, and transport of nucleic acids.

SYNCHRONIZATION OF CIRCADIAN RHYTHMS


BY THE SUPRACHIASMATIC NUCLEUS
To influence so many aspects of physiology and behavior (including sleep–wake cycles,
cardiovascular activity, endocrine system, body temperature, renal activity, physiology
of the GI tract, and hepatic metabolism), the periods of the cellular oscillators need to be
synchronized among the individual tissues of the organism. A single master clock respon-
sible for this synchronization resides in the paired suprachiasmatic nuclei (SCN) of the
anterior hypothalamus. It is light entrainable through the input from a retinohypothalamic
tract (RHT) and controls most circadian periodicities throughout the body, which become
desynchronized after SCN damage.
206 Advanced Exercise Endocrinology

Researchers have extensively studied the synchronization of circadian rhythms by the


SCN. The paired SCN master clock appears to control peripheral rhythms through both
hormonal and neuronal connections (figure 9.3; Froy, 2010; Kalsbeek et al., 2011; Maywood
et al., 2007). The ability of fetal SCN tissue transplanted in animals bearing SCN lesions to
reinstate behavioral rhythmicity indicates that SCN exert their control over locomotion at
least in part by releasing paracrine factors into the surrounding brain nuclei. The candidate
peptides for paracrine control of circadian rhythms by SCN include vasopressin, TGFα,
prokineticin 2, and a cardiotrophin-like cytokine. Evidence for neural control of periph-
eral rhythms by SCN derives from experiments in which sleep and locomotor and cortisol
rhythms split into two 12 h cycles. These cycles appear to be independently controlled by
left and right SCN if they are 12 h out of phase.

Control of Endocrine Circadian Rhythms by Neuronal


Output of SCN
The ventrolateral core of the SCN contains cells that express VIP (vasoactive intestinal
polypeptide) and GRP (gastrin-releasing peptide), whereas the dorsomedial shell contains
AVP-positive cells. The core is entrained to light–dark cycles relayed by the RHT, which
transmits photic stimuli detected by the photopigment melanopsin in retinal ganglion cells.
Light initiates clock gene (Per) expressions in the core and their propagation to the shell
by way of GABA, VIP, and GRP signaling pathways. Although the SCN core does not
display distinct electrical or molecular rhythmicity, its destruction results in the loss of all
measured circadian responses, including rhythmic melatonin and corticosterone secretion,

IGL
PVN Raphe

VLPO DMH
SPZ
MPOA
ORX MCH

RHT SCN VMH


LH
Autonomic
NPY/AgRP POMC nervous system
ARC

Ghrelin,
insulin,
leptin,
glucose Gut-derived hormones,
nutrients,
Metabolism, abdominal distention
hepatic glucose production,
hormone secretion

Figure 9.3 Reciprocal neural and hormonal interactions between the SCN, brain pathways, and hormones
responsible for circadian control of hormone secretion and metabolism.
Reprinted, by permission, from O. Froy, 2010, “Metabolism and circadian rhythms—Implications for obesity,” Endocrinology Review 31: 1-24.
E4517/Borer/fig.9.3/428902/TB/R2-alw
Exercise and Endocrine Rhythms 207

body temperature, locomotor activity, and drinking (Kriegsfeld et al., 2004; Stephan &
Zucker, 1972). This suggests that the core region is essential for the production of the strong
circadian rhythmic output of gene expression in the SCN shell, which is therefore considered
to be the oscillating portion of the clock. In this way the master clock is synchronized to the
24 h solar time cycle. Upon photic stimulation, glutamate and PACAP (pituitary adenylate
cyclase–activating peptide) are coreleased in the SCN (Kalsbeek et al., 2011). The core clock
genes in SCN maintain the 24 h feedback loops and communicate the photic information
to the rest of the body by driving a circadian pattern of neuronal firing from the SCN. AVP
fibers from SCN play an important role in the synchronization and circadian control of
neuroendocrine functions.
Neural pathways through which the photoentrainable SCN control endocrine and physi-
ological rhythms include direct projections to neuroendocrine MPOA (medial preoptic area)
and preautonomic centers (i.e., the PVN of the hypothalamus, SPZ [subparaventricular zone],
DMH [dorsomedial hypothalamic nucleus]; see figure 9.3 and NA and DMV; Froy, 2010;
Kalsbeek et al., 2011). The acrophase of cortisol circadian rhythm occurs in all mammals
at the start of their active period and provides them with energy to initiate food-seeking
behaviors. SCN is responsible for circadian timing of this rhythm by activating CRH produc-
tion from the medial parvocellular PVN through an indirect neuronal route that includes the
stimulatory AVP and inhibitory GABA projections to SPZ and DMH (Kalsbeek et al., 2011).
In the adrenal cortex, the light and circadian signals from the SCN can drive the striking
circadian pattern of glucocorticoid synthesis and release independently of hypothalamo-
adenohypophyseal activation (Brandenberger & Follenius, 1975).
SCN controls the nocturnal secretion of melatonin through a similar indirect influence
over the pre-S regions of the brain. The SCN controls S outflow through direct innerva-
tions of the dorsal and ventral PVN, the source of neurons that activate preganglionic S
neurons (see figure 3.18). Melatonin synthesis from a serotonin precursor is initiated by
postganglionic S fibers from the superior cervical ganglion. After NE binds to beta AdRs
in the pineal gland, activity of the rate-limiting melatonin biosynthetic enzyme increases
100-fold. SCN inhibits melatonin release through activation of inhibitory GABA projections
to the PVN during the diurnal phase of the photoperiod. At the onset of darkness, GABA
output is curtailed and continuously active glutamergic neurons stimulate PVN to activate
preganglionic S fibers to the superior cervical ganglion and postganglionic fibers to the
pineal gland (Kalsbeek et al., 2011). Similar to the circadian control of melatonin secretion,
SCN exerts control over the preautonomic brain regions (NA) that are responsible for the
circadian rhythms of the heart, respiration, circulation, and blood pressure (see figure 3.8).
SCN also drives the circadian rhythm of male sex hormone secretion and the menstrual
rhythm in human and primate females (Karatsoreos & Silver, 2007). Late-nocturnal and
early-morning increases in T secretion are expressed particularly strongly during the pubertal
period and are controlled by circadian increases in LH secretion. The timing of ovulation is
also under precise circadian control that may be mediated by SCN projections to the hypo-
thalamic GnRH neurons. In seasonal breeders, annual changes in photoperiod length affect
reproduction and fat store size through SCN control over the circadian pattern of melatonin
secretion by the pineal gland (Morgan et al., 2003). SCN also regulates the circadian pattern
of leptin secretion and leptin-induced lipolysis (Kalsbeek et al., 2001; Shen et al., 2007).
Hormonal oscillations do not simply reflect the output of the SCN clock. Instead, daily
hormonal profiles are a product of the complex interaction between the output of the cir-
cadian pacemaker, periodic changes in behavior, light exposure, exposure to food under
conditions of variable availability, neuroendocrine feedback mechanisms, sex, age, and the
timing of sleep and wakefulness. SCN controls the secretion of melatonin, cortisol, leptin,
208 Advanced Exercise Endocrinology

ghrelin, insulin, catecholamines, and many other metabolic hormones, and these endocrine
rhythms in turn influence peripheral metabolism and physiology. For instance, circadian
and seasonal changes in melatonin secretion affect, respectively, sleep and seasonal breed-
ing and the onset of puberty. The hormone also has effects on immune function and cancer
suppression (Macchi & Bruce, 2004).

Control of Behavioral and Metabolic Circadian Rhythms by SCN


Through the dorsal SPZ, the SCN controls the circadian rhythm of body temperature, and
through the ventral part, the SCN controls the circadian sleep–wakefulness and spontane-
ous physical activity (locomotion). Cell bodies in the DMH, which are innervated by both
the SCN and the SPZ, also control the circadian pattern of sleep–wakefulness, locomotor
activity, and feeding. Circadian control of food intake, metabolism, and energy expenditure
depends on SCN connections with ARC and VMH nuclei and the ventral part of the LHA.
Inhibitory GABA interneurons from the SCN to the PVN facilitate PS functions. Conse-
quently, most viscera and their endocrine cells receive SCN-dependent circadian time cues
via their PS or S innervations.
Circadian variation in peripheral glucose uptake and hepatic glucose production depends
on the integrity of the SCN and their projections to PVN and S outflow (Cailotto et al.,
2005; Kalsbeek et al., 2006). Plasma glucose progressively increases to a peak at the start
of, and decreases halfway through, the active part of the photoperiod in both animals and
humans; the effect is independent of feeding status. This rhythm is abolished after lesions
of SCN (La Fleur et al., 1999). SCN appears to generate the rhythm of hepatic glucose
production by exerting GABA inhibition over perifornical orexin neurons; an acrophase
occurs during the second hour of the inactive period. Disinhibition of this GABA block-
ade stimulates orexin neurons in the perifornical hypothalamus to activate preganglionic
S neurons to the liver and increase the rate of glycogenolysis (Kalsbeek et al., 2011). The
circadian clock regulates many enzymes involved in energy metabolism, such as expression
of lactate dehydrogenase, ACC, malic enzyme, fatty acid synthetase, G6P dehydrogenase,
and many others (Froy, 2010). Circadian variation in lipid metabolism appears to be linked
to circadian changes in S and PS activity via SCN neural projections to the MPOA (see
figure 9.3; Bamshad et al., 1998).
Metabolic events at peripheral sites may influence the SCN master clock through affer-
ent neuronal and hormonal feedback loops (Froy, 2010). Thus, metabolic and physiological
information may reach the SCN through the binding of leptin and ghrelin to their receptors
on SCN neurons in addition to neural afferent signals. The latter include afferent S nerves and
PS afferents to the NTS and neuronal connections between the ARC—the principal site of
ghrelin and leptin action in the brain—and the SCN. The interaction between the molecular
clock, MyoD, and metabolic factors in the skeletal muscle, such as PGC-1α, provide a potential
system of feedback loops that may be critical for both the maintenance and adaptation of this
tissue (Lefta et al., 2011; McCarthy et al., 2007). Sex hormones also affect the operation of
the master clock. Estrogen and T increase the circadian amplitude of spontaneous locomotion,
shorten the free-running period, and advance the phase angle of entrainment of locomotor
rhythm (Karatsoreos & Silver, 2007). The influence of estrogen is probably indirect because
few ERs exist in the SCN. Instead, the effects occur through estrogenic action on the preoptic
area, corticomedial amygdala, bed nucleus of the stria terminalis, ARC, and serotonergic
neurons of the dorsal raphe nucleus and their connections to the SCN. On the other hand,
ARs found in the SCN core and in the retina may alter the photoreception and the opera-
tion of the master clock. The importance of hormonal feedback in the operation of the SCN
Exercise and Endocrine Rhythms 209

is illustrated by the loss of circadian rhythmicity of feeding after VMH–ARC lesions that
destroy the hypothalamic site of leptin action (Kakolewski et al., 1971).

CONTROL OF BIOLOGICAL RHYTHMS BY THE


FOOD-ENTRAINABLE OSCILLATOR
Another circadian clock synchronizes behaviors and physiology to food availability. Circa-
dian periodicities are food entrainable in that even in the absence of SCN, food seeking, liver
enzymes (Escobar et al., 1998; Hara et al., 2001; Honma et al., 1983) and many metabolic
functions display circadian rhythm related to the timing of restricted and predictable food
access. The food-entrainable oscillator overrides the photoentrainable SCN clock when
access to limited quantities of food is restricted to certain times of the day. Spontaneous loco-
motor activity is greatly increased before the scheduled presentation of a restricted amount of
food and is therefore called food-anticipatory activity. Food-anticipatory locomotor activity
is phase shifted to 2 to 4 h before the meal, and circadian body temperature, corticosterone
secretion, GI motility, activity of digestive enzymes, and enzyme oscillations in the liver,
kidney, heart, and pancreas all are entrained to the time of food access independently of SCN
control (Hara et al., 2001). However, all of these circadian rhythms revert to SCN control
when access to food becomes unrestricted and unpredictable (Stephan, 2002). Thus, phase
shifting is contingent on loss of body weight and negative energy balance during restricted
access to food and disappears when the quantity of food during timed access is sufficient
to prevent negative energy balance (Mendoza et al., 2008). When caloric restriction and
intermittent feeding are implemented on alternate days, they affect circadian rhythms by
synchronizing the photoentrainable SCN, probably because they cause weight loss. Both
food presentation schedules produce favorable reductions in morbidity and extend the pos-
sibility of survival, possibly by activating the cellular stress pathways (Anson et al., 2003).
The anatomical substrate of the food-entrainable central clock is less defined but appears
to involve the corticolimbic substrates that control salience of food stimuli and food seek-
ing locomotion under the conditions of energy deficit. These structures include the NAc,
basolateral and central amygdala, bed nucleus of the stria terminalis, lateral septum, pre-
frontal cortex, paraventricular thalamic nucleus, DMH, and parabrachial nuclei (Angeles-
Castellanos et al., 2007; Mieda et al., 2006; Mistlberger & Mumby, 1992).

CONTROL OF ULTRADIAN RHYTHMS


Secretory patterns of the majority of hormones and cardiorespiratory and most other vital
functions display endogenous ultradian rhythmicity. Rhythmic discharges of the SA node
at about 100 beats per min and of the AV node at about 60 beats per min provide baseline
heart muscle contractility that is further modified by autonomic, hormonal, and mechanical
stimuli. Ultradian pulses of pituitary GnRH are critically important for the maintenance
of mammalian gametogenesis and ovulation (chapter 7), and the ultradian alternating
rhythm of hypophysiotropic GHRH and inhibiting factor SRIF control the pulsatile release
of pituitary GH (Giustina & Veldhuis, 1998) that is necessary for bone and muscle growth
and development. Rhythmic ultradian secretion of releasing or inhibiting hormones persists
when these cells are studied in tissue cultures (Chappell et al., 2003). Although GnRH cells
are diffusely dispersed in the hypothalamus, they, like heart cells in tissue culture, adopt
and display collective synchronization to a common frequency (Lloyd & Murray, 2005).
Compared with the wealth of information about the operation of the SCN circadian clock,
much less is known about the ultradian rhythm mechanism. The generation of ultradian
210 Advanced Exercise Endocrinology

cardiac pacemaker activity requires coordinated activity of different membrane ionic chan-
nels as well as intracellular signaling factors, including Ca2+, and second messengers such
as cAMP (Bucchi et al., 2012). Spontaneous discharges of the SA and AV node result from
the flow of funny electrical current caused by a family of hyperpolarization-activated, cyclic
nucleotide-gated HCN channels. During diastole, the pacemaker membrane is hyperpolarized
from −40 mV to −70 mV through the vagal suppression of cAMP generation and the open-
ing of a K channel. The pacemaker cell membrane repolarizes below the funny electrical
current threshold (about −40/−50 mV) at the end of an SA action potential. This activates
the funny current and supplies inward the mixed Na–Ca current, which is responsible for
starting the diastolic depolarization and generating rhythmic pacemaker discharge. Of the
four cyclic nucleotide-gated (HCN) channels, HCN4 appears to be essential for the ultradian
heart pacemaker discharges because deletion of its Hcn4 gene reduces funny electrical cur-
rent by approximately 70%, and spontaneous pacemaker discharge rate by approximately
60% (Baruscotti et al., 2011).
It appears that the components of the core clock genes are responsible for the ultradian
pattern of GnRH secretion because disruption of Clock gene function slows GnRH pulse
frequency and increases its amplitude, whereas overexpression of the Cry1 gene results
in significantly higher GnRH amplitude that may control ovulatory LH release. At the
moment, it is uncertain whether the involvement of core clock genes in ultradian physiotropic
hormone release involves transcriptional or posttranslational mechanisms, effects on other
modulating genes, or an effect on membrane ion channels and depolarization. In addition
to its circadian control and a homeostatic component that adjusts sleep latency, duration,
and depth of sleep to sleep deprivation, sleep also displays ultradian periodicity (Landolt,
2011). Ultradian 70 min oscillations of different phases of sleep are characterized by differ-
ent patterns of EEG waves. Deep sleep is characterized by high-amplitude, low-frequency
EEG oscillations of 0.5 to 2 Hz (cycles/s) called delta waves. Sleep of intermediate depth
elicits faster theta waves (5 Hz), and light sleep displays phasic EEG spindles at 12 to 16 Hz,
known as sigma activity. Quiet wakefulness elicits low-amplitude, higher-frequency alpha
waves (9-12 Hz). These three stages of sleep are known as NREM (nonrapid eye move-
ment) sleep, in contrast to REM sleep that periodically interrupts slow delta waves. REM
sleep, during which humans experience dreaming, is identified by low-amplitude theta and
higher-frequency EEG activity, rapid eye movements, and absence of tone in antigravity
muscles. As sleep proceeds, its ultradian rhythmicity is manifested as declining prevalence
of delta oscillations, sleep depth, and the ratio of NREM to REM phases of sleep. Involve-
ment of core genes of the circadian clock in the control of ultradian sleep rhythm is seen
in changes in the duration of the deep-sleep phase in individuals carrying an isoform of
the Per3 gene on chromosome 1 (Landolt, 2011). Additional nonclock genes that appear to
modulate human ultradian sleep rhythm include Ada and genes for A1 and A2A adenosine
receptors that affect individual sensitivity to the effects of caffeine on sleep, as well as the
gene encoding the TrkB receptor for the BDNF that affects the duration of deep sleep. That
phases of sleep may act as synchronizers of hormone secretion is seen in the connections
between sleep and the secretion of GH, PRL, TSH, and cortisol. Pulses in GH and PRL are
positively linked to increases in delta wave activity of deep sleep, whereas TSH and cortisol
pulses are related to decreases in delta wave activity (Gronfier & Brandenberger, 1998).
Ultradian hormone pulses activate particular metabolic processes. This is especially
evident for cortisol, which shows high diurnal secretory activity of both the circadian and
ultradian kind (figure 9.4). Experimentally simulated ultradian cortisol peaks reach hormone
concentration levels that are sufficient to trigger genomic effects. These include expres-
sion of both core circadian clock genes (Bmal 1, Per2), genes involved in the suppression
Exercise and Endocrine Rhythms 211

Plasma cortisol (μg/dI) 20

15
Nonexercise day
10
Exercise day

Exer M
0
20
Plasma cortisol (μg/dI)

15

10

M Exer
0
20
Plasma cortisol (μg/dI)

15

10

M M Exer
0
8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
Time of day (h)

Figure 9.4 Circadian, prandial, and exercise effects on cortisol secretion. The influence .of time of day on the
E4517/Borer/Fig9.4/428904/alw/R2
magnitude of cortisol elicited by exercise. Ninety minutes of exercise at 25% to 55% of VO2max stimulated a
large peak of cortisol release when exercise was performed during the time of day when cortisol is not secreted
(1000h-1130h and 2130h-2300h). When exercise was performed during the midday cortisol peak (1300h-1430h),
it elicited no additional cortisol secretion.
Reprinted, by permission, from C. Brandenberger et al., 1982, “Feedback from meal-related peaks determines diurnal changes in cortisol
response to exercise,” Journal of Clinical Endocrinology and Metabolism 54: 592-596.

of protein translation (glutamine synthetase gene), lipid metabolism genes (srebpf1), liver
enzyme genes for utilization of amino acids for gluconeogenesis, and genes involved in
innate immune responses (Almon et al., 2008).

CONTROL OF BIOLOGICAL RHYTHMS BY EXERCISE


Physical performance displays circadian periodicity and thus is of interest in both science
and sport. The involvement of physical activity with biological rhythmicity is complex and
reciprocal. Coordinated endogenous hormonal and physiological rhythms affect physical
activity, including optimal outcomes of exercise training, as well as health and well-being.
Spontaneous locomotor activity also may be governed by a central neural clock because it
has the capacity to synchronize and control circadian rhythms in the muscle (Yamanaka
et al., 2008), phase shift some circadian rhythms, increase the amplitude of ultradian and
endocrine rhythms, and phase shift seasonal and lifetime rhythms. The identity of neural
circuits involved in this control is currently not known.
212 Advanced Exercise Endocrinology

Effects of Physiological Circadian Periodicities


on Physical Performance
Researchers have recognized for some time that human performance displays circadian
periodicity (Minors & Waterhouse, 1981; Reilly et al., 1997a). Oxygen consumption, CO2
production, minute ventilation, flexibility, peak knee-extension torque, and hand and leg
strength all have a nadir at 0600h and an acrophase at 1800h, mirroring circadian changes in
core temperature (Tc). Consequently, the fastest swim times were recorded mostly between
1800h and 2200h and rarely, if ever, at 0600h. Rather than reflecting a causal relationship
with circadian temperature changes, physical performance may be coincidental to tempera-
ture changes that result from circadian rhythms in metabolic processes in the muscle and
organs supporting the energy demands of exercise (Froy, 2010).
Jet lag is an issue of importance in international sport competitions. Jet lag occurs when
humans are transported across time zones by transmeridian travel. This throws endogenous
circadian rhythms out of phase with the solar day. Such misalignments between exogenous
and solar cycles produce inappropriate timing of sleepiness and wakefulness, urine pro-
duction, and metabolic rhythms and disturbances in appetite, digestion, and mood. As a
result, physical performance is impaired through slower reaction and locomotor times and
decreased strength (Reilly et al., 1997b; Shephard, 1984). Entrainment to a new light–dark
cycle usually takes 1 d per time zone displacement but can be accelerated, as mentioned
earlier, by using light, naps, melatonin, and exercise as zeitgebers. Zeitgebers that facilitate
phase delays for adjustments to westward time-zone crossings are more effective than those
entraining eastward phase advances, possibly because human circadian rhythms generally
have periods that are longer than 24 h.
The recommendation that humans should consume protein-rich meals in the morning
to increase alertness and carbohydrate-rich meals in the evening to induce sleepiness as a
means of phase shifting to a new light–dark schedule is supported by the capacity of certain
nutrients to phase shift the clock. Nutrients that are capable of resetting or phase shifting
circadian rhythms are amino acids (Iwanaga et al., 2005), thiamine (Langlais & Hall, 1998),
retinoic acid (Shirai et al., 2006), Na (Mohri et al., 2003), ethanol (Spanagel et al., 2005),
caffeine (Antle et al., 2001), and glucose (Stephan & Davidson, 1998).

Exercise as a Rhythm Modulator and a Zeitgeber


Exercise can both stimulate and suppress ultradian hormone secretion (see chapter 5).
Stimulation by exercise of S activity inhibits insulin secretion. Exercise also can affect
hormone secretion by producing an energy drain and reduction in the size of energy stores
that then exerts indirect effects on secretion of hormones such as LH (see figure 7.18, e and
f; Loucks & Thuma, 2003).
Exercise can increase both the frequency (Borer et al., 1986) and, more often, the ampli-
tude (Pritzlaff et al., 1999; Figure 5.10) of ultradian pulsatile secretion of GH and other coun-
terregulatory hormones in a dose-dependent fashion. The elicitation of endocrine responses
by exercise is not predictable but is in some cases dependent on ultradian, circadian, or
sleep-entrained rhythms. For instance, time of day affects cortisol responses to 90 min of
moderate-intensity exercise at between 25% and 55% of maximal effort (Brandenberger &
Follenius, 1975). Exercise triggers an ultradian cortisol pulse at times when the secretion of
hormone is declining from its early-morning circadian acrophase (1000h-1200h) or is at its
circadian nadir (2200h-2400h) but not when an endogenous ultradian pulse is in progress
(1200h-1400h; figure 9.4). Reflecting these pronounced circadian and ultradian rhythms
Exercise and Endocrine Rhythms 213

.
of secretion, 30 min of intense exercise at 85% of VO2max elicits greatest cortisol response
(a 600% increase in the area under the curve) at the hormone’s circadian nadir but smaller
increases in concentration of between 150 and 200% at the hormone’s acrophase at 0700h
and during an early-evening (1900h), meal-associated ultradian pulse (Kanaley et al., 2001).
In the case of GH, exercise affects circhoral hormone secretion by influencing hypotha-
lamic releasing and inhibitory factors, which also can interfere with GH pulse elicitation
during ultradian refractory periods that usually follow ultradian hormone bursts (Borer et
al., 1986; Thompson et al., 1993; Veldhuis et al., 2004). Thus, when the same 30-min intense
exercise stimulus is applied at 0700h, 1900h, or 2400h, it produces similar increases in GH
areas under the curve, possibly because these times do not coincide with the ultradian . GH
peaks (Kanaley et al., 2001). However, if 3 bouts of 30 min of exercise at 70% of VO2max
are performed sequentially at intervals shorter than the duration of the 90 min morning
GH pulse (1000h, 1130h, and 1200h), the amplitude of 3 elicited GH pulses is about one
half the size of the amplitudes achieved by exercise performed at 1000h, 1400h, and 1800h,
intervals longer than the combined ultradian GH pulse and postpulse refractory period of
about 240 min (Kanaley et al., 1997). The 4 h spaced timing of exercise produces progres-
sive amplification of ultradian GH pulse amplitudes, but both exercise timing paradigms
produce the same 150% to 160% increase in 16 h cumulative GH secretion, indicating that
difference in the timing of exercise may alter the amplitude of individual hormone pulses
but preserve total integrated hormone response. The stimulatory effects of exercise on deep
sleep as well as associated increases in GH secretion during its delta-wave phase (Van Cauter
et al., 1992) may be mediated by increases in Tc when exercise lasts more than 1 h because
experimentally reduced Tc prevents exercise-associated GH response (Horne & Staff, 1983).
In addition to modulating the amplitude of ultradian hormone pulses, exercise can phase
shift the circadian, annual, and lifespan clocks. For instance, a 3 h bout of moderate-intensity
exercise at the time of Tc nadir at between 0200h and 0600h phase delays the secretion of
melatonin and TSH by 1 h (figure 9.5; Van Reeth et al., 1994). This evidence shows that
humans can use exercise as a phase-delay Zeitgeber.

1.2 1.05
Exercise
1.0 No exercise
0.95
Melatonin (% of mean)

TSH (% of mean)

0.8

0.6 0.85

0.4
0.75
0.2

0.0 0.65
−01 00 +01 −01 00 +01
Hours before or after onset of melatonin rise on day 1 Hours before or after onset of melatonin rise on day 1

Figure 9.5 A 3 h bout of moderate-intensity exercise performed at the time of the Tc nadir phase delays melatonin
and TSH rhythm by 1 h.
E4517/Borer/Fig9.5/401622/alw/pulled-R2
Reprinted, by permission, from O. Van Reeth et al., 1994, “Nocturnal exercise phase delays circadian rhythms of melatonin and thyrotropin secretion in
normal men,” American Journal of Physiology 266: E964-E974.
214 Advanced Exercise Endocrinology

Voluntary running reinstates fertility in seasonally breeding golden hamsters that are
maintained in a short photoperiod that is representative of the nonbreeding season (Borer
et al., 1983). This demonstrates that exercise can phase shift the annual rhythm of seasonal
breeding. Voluntary running reinstates high growth hormone pulsatility (Borer & Kelch,
1978) and accelerates somatic and skeletal growth at the adult stage of hamster life that is
ordinarily characterized by slow growth (Borer & Kuhns, 1977) but is ineffective during
the initial rapid-growth phase that is characteristic of immature animals (Borer & Kaplan,
1977). This suggests that exercise has the capacity to phase delay or prolong the life span
growth rhythm.

Effects of Hormone Rhythms on Training Adaptations


Many biological effects of hormones depend critically on the temporal pattern of hormone
release. This is because the ultradian patterning of hormone secretion plays an exception-
ally important role in controlling cellular growth and fertility, the two processes that also
display a periodicity in the lifespan. Cellular proliferation in the bone chondrocytes and
muscles of hypophysectomized rats depends critically on the pulsatile secretion, rather than
continuous delivery, of GH (Isgaard, Carlsson et al., 1988; Isgaard, Moller et al., 1988), and
the highest rate of ultradian GH secretion occurs during the peripubertal period of rapid
statural growth. After the termination of the growth phase in adult humans, pulsatile GH
delivery increases lipolysis, and continuous infusion of GH stimulates IGF-I expression and
its anabolic effects in skeletal muscle (Surya et al., 2009; see figure 5.8). Intermittent secre-
tion or administration of PTH stimulates bone mineral accretion in the cancellous bone,
and continuous elevation of PTH causes bone resorption (Jilka, 2007). Researchers have
attributed this effect to a cellular negative feedback consisting of acute facilitation by PTH
of the expression of antiapoptotic genes in osteocytes or mesenchymal osteoblast progenitor
cells in the bone marrow and the expression of proapoptotic genes 6 h later (Bellido et al.,
2003). The 6 h rhythm of stimulatory PTH action is mediated by the cAMP and PKA trans-
duction pathways (see figure 8.4). Activation by PKA of transcription of the antiapoptotic
cell survival genes such as Bcl-2 is mediated by Runx2 and CREB. PTH also facilitates
expression and anabolic action of several growth factors (e.g., IGF-I, FGF-2, TGFβ, and
BMP), cytokines (e.g., IL-6), and autocrine–paracrine factors (e.g., Wnt and Hedgehog).
The result is an increase in the number of osteoblasts, osteogenic activity in osteoblasts,
and suppression of adipogenic differentiation of MSC by PPARγ.
Prolonged elevation of PTH concentration is triggered by low plasma Ca concentrations.
It leads to bone mineral resorption and manifests as secondary hyperparathyroidism in Ca-
deficient postmenopausal women. Under such conditions, exposure to chronically elevated
PTH concentration leads to Runx2 degradation in the proteasomes through the action of
Smurf1, and a delayed action of PKA allows for expression of proapoptotic genes such as
Bad, and both actions result in osteoblast apoptosis and bone resorption. The dependence
of anabolic bone mineral accretion on intermittent but short-lasting exposure to PTH is the
basis for the treatment of osteoporosis with daily injections of the PTH analog teriparatide
(Sikon & Batur, 2010). Because exercise and mechanical loading stimulate a PTH pulse,
6 h anabolic action of PTH pulse most likely accounts for why rodent bone growth and
mineralization double when mechanical loading is spaced at 8 h intervals compared with
the effect of a single, consolidated episode of loading (Robling et al., 2001). In terms of the
role of ultradian hormone secretion on mammalian fertility, pulsatile administration—but
not a continuous infusion—of hypothalamic GnRH maintains LH pulsatility and fertility in
human and primate females (Belchetz et al., 1978). As discussed in chapter 7, suppression
of LH pulsatility below a critical circhoral LH frequency by the energy drain of exercise
Exercise and Endocrine Rhythms 215

training suppresses the fertility of young female athletes who experience exercise-induced
amenorrhea.
Coaches have practiced the empirical principle of periodicity in training for many years.
A tenet of periodicity is that exercise needs to be interspersed with hours, days, or weeks of
rest or reduced physical activity to allow maximal training adaptations to develop. These
practices implicitly acknowledge that many functions (e.g., those having to do with circa-
dian changes in anabolic and catabolic hormone release, timing of resynthesis of metabolic
fuels during exercise recovery, or timing of postprandial nutrient absorption for synthesis
of additional contractile proteins) must be accommodated to the rhythmic processes in the
human body. These considerations often do not receive sufficient attention because the
scientific mindset is strongly biased toward examining the effects of individual independent
variables outside of a chronobiological context.
One example of the role of circadian influences in muscle hypertrophy is the observa-
tion that optimal adaptations to resistance training (i.e., muscle hypertrophy and strength
increases) seem to occur when an athlete performs training in the late afternoon. At that
time, exercise-induced increases in T concentration coincide with a circadian decline in
concentrations of the catabolic hormone cortisol (Hayes et al., 2010). Researchers have
also recently recognized that timing of exercise and meals and intake of critical nutrients
affects the magnitude of skeletal muscle protein synthesis. Adding carbohydrate to amino
acid intake amplifies hypertrophic response, which strongly suggests that ultradian meal-
associated insulin release plays an important role in contractile protein synthesis. The role of
the ultradian relationship between phases of catabolic cortisol and meal-associated anabolic
insulin release (see figure 8.6, a and b) underscores the importance of ultradian timing of
the hormonal milieu in adaptations to exercise training. The timing of hepatic and muscle
glycogen depletion can substantially alter the effect of exercise on regulatory functions
such as blood glucose setpoint because the Km of HK, the glucose-capturing enzyme in the
muscle, is 100 times lower than that of hepatic GK. Thus, depending on whether exercise
is performed in a postprandial or fasted state, the unequal rate of postexercise glycogen
resynthesis in the two organs may alter the glucagon glycemic threshold and the level at
which blood glucose is regulated (Borer et al., 2009).

Chronomedicine and Chronopharmacology:


Circadian Rhythms and Human Health
The early observation by Franz Halberg that the anticoagulative efficacy of aspirin varies
depending on the time of day it is taken (Halberg et al., 1991) highlighted the importance of
circadian and other biorhythmic processes on human physiology and health. Aspirin prevents
blood clotting by inhibiting thromboxane synthesis and cyclo-oxygenase-dependent platelet
aggregation, processes that are more active in early morning. A 100 mg dose of aspirin was
most effective in suppressing thromboxane synthesis and enhancing the number of β2 AdRs
on lymphocytes when taken within 3 h of waking. It was less effective if taken 6 to 9 h later,
and ineffective if taken 12 h later. Heart attacks, strokes, pulmonary edema, hypertensive
crises, asthma, and allergic rhinitis attacks all peak at certain times during the day (Singh
et al., 2002). A circadian pattern of cardiovascular crises is related to the circadian pattern
of catecholamine secretion and action that peaks around midday; this pattern is abolished
by administration of beta blockers. The circadian pattern of allergy-related crises has a
nocturnal or early-morning acrophase and is associated with the circadian pattern of secre-
tion of histamine and other mediators of immune reaction that is about 180° out of phase
with the circadian acrophase of bronchodilatating catecholamine secretion. Disruption of
the functioning of the central biological clock can also cause sleep disorders and affect
216
217
218 Advanced Exercise Endocrinology

psychological functioning; the key determinants are the circadian rhythms in the secretion of
metabolic hormones or action of lipogenic and lipolytic enzymes (Reppert & Weaver, 2002).
Disturbances in the operation of circadian physiology also may play an important role in
the development of obesity and metabolic syndrome (Froy, 2010; Staels 2006). In general,
conditions such as caloric restriction and associated negative energy balance and weight
loss reduce morbidities associated with metabolic syndrome and extend the life span in a
number of organisms. These effects are associated with improved synchrony and higher
amplitudes of central and peripheral ultradian and circadian rhythms. Activation of AMPK
and increased concentrations of the oxidized forms of NAD and the sirtuin SIRT1 may
play a role in this outcome because these enzymes display circadian activity patterns, are
responsive to negative energy states, and have the ability to alter the molecular mechanics
of the circadian clock. By contrast, high-fat diets and associated weight gain and obesity
are associated with phase shifting and declines in the amplitude of locomotor and feeding
rhythms. Seasonal fluctuations in body fat stores attest to the functional control of adipose
tissue lipid metabolism by the clock mechanism. Changes in the functional control of adi-
pose tissue lipid metabolism by the clock mechanism also may be mediated by changes in
cortisol secretion and its lipogenic actions.
Sleep is a clock-controlled behavior. It is photoentrained to the solar cycle, but this
entrainment is often disrupted by the contemporary lifestyle in which the timing of solar
illumination is overridden by artificial lighting and sleepiness is overruled by protracted
lighting and other imperatives. Volitional shortening of the sleep period and shift work are
associated with rapid development of several morbidities such as increased BMI and height-
ened risk of T2D and hypertension. These effects may be mediated, in part, by a disruption
of circadian rhythms of hormones such as ghrelin and leptin and of their influence on the
motivation to eat and on the control of energy metabolism (Spiegel et al., 1999).

SUMMARY
Most of human physiology, including hormone secretion and action, is organized in a rhyth-
mic fashion to produce specific biological effects, increase energy efficiency, or coordinate
behavior and physiology with environmental opportunities. This chapter first provides basic
information about biological rhythms, the molecular mechanism of the circadian clock
that operates in most tissues, and the coordination and synchronization of these peripheral
oscillations by hypothalamic suprachiasmatic nuclei. This chapter then describes how these
nuclei control hormonal secretion and synchronize peripheral periodicities in the heart,
skeletal muscle, and liver. Because hormones, heart pacemakers, and sleep also display
ultradian periodicities, the more limited insight into the mechanisms of these clocks and
the food-entrainable clock is discussed next. The chapter also addresses the importance of
biological periodicities in human performance and the capacity of physical activity to affect
amplitude and frequency of hormone secretion and to phase shift endocrine, seasonal, and
lifetime rhythms. This chapter then describes the significance of hormone secretion pattern
in facilitating cellular growth fertility and bone mineral accretion and provides a cautionary
note about the variability in elicitation by exercise of ultradian hormone pulses depend-
ing on the phase of their endogenous control. The final section highlights the emerging
fields of chronomedicine and chronopharmacology in recognition of the important role of
hormonal and physiological rhythms in the maintenance of health and efficacy of drugs.
An appreciation of the dependence of many biological processes—including those elicited
by exercise—on circadian and ultradian timing of hormone secretion or nutrient intake
and exercise is bound to significantly increase one’s understanding of exercise physiology.
H APTER

10
C

Measuring Hormones
B
ecause hormones affect human behavior, physiology, and metabolism at rest and
during exercise, the need to measure them is both important and understandable.
However, the appropriate selection of tissue or body-fluid sources of hormones,
aspects of hormone secretion and action, and approaches of measurement can present chal-
lenges. The basic question that motivates the desire to measure hormones will determine
the practical problems that one needs to resolve before one can successfully carry out mea-
surement. This chapter discusses how to best measure hormone concentration, hormone
biological action, hormone synthesis and site of release, and rate of hormone secretion.

MEASURING HORMONE CONCENTRATION


The usual approach to understanding the role of hormones in a given behavior such as move-
ment or feeding is to assess changes in the pattern of hormone concentration as a function of
time of occurrence and parameters of the independent variable. This useful, albeit correla-
tional, approach requires one to make decisions about the medium from which the hormone
will be sampled, the timing and frequency of sampling, and collection method to be used.

Choosing the Medium for Hormone Collection


Because hormones are traditionally defined as biologically active substances that reach their
targets through circulation, measurement of hormone concentration typically is based on
plasma derived from unclotted blood or serum that remains after clotting. The amount of
plasma or serum obtained is usually about one half of the blood volume. For measurements
using radioimmunoassay methods (RIAs) and depending on assay sensitivity, quantities
of plasma or serum needed usually range between 50 and 100 μl (double this amount for
duplicate determinations) and are sometimes as much as 1 ml when the hormone must be
extracted from plasma or serum. The extracted hormone is then measured in duplicate. For
measurement of hormones by the more sensitive chemiluminescent techniques required to
detect low hormone concentrations, a single aliquot of plasma (~200 to 300 μl) is required
for duplicate or multiple measurements of the hormone in the luminometer. The size of
plasma or serum volume can be reduced to 10 to 25 μl in recently developed multiplex
chemiluminescent hormone assays. With this approach, it is possible to measure multiple
hormones simultaneously in the small plasma or serum sample utilizing fluorescence
produced by reagents applied to either hormones or hormone antibodies. Whether one

219
220 Advanced Exercise Endocrinology

selects acute venipuncture or an indwelling catheter in a peripheral vein for the withdrawal
of a single or repeated blood samples depends on the circumstances. It may be difficult to
collect serial blood samples from an indwelling catheter from, for example, a competitor
during a marathon. On the other hand, study of hormone secretion in a controlled clinical
or laboratory setting may create an unnatural situation that may not properly reflect the
phenomenon being studied. One must weigh the potential stress of acute venipuncture,
which may confound the hormonal profile by stimulating the release of stress hormones,
against the reduced stress of collecting blood from an indwelling catheter in a more con-
trived, controlled setting. Blood samples can be collected automatically with pumps when
sequential blood samples are desired in ambulatory humans. Small, portable pumps are
attached to the clothing and are programmed to withdraw blood into appropriate collection
tubes at timed intervals. This approach was successfully used in studies of hormones with
distinct pulsatile characteristics, such as GH (Bjarnason et al., 2002).
A possible complication in measuring blood hormone concentration is the acute decline in
PV (hemoconcentration) during exercise and adaptive increases in PV in response to endur-
ance training. Acute hemoconcentration depends on the intensity of exercise and involves
PV declines that change from 6% to 12% as relative exercise intensity increases from 50%
to 90% (Wilkerson et al., 1977). These shifts in plasma from the extracellular compartment
to the intracellular compartment are associated with intensity-dependent increases in plasma
renin and AVP concentrations (Convertino et al., 1983). When it is important to determine
whether the change in hormone concentration results from increased release of hormone
or from reduction in PV, the magnitude of hemoconcentration is measured from changes
in the hematocrit (the relative volume of packed red blood cells after centrifugation in a
blood-filled capillary). When the biological action of a hormone is of greater interest, the
change in plasma or serum concentration of hormone during exercise is assessed without
correction for hemoconcentration because actual concentration of a hormone determines
its effect. This issue usually arises in situations where changes in hormone concentration
can be accounted for by changes in the hematocrit. This brings up the question whether
exercise actually triggered hormone secretion or the apparent hormone concentration change
was due to hemoconcentration. Regardless of the answer to the question, an increase in the
concentration of circulating hormone likely has a biological effect even if it was caused by
reduced PV.
Depending on the susceptibility of a given hormone to oxidative or enzymatic degrada-
tion, collection of plasma or serum may require appropriate enzyme blockers. EDTA-coated
tubes or the addition of proteolytic enzyme blockers to collected blood are used to measure
catecholamines and many other rapidly degraded hormones. Additional blockers of the
enzyme dipeptidyl peptidase 4 or serine protease inhibitor need to be added to collected blood
to protect some gut peptides from rapid degradation, such as such as GLP-1 and acylated
ghrelin. Neglecting this precaution may result in the loss of most or all hormone activity.
Tissue sources of hormones are sometimes obtained with microdialysis. Hormones can
be measured in microdialysate from the brains of animals (Kitaoka et al., 2010) and the
adipose and other tissues of humans (Lange et al., 2002) and animals (Mano-Otagiri et al.,
2009). This requires the insertion of small-diameter tubing into the tissue in question and
using pumps that generate negative pressure to produce a slow flow of small volumes of
dialysate into collection tubes.
Less-invasive methods of measuring hormone concentration involve the use of saliva and
urine. For salivary measurements, subjects are given a small cotton plug to chew; saliva can
then be extracted from the plug by centrifugation. When salivary flow is insufficient, subjects
Measuring Hormones 221

can increase salivation by chewing a piece of paraffin. Steroid hormones (Ives et al., 2011)
and melatonin (Kozaki et al., 2011) have been successfully measured with this approach.
Most hormones undergo enzymatic degradation, and their byproducts are eliminated
in the urine. Therefore, urine is another source of information about hormone produc-
tion. Because urine accumulation over time precludes information about short-term
changes in hormone secretion, urinary hormone measurements are appropriate when
daily changes in hormone production are of interest. This approach has been useful in
assessing changes in daily catecholamine secretion in response to stress (Froberg et
al., 1972) or reduction in urinary estrone glucuronide and pregnanediol glucuronide in
response to energy restriction during endurance training (Loucks et al., 1989).

Choosing the Timing and Frequency of Hormone Collection


The biological action of some hormones (e.g., GH, PTH, and LH) is critically dependent
on their pattern of secretion (see chapters 5 and 9). Pulsatile delivery of GH is essential for
cellular proliferation in developing bone and muscle of growing rodents (Isgaard, Carlsson et
al., 1988; Isgaard, Moller et al., 1988) and for lipolysis in adult humans (Surya et al., 2009),
whereas constantly elevated levels of the respective hormones are ineffective or result in
different hormone action (e.g., muscle protein synthesis instead of lipolysis). Pulsatile PTH
secretion promotes bone mineral accretion, whereas constant elevation of PTH concentrations
leads to bone mineral loss (Jilka, 2007). LH pulsatility is necessary for the maintenance of
normal menstrual cycles, and its reduction or absence abolishes them (see chapter 7). Thus,
one must make an appropriate choice regarding the frequency of blood sampling in order
to document the pattern of hormone release and fully understand its action.
The frequency of sampling depends in part on hormone half-life. Hormones with a short
half-life such as GH and LH require sampling at least every 10 to 20 min to accurately
capture fluctuations in hormone concentrations, although sampling every 5 min captures
significantly more hormone peaks than does sampling every 10 or more min (Evans et al.,
1987). Insulin concentrations oscillate at 4 to 5 min when blood is sampled from the hepatic
portal vein (Pørksen, 2002). Oscillatory secretion patterns of insulin are important in con-
trolling target tissue sensitivity to the hormone and are attenuated in individuals with T2D
(Polonsky et al., 1988). Hormones with longer half-life, such as FSH, require less-frequent
sampling to establish a true pattern of fluctuations.
Several statistical procedures for detecting true hormone pulses among random measure-
ment noise have been developed. The procedure used most frequently is cluster analysis
(Veldhuis & Johnson, 1986). In this computerized algorithm, a pulse is defined as a statisti-
cally significant increase in a cluster of hormone values followed by a statistically significant
decrease in a second cluster of values. The increase or decrease is judged in relation to the
actual experimental error expressed by the replicates in the presumptive nadir and peak
data clusters. The program permits the operator to specify the cluster sizes of test peaks and
pre- and postpeak nadirs. This method is largely insensitive to unstable baseline hormone
concentrations and is not adversely affected by varying pulse amplitudes, widths, or con-
figurations in the endocrine series. In addition, the simple statistical basis for this algorithm
renders it minimally dependent on explicit or a priori assumptions about rates of hormone
secretion or disappearance. The program has been validated for false-positive errors against
a wide range of intra-series coefficients of variation (4%-52%). The program is applicable
for profiles of LH, FSH, GH, PRL, ACTH, and cortisol, which exhibit pulsatile fluctuation.
When it is of interest to determine endocrine consequences of an acute event such as
exercise rather than to accurately document the pattern of hormone fluctuation, one still
222 Advanced Exercise Endocrinology

needs to collect blood at a number of time points in order to define the time course of the
change in hormone concentration. For instance, peak concentration of GH is attained some
time after the termination of exercise, depending on the intensity of exercise (see figure 5.10).
Changes in plasma leptin in response to the energy expenditure of exercise are seen several
hours into recovery (figure 6.9b, Borer et al., 2009) and were originally missed in studies in
which only a single or an early postexercise blood sample was taken (Weltman et al., 2000).

Choosing the Method for Measuring Hormone


Choosing an appropriate method for measuring hormone may depend on limitations to total
permissible volume of blood collection, sensitivity of the measurement method, availability
of measurement equipment, and cost of the assays. Institutional review boards place a limit
on the quantity of blood that can be safely withdrawn from healthy individuals per unit of
time. One must take this consideration into account when selecting both a measurement
approach and a sample volume. Multiplex assays make the most economical use of blood
or other sample volumes, because one can measure multiple hormones simultaneously in a
10 to 25 μl volume of serum or plasma assayed in 96-well microtiter plates. The use of this
approach is limited by the availability of the appropriate luminometer apparatus. Because
the sensitivity of chemiluminescent assays can be up to 500 times greater than that of
corresponding RIA, chemiluminescent assays may reveal much lower ranges of hormone
concentrations when that is the desired goal. Both methods depend on the availability of
appropriate equipment.
RIA generally are more economical in terms of cost but not necessarily in terms of
amount of blood because they require duplicate measurements of between 50 and 100 µl
of plasma or serum. They also expose the technician to modest gamma irradiation because
the isotope with which the hormone is usually tagged is 125iodine. The principle of RIA
and most hormone assays involves the use of an antibody specific to a given hormone. The
antibody, a gamma globulin, is usually developed by a genetically unrelated species (e.g., a
rodent or lagomorph for primate hormones and vice versa) immunized against the hormone
through serial subcutaneous hormone injections. The specificity of the antibody is a very
important consideration because many hormones with structural similarities (e.g., GH and
PRL, CCK and gastrin, β-endorphin and β-LPH) generate antibodies to the conserved regions
of the molecule and therefore cross-react with the hormones that are structurally similar.
This is sometimes solved by the use of monoclonal antibodies developed to the dissimilar
part of the molecule or by use of discriminating polyclonal antibodies. In the case of CCK,
the problem is solved by alcohol extraction of CCK to prevent it from cross-reacting with
gastrin. It is therefore important to check the information about hormone cross-reactivity in
the specifications of any individual hormone assay. Another consideration when selecting an
RIA for measuring hormones is whether antibodies for a hormone in a given species have
been validated for measuring this hormone in a different species. If the standard curves
measured by the same antibody in the two species are not parallel, concentration measure-
ments will be distorted and inaccurate (Borer et al., 1982).
RIA and most hormone assays using other kinds of labels are based on the principle of
competitive binding between the antibody and, in this case, the nonradioactive and radioac-
tive hormone. A standard curve pits known and systematically increasing concentrations of
nonradioactive hormone against a constant amount of radioactive hormone that is added to
all assay samples. As the proportion of nonradioactive hormone increases, it competitively
displaces the radioactive hormone from binding with the antibody. After appropriate sched-
ules of incubation that allow for hormone–receptor binding, the unbound hormone is removed
Measuring Hormones 223

by any number of separation techniques. In solid-phase assays, the antibody is coated to the
inner wall of the assay tube or well wall and the unbound hormone is simply poured out. In
competitive binding assays where the specific antibody is added in liquid form at high dilu-
tions (usually between 1 and 50,000 or 1 and 100,000), the hormone bound to the antibody
needs to be precipitated. This is usually achieved with a second antibody that binds to the
gamma globulin of the first antibody. After centrifugation, the bound hormone–antibody
complex is compacted at the bottom of the tube and the unbound radioactive hormone is
poured off. Other separation techniques remove unbound hormone with activated charcoal
or other adsorptive substances. The percentage of the bound radioactive hormone (less the
nonspecific binding of radioactive hormone to the tubes in the absence of the antibody) can
then be calculated against the antibody-bound hormone in the absence of the competition.
These reference or B0 tubes provide the information about the level of hormone–antibody
binding in the assay, which, for best competitive hormone displacement, should be close to
40%. Assay sensitivity in some RIAs is increased by delaying the addition of radioactive
hormone. In such assays, a 1 d incubation of the nonradioactive hormone in the standards
and samples with the antibody results in better competitive hormone displacement from
the antibody after delayed addition of the tracer a day later. The binding curve is sinusoidal
and is usually linearized by using a log-logit plot (logit of the ratio of the binding in tubes
containing the hormone over B0 tubes against the log of standard hormone concentrations).
The standard curve allows extrapolation of concentration of the hormone in standards to
hormone in plasma or serum samples. Before the recent introduction of the chemilumines-
cent and multiplex technologies, the use of RIA revolutionized measurement of hormone
concentration. In 1977, its developer, Rosalyn Yalow, received a Nobel Prize in physiology
and medicine for this achievement.
Variations on the RIA theme are IRMA and ELISA. IRMA (immunoradiometric assay)
involves applying a hormone to the appropriate adsorbent on the surface of a 96-well
microtiter plate, to which a radioactive antibody is applied. After incubation and washing
off of the unbound hormone in standards or samples, the amount of radioactive antibodies
bound to the hormone is quantified in plate readers that can detect isotope disintegrations.
ELISA (enzyme-linked immunosorbent assay) was developed to reduce operator exposure
to radioactive isotopes by using chemical reactions that produce a change in fluorescence or
optical absorbance. It can be used for determination of either a hormone or an antibody in
a sample. Several variations of the ELISA procedure are available. All use two antibodies:
one to the hormone and one to the first antibody. The second antibody generally is linked
to an enzyme that changes optical density or fluorescence in the presence of an appropri-
ate substrate. After separating unbound hormone–enzyme complexes by washing, one can
quantify hormone concentration from the color or fluorescence intensity in a spectrometer
plate reader. As in the RIA, in each assay a known series of increasing concentrations of
the hormone serves as a reference standard. In indirect ELISA, the hormone is adsorbed to
the plate well surface. A neutral protein wash removes competing proteins from the well
surface. The primary antibody is then allowed to interact with the hormone. A secondary
enzyme-linked antibody is next added, and the color or fluorescence is measured after the
enzyme–substrate reaction occurs. In the sandwich ELISA, two antibodies to the hormone
are used: a capture antibody to coat the walls of the plate wells and a detecting antibody to
attach to the hormone after it binds to the capture antibody. A third, enzyme-linked second-
ary antibody that has the property of binding to the detecting antibody is then applied. After
application of the enzyme substrate, the color or fluorescence is measured. Competitive
ELISA is the third variant of this approach. Here, the hormone is first applied to the walls
224 Advanced Exercise Endocrinology

of the plate well. Hormone is then allowed to separately incubate with the antibody before
it is applied to the plate wells. The hormone on the walls competes with the hormone in the
hormone–antibody solution and reduces the amount of antibody binding to the hormone
on the well walls. The secondary antibody to the first antibody is enzyme linked and will
produce different intensity of color or fluorescence in interaction with an enzyme substrate.

MEASURING BIOLOGICAL ACTIONS OF HORMONES


The most elementary approach to understanding hormone action is to remove the gland pro-
ducing the hormone to observe the consequences on the physiology and behavior. Although
this approach usually cannot detect multiple actions of a given hormone and may be lethal
when deleted hormones support vital functions, it can produce remarkable results. An
example is the series of Alfred Goldberg experiments (see chapter 8) demonstrating that GH
and insulin are not necessary for hypertrophic muscle adaptation to mechanical loading in,
respectively, hypophysectomized rats and rats rendered diabetic by administration of alloxan.
A complementary approach to the study of hormone function is to administer the hormone
in situations where it is absent, inactive, or blocked. Classical examples include administra-
tion of insulin to individuals with T1D or GH to pituitary dwarfs to correct physiological
deficits caused by hormone deficiencies. To study GH or other hormone action following the
infusion of the hormone without the interference of endogenous hormone release, a blocker
of GHRH or other releasing hormone can be applied (Jessup et al., 2004). The major caveat
in the interpretation of hormone substitution experiments is that one must apply the hor-
mone in a physiological concentration and a pattern that simulates the normal concentration
and secretory pattern. Applying pharmacological doses of the hormone may not reveal its
physiological role and may lead to wrong inferences about its actions. For example, result
may differ when administration of a hormone is pulsatile rather than sustained and of a
constant concentration. IGF-I mRNA production in muscle and bone of hypophysectomized
rats increases in response to pulsatile but not constant administration of GH (Isgaard, Carls-
son et al., 1988; Isgaard, Moller et al., 1988), LH pulsatility and reproductive cycles are
maintained after pulsatile but not constant administration of GnRH (Belchetz et al., 1978),
and pulsatile GH administration increases lipolysis in muscle of adult humans whereas an
increase in the invariant concentration of GH increases IGF-I mRNA production (Surya
et al., 2009). Additional inferences about hormone action can be obtained by applying a
hormone receptor blocker. Examples of successful application of this strategy are studies
demonstrating the role of type A CCK receptor in appetite suppression (Corwin et al., 1991)
and the leptin receptor for elicitation of lipolysis (Siegrist-Kaiser et al., 1997). A hormone-
blocking action of other hormones is also sometimes used. For instance, constant infusion
of SRIF can be used to suppress pancreatic release of insulin and glucagon so that the role
of the latter two hormones in exercise can be studied (Pencek et al., 2004).
A bioassay is the prototypic test of biological action of the hormone. It is used when
one is uncertain whether changes in hormone concentration measured by an RIA reflect
changes in the biological action of the hormone. The bioassay entails administration of dif-
ferent doses of a hormone to tissues or cell cultures in which the hormone is known to act.
One of the oldest bioassays is a tibia test (Li & Lostroh, 1957). In this assay, a preparation
containing purified GH or extract of pituitary gland is injected into hypophysectomized
rats. The change in the width of the epiphyseal growth plate in the proximal tibial epiphysis
is a measure of the potency of tested GH relative to a standard concentration of human GH.
Another example of a bioassay is the use of Leydig cell culture to measure androgenic action
of circulating LH and its second messengers (Payne & Youngblood, 1995).
Measuring Hormones 225

Another approach to measuring hormone action is to clamp hormone concentration in


circulation and quantify the magnitude of its biological effect. Insulin clamps have had
wide application in the assessment of the capacity of insulin to facilitate peripheral glucose
uptake as a measure of tissue sensitivity to insulin (Foley et al., 1985). Glucose is infused
to maintain euglycemia, and the total amount or rate of glucose infused reflects the hypo-
glycemic effectiveness of insulin. Hormone action can also be assessed by quantifying in
a standardized fashion a hormone response to a reference challenge or metabolic response
to a reference hormone challenge. This is most often used when insulin resistance is sus-
pected. A well-known example is the oral glucose tolerance test, where changes in insulin
and glucose concentrations are evaluated over time against a reference response after a
standard amount of glucose is administered by mouth. A more informative test of insulin
sensitivity is the frequently sampled intravenous glucose tolerance test (Bergman, 2003).
Here, tissue sensitivity to insulin action is assessed by measuring the magnitude of insulin
response to a glucose stimulus as well as the effectiveness of insulin in removing the glucose
from plasma after a subsequent insulin infusion. A meal tolerance test measures insulin
and glucose responses to a standardized meal containing carbohydrate in addition to other
nutrients. An insulin tolerance test similarly examines the sensitivity of peripheral tissues
to insulin action by measuring the extent to which different concentrations of the hormone
lower plasma glucose.

MEASURING HORMONE SYNTHESIS AND SITE


OF PRODUCTION
Because the majority of hormones are peptides or proteins, a frequently used approach is
measuring the synthesis of an mRNA for a given hormone in response to an appropriate
stimulus. A good example is tracking changes in tissue IGF-I mRNA expression to determine
the response to an anabolic stimulus (see figure 2.5). mRNA in tissue extracts is usually
quantified with hybridization assays (Durnam & Palmiter, 1983). In these assays a specific
reverse and tritium-labeled mRNA is used to hybridize under appropriate conditions with
an mRNA of a particular hormone. This procedure creates a specific double-stranded
mRNA hybrid that is protected from degradation to which single-strand mRNA is suscep-
tible. Breakdown products of unprotected single-strand mRNA in the extract are washed
away, and the specific mRNA hybrids are collected on a filter paper and quantified against
a standard curve in a scintillation counter. A similar procedure can be used to determine
the site of hormone synthesis by mRNA hybridization in histological sections of a tissue.
To track synthesis of nonprotein hormones such as catecholamines (Arbogast & Voogt,
2002) and steroids (Aizawa et al., 2010), changes in the concentrations of their biosynthetic
enzymes are usually measured.

MEASURING THE RATE OF HORMONE SECRETION


Some hormones produce a slowly metabolized byproduct during their conversion from a
prohormone to a more rapidly metabolized, biologically active hormone. The concentration of
the byproduct or the ratio of byproduct to active hormone can then provide a better measure
of the dynamics of hormone secretion. An example is assessment of insulin synthesis from
changes in the ratio of C peptide (figure 1.4) to proinsulin or insulin (Schölin et al., 2011).
However, deconvolution analysis is a more accurate and powerful approach to measuring
the rate of hormone synthesis (Johnson et al., 2004). Deconvolution analysis identifies hor-
mone secretory burst mass [amount of hormone secreted per pulse per distribution volume
226
227
228 Advanced Exercise Endocrinology

(μg/dl)], pulse frequency, and basal secretion rate using an algorithm that utilizes the rapid
phase and slow phase of hormone half-life and filters out secretory noise. Approximate
entropy, another parameter of interest in the analysis of pulsatile hormone secretion, mea-
sures relative orderliness of hormone secretion. Irregularity in hormone secretion signifies
loss of negative feedback or feed forward control (Veldhuis et al., 2001).

SUMMARY
This chapter outlines how to effectively measure hormone concentration, biological action,
synthesis and site of release, and rate of secretion. To measure hormone concentration, one
needs to choose the medium from which the hormone will be sampled: plasma, serum, tissue,
saliva, or urine. One also must choose an appropriate frequency of sampling based on the
properties of spontaneous or exercise-induced hormone secretory pattern or examination
of hormone biological action. When frequent hormone sampling is used, one must use an
appropriate statistical method for identifying hormone pulses. This chapter describes sev-
eral methods for measuring hormones (RIA, IRMA, ELISA) and their competitive binding
principles. The relative benefits of using endocrine-gland removal, hormone replacement,
administration of constant hormone levels, or standardized challenges to assess hormone
action are described. The use of mRNA measurements for the study of the rate and location
of protein hormone synthesis is also described. For other hormones such as catecholamines,
measurement of their biosynthetic enzymes is used instead. Finally, this chapter outlines
approaches for determining secretory rates of hormone.
References
Chapter 1 BM. A PGC1-α-dependent myokine that drives
brown-fat-like development of white fat and
Bayliss WM, Starling EH. The mechanism of pan-
thermogenesis. Nature. 2012; 481(7382): 463-468.
creatic secretion. J Physiol. 1902; 28: 325-353.
Cartee GD, Funai K. Exercise and insulin: Con-
Fitzsimons JT. Angiotensin, thirst, and sodium
vergence or divergence at AS160 and TBC1D1?
appetite. Physiol Rev. 1998; 78: 583-686.
Exer Sport Sci Rev. 2009; 37: 188-195.
Lang J, Nishimoto I, Okamoto T, Regazzi R, Kiraly
Cartee GD, Wojtaszewski JFP. Role of Akt sub-
C, Weller U, Wollheim CB. Direct control of
strate of 160 kDa in insulin-stimulated and
beta cell exocytosis by receptor-mediated acti-
contraction-stimulated glucose transport. Appl
vation of the heterotrimeric GTPases Gi and
Physiol Nut Metab. 2007; 32: 557-566.
G(o) or by the expression of their active G alpha
DeVol DL, Rotwein P, Sadow JL, Novakofski J,
subunits. EMBO J. 1995; 14: 3635-3644.
Bechtel PJ. Activation of insulin-like growth
Lanning HJ, Carter-Su C. Recent advances in
factor gene expression during work-induced
growth hormone signaling. Rev Endocr Metab
skeletal muscle growth. Am Journal Physiol.
Disord. 2006; 7: 225-235.
1990; 259: E89-E95.
Li J, Al-Azzawi F. Mechanism of androgen receptor
action. Maturitas. 2009; 63: 142-148. Ervin RB. Prevalence of metabolic syndrome
Little TJ, Horowitz M, Feinle-Bisset C. Role of among adults 20 years of age and over, by sex,
cholecystokinin in appetite control and body age, race and ethnicity, and body mass index:
weight regulation. Obesity Rev. 2005; 6: 297-306. United States, 2003-2006. Natl Health Stat
Porges SW. The polyvagal theory: Phylogenetic Report. 2009; 5: 1-7.
substrates of a social nervous system. Psycho- Fatone C, Guescini M, Balducci S, Battistoni S,
neuroendocrinol. 1998; 23: 837-861. Settequattrini A, Pippi R, Stocchi L, Mantuano
Wilkins L, Richter CP. A great craving for salt by a M, Stocchi V, De Feo P. Two weekly sessions of
child with cortico-adrenal insufficiency. JAMA. combined aerobic and resistance exercise are
1940; 114: 866-873. sufficient to provide beneficial effects in subjects
with Type 2 diabetes mellitus and metabolic
Chapter 2 syndrome. J Endocr Invest. 2010; 33: 489-495.
Allen RG, Tresini M. Oxidative stress and gene reg- Favier FB, Benoit H, Freyssenet D. Cellular and
ulation. Free Radic Biol Med. 2000; 28: 463-499. molecular events controlling skeletal muscle
Baar K, Blough E, Dineen B, Esser KA. Transcrip- mass in response to altered use. Pflugers Arch-
tional regulation in response to exercise. Exer Eur J Physiol. 2008; 456: 587-600.
Sport Sci Rev. 1999; 27: 333-379. Frøsig C, Richter EA. Improved insulin sensitiv-
Balkau B, Mhamdi L, Oppert JM, Nolan J, Golay ity after exercise: Focus on insulin signaling.
A, Porcellati F, Laakso M, Ferrannini E, EGIR- Obesity (Silver Spring). 2009; 17(suppl 3): S15-20.
RISC Study Group. Physical activity and insulin Fukada E, Yasuda I. On the piezoelectric effect on
sensitivity: The RISC study. Diabetes. 2008; 57: bone. J Physical Soc Japan. 1957; 10: 1158-1169.
2613-2618. Garetto LP, Richter EA, Goodman MN, Ruder-
Bassett CAL, Pawluk RJ, Pilla AA. Generation man NB. Enhanced muscle glucose metabolism
of electric potentials by bone in response to after exercise in the rat: The two phases. Am J
mechanical stress. Science. 1962; 137: 1063-1064. Physiol. 1984; 246: E471-E475.
Benzet S, Sanchez H, Chapot R, Peinnequin A, Goldberg AL. Work-induced growth of skeletal
Bigard X, Kouklmann N. Basal peroxisome pro- muscle in normal and hypophysectomized rats.
liferator activated receptor gamma coactivator Am J Physiol. 1967; 213: 1193-1198.
1α expression is independent of calcineurin in Grundy S, Brewer B, Cleeman J, Smith S, Lenfant
skeletal muscle. Metabolism. 2012; 61: 389-394. C. Definition of metabolic syndrome. Report of
Bonewald LF, Johnson ML. Osteocytes, mecha- the National Heart, Lung, and Blood Institute/
nosensing and Wnt signaling. Bone. 2008; 42: American Heart Association Conference on
606-615. Scientific Issues Related to Definition. Circula-
Boström P, Wu J, Jedrychowski MP, Korde A, Ye tion. 2004; 109: 433-438.
L, Lo JC, Rasbach KA, Boström EA, Choi JH, Hawley JA, Zierath JR. Integration of metabolic
Long JZ, Kajimura S, Zingaretti MC, Vind BF, and mitogenic signal transduction in skeletal
Tu H, Cinti S, Højlund K, Gygi SP, Spiegelman muscle. Exer Sport Sci Rev. 2004; 32: 4-8.

229
230 References

Hornberger TA, Armstrong TA, Koh DD, Burk- Lin J, Wu H, Tarr PT, Zhang CY, Wu Z, Boss O,
holder TJ, Esser KA. Intra cellular signaling Michael LF, Puigserver P, Isotani E, Olson EN,
specificity in response to uniaxial vs. multiaxial Lowell BB, Bassel-Duby R, Spiegelman BM.
stretch: Implications for mechanotransduction. Transcriptional co-activator PGC-1alpha drives
Am J Physiol. 2005; 288: C185-C194. the formation of slow-twitch muscle fibres.
Hornberger TA, Esser KA. Mechanotransduction Nature. 2002; 418: 797-801.
and the regulation of protein synthesis in skeletal Liu L, Yu B, Chen J, Tang Z, Zong C, Shen D,
muscle. Proc Nutrit Soc. 2004; 63: 331-335. Zheng Q, Tong X, Gao C, Wang J. Different
Hornberger TA, Stuppard R, Conley KE, Fedele effects of intermittent and continuous f luid
MJ, Fiorotto ML, Chin ER, Esser KA. Mechani- shear stresses on osteogenic differentiation of
cal stimuli regulate rapamycin-sensitive signal- human mesenchymal stem cells. Biomech Model
ing by a phosphoinositide 3-kinase, protein Mechanobiol. 2012; 11: 391-401.
kinase B- and growth factor-independent Long YC, Zierath JR. Influence of AMP-activated
mechanism. Biochem J. 2004; 380: 795-804. protein kinase and calcineurin on metabolic
Hwang SY, Putney JW Jr. Calcium signaling in networks in skeletal muscle. Amer J Physiol.
osteoclasts. Biochim Biophys Acta. 2011; 1813: 2008; 295: E545-E552.
979-983. Luu YK, Capilla E, Rosen CJ, Gilsanz V, Pessin
Jackson MJ, Papa S, Bolanos J, Bruckdorfer R, JE, Judex S, Rubin CT. Mechanical stimulation
Carlsen H, Elliott RM, Flier J, Griffiths HR, of mesenchymal stem cell proliferation and
Heales S, Holst B, Lorusso M, Lund E, Oivind differentiation promotes osteogenesis while pre-
Moskaug J, Moser U, Di Paola M, Polidori MC, venting dietary induced obesity. J Bone Miner
Signorile A, Stahl W, Vina-Ribes J, Astley SB. Res. 2009; 24: 50-61.
Antioxidants, reactive oxygen and nitrogen spe- Maarbjerg SJ, Sylow L, Richter EA. Current under-
cies, gene induction and mitochondrial function. standing of increased insulin sensitivity after
Mol Aspects Med. 2002; 23: 209-285. exercise—Emerging candidates. Acta Physiol
Jespersen JG, Nedergaard A, Andersen LL, (Oxf). 2011; 202: 323-335.
Schjerling P, Andersen JL. Myostatin expres- Matheny RW Jr, Nindl BC, Adamo ML. Minire-
sion during human muscle hypertrophy and view: Mechanogrowth factor: a putative product
subsequent atrophy: Increased myostatin with of IGF-I gene expression involved in tissue repair
detraining. Scand J Med Sci Sports. 2011; 21: and regeneration. Endocrinol 2010; 151: 865-875.
215-223. Morino K, Petersen KF, Dufour S, Befroy D, Frat-
Jessen N, Goodyear LJ. Contraction signaling tini J, Shatzkes N, Neschen S, White MF, Bilz
to glucose transport in skeletal muscle. J Appl S, Sono S, Pypaert M, Shulman GI. Reduced
Physiol. 2005; 99: 330-337. mitochondrial density and increased IRS-1
Jeyasingham MD, Artigues A, Nadeau OW, Carl- serine phosphorylation in muscle of insulin-
son GM. Structural evidence for co-evolution resistant offspring of type 2 diabetic parents. J
of the regulation of contraction and energy Clin Invest. 2005; 115: 3587-3593.
production in skeletal muscle. J Mol Biol. 2007; Niess AM, Simon P. Response and adaptation of
377: 623-629. skeletal muscle to exercise—The role of reac-
Jorgensen SB, Jensen TE, Richter EA. Role of tive oxygen species. Frontiers Biosci. 2007; 12:
AMPK in skeletal muscle gene adaptation in 4826-4838.
relation to exercise. Appl Physiol Nut Metab. Nikander R, Kannus P, Rantalainen T, Uusi-Rasi
2007; 32: 904-911. K, Heinonen A, Sievänen H. Cross-sectional
Jorgensen SB, Rose AJ. How is AMPK activity geometry of weight-bearing tibia in female
regulated in skeletal muscles during exercise? athletes subjected to different exercise loadings.
Frontiers Biosci. 2008; 13: 5589-5604. Osteoporos Int. 2010; 10: 1687-1694.
Koh H-J, Brandauer J, Goodyear LJ. LKB1 and Patterson TE, Sakai Y, Grabiner MD, Ibwoye M,
AMPK and the regulation of skeletal muscle Midura RJ, Zborowski M, Wolfman A. Expo-
metabolism. Curr Opinion Clin Nut Metab Care. sure of murine cells to pulsed electromagnetic
2008; 11: 227-232. fields rapidly activates the mTOR signaling
Kostek MC, Chen Y-W, Cuthbertson DJ, Shi pathway. Bioelectromagnetics. 2006; 27: 535-544.
R, Fedele MJ, Esser KA, Rennie MJ. Gene Petersen KF, Shulman GI. Etiology of insulin resis-
expression responses over 24 h to lengthening tance. Am J Med. 2006; 119(5 suppl 1): S10-S16.
and shortening contractions in human muscle: Phielix E, Meex R, Ouwens DM, Sparks L, Hoeks L,
Major changes in CSRP3, MUSTN1, SIX1, and Schaart G, Moonen-Kornips E, Hasselink MKC,
FBX32. Physiological Genomics. 2007; 31: 42-52. Schrauwen P. High oxidative capacity due to
Kreke MR, Sharp LA, Lee YW, Goldstein AS. chronic exercise training attenuates lipid-induced
Effect of intermittent sheer stress on mechano- insulin resistance. Diabetes. 2012; 61: 2472-2478.
transductive signaling and osteoblastic differ- Pienkowski D, Pollack SR. The origin of stress-
entiation of bone marrow stromal cells. Tissue generated potentials in fluid-saturated bone. J
Engin A. 2008; 14: 529-537. Orthop Res. 1983; 1: 30-41.
References 231

Powers SK, Jackson MJ. Exercise-induced oxida- Urso ML. Disuse atrophy of human skeletal muscle:
tive stress: Cellular mechanisms and impact on Cell signaling and potential interventions. Med
muscle force production. Physiol Rev. 2008; 88: Sci Sports Exerc. 2009; 41: 1860-1868.
1243-1276. Violett B, Mounier R, Leclerc J, Yazigi A, Foretz
Rath B, Nam J, Knobloch TJ, Lannutti JJ, Agarwal M, Andrelli F. Targeting AMP-activated protein
S. Compressive forces induce osteogenic gene kinase as a novel therapeutic approach for the
expression in calvarial osteoblasts. J Biomech. treatment of metabolic disorders. Diab Metab.
2008; 41: 1095-1103. 2007; 33: 395-402.
Richter EA, Mikines KJ, Galbo H, Kiens B. Effect Ward DF Jr, Salasznyk RM, Klees RF, Backiel
of exercise on insulin action in human skeletal J, Agius P, Bennett K, Boskey A, Plopper
muscle. J Appl Physiol. 1989; 66: 876-885. GE. Mechanical strain enhances extracellular
Richter EA, Ploug T, Galbo H. Increased muscle matrix-induced gene focusing and promotes
glucose uptake after exercise. No need for insu- osteogenic differentiation of human mesenchy-
lin during exercise. Diabetes. 1985; 34: 1041-1048. mal stem cells through an extracellular-related
Röckl KSC, Witczak CA, Goodyear LJ. Signaling kinase-dependent pathway. Stem Cells Develop.
mechanisms in skeletal muscle: Acute responses 2007; 16: 467-479.
and chronic adaptations to exercise. IUBMB Wasserman DH. Four grams of glucose. Am J
Life. 2008; 60: 145-153. Physiol. 2009; 296: E11-E21.
Rommel C, Bodine SC, Clarke BA, Rossman R, Wasserman DH, Kang L, Ayala JE, Fueger PT,
Nunez L, Stitt TN, Yancopoulos GD, Glass DJ. Lee-Young RS. The physiological regulation
Mediation of IGF-1-induced skeletal myotube of glucose flux into muscle in vivo. J Exp Biol.
hypertrophy by PI(3)K/Akt/mTOR and PI(3) 2011; 214: 254-262.
K/Akt/GSK3 pathways.  Nat Cell Biol. 2001; Winder WW, Holmes BF, Rubink DS, Jensen EB,
11: 1009-1113. Chen M, Holloszy JO. Activation of AMP-acti-
Rose AJ, Richter EA. Skeletal muscle glucose vated protein kinase increases mitochondrial
uptake during exercise: How is it regulated? enzymes in skeletal muscle. J Appl Physiol. 2000;
Physiology (Bethesda). 2005; 20: 260-270. 88: 2219-2226.
Rubin CT, Capilla E, Luu YK, Busa B, Crawford Winder WW, Thompson DM. Cellular energy
H, Nolan DJ, Mittal V, Rosen CJ, Pessin JE, sensing and signaling by AMP-activated protein
Judex S. Adipogenesis is inhibited by brief, kinase. Cell Biochem Biophys. 2007; 47: 332-347.
daily exposure to high-frequency, extremely Witzak CA, Sharoff CA, Goodyear LJ. AMP-
low-magnitude mechanical signals. Proc Natl activated protein kinase in skeletal muscle: From
Acad Sci. 2007; 104: 17679-17884. structure and localization to its role as a master
Rubin J, Murphy TC, Zhu L, Roy E, Nanes MS, Fan regulator of cellular metabolism. Cell Mol Life
X. Mechanical strain differentially regulates Sci. 2008; 65: 3737-3755.
endothelial nitric oxide synthase and receptor Wojtaszewski JF, Nielsen JN, Richter EA. Invited
activator of nuclear kappa B ligand expression review: Effect of acute exercise on insulin signal-
via ERK 1/2 MAPK. J Biol Chem. 2003; 278: ing and action in humans. J Appl Physiol. 2002;
34018-34025. 93: 384-392.
Rubin J, Rubin C, Jacobs CA. Molecular pathways Yamashita M, Katsumata M, Iwashima M, Kimura
mediating mechanical signaling in bone. Gene. M, Shimizu C, Kamata T, Shin T, Seki N, Suzuki
2006; 367: 1-16. S, Taniguchi M, Nakayama T. T cell receptor-
Sakuma K, Yamaguchi A. The functional role induced calcineurin activation regulates T
of calcineurin in hypertrophy, regeneration, helper type 2 cell development by modifying the
and disorders of skeletal muscle. J Biomed interleukin 4 receptor signaling complex. J Exp
Biotechnol. 2010; 2010: 721219. Epub 2010 Apr Med. 2000; 191: 1869-1879.
1. PMID:21960735. [PubMed - indexed for
MEDLINE]. Chapter 3
Schenk S, Cook JN, Kaufman AE, Horowitz JF. Altura BM, Altura BT. Actions of vasopressin, oxy-
Postexercise insulin sensitivity is not impaired tocin and synthetic analogs on vascular smooth
after an overnight lipid infusion. Am J Physiol. muscle. Fed Proc. 1984; 43: 80-86.
2005; 288: E519-E525. Bishop VS, Ryuzaki M, Cai Y, Nishida Y, Cox BF.
Schiaffino S, Reggiani C. Molecular diversity Angiotensin II-dependent hypertension and
of myofibrillar proteins: Gene regulation and the arterial baroreflex. Clin Exp Hypertension.
functional significance. Physiol Rev. 1996; 76: 1995; 17: 29-38.
371-423. Bittar E. Pulmonary Biology in Health and Disease.
Schwartz MA. Integrins and extracellular matrix New York: Springer Verlag; 2002.
in mechanotransduction. Cold Spring Harb Brooks VL. Interactions between angiotensin
Perspect Biol. 2010; 2: a005066. Epub 2010 Nov II and the sympathetic nervous system in the
17. PMID: 21084386. [PubMed - indexed for long-term control of arterial pressure. Clin Exp
MEDLINE]. Pharmacol Physiol. 1997; 24: 83-90.
232 References

Cervero F, Foreman RD. Sensory innervations Ponchon P, Elghozi JL. Contribution of humoral
of the viscera. In: Loewy AD, Spyer KM, eds. systems to the recovery of blood pressure fol-
Central Regulation of Autonomic Functions. New lowing severe hemorrhage. J Auton Pharmacol.
York: Oxford University Press; 1990: 104-125. 1997; 17: 319-329.
Convertino VA, Keil LC, Bernauer EM, Greenleaf Rädegran G, Hellsten Y. Adenosine and nitric
JE. Plasma volume, osmolality, vasopressin, and oxide in exercise-induced human skeletal muscle
renin activity during graded exercise in man. J vasodilatation. Acta Physiol Scand. 2000; 168:
Appl Physiol. 1981; 50: 123-128. 575-591.
Deuster PA, Chrousos GP, Luger A, DeBolt JE, Richter DW, Spyer KM. Cardiorespiratory control.
Bernier LL, Trostmann UH, Kyle SB, Montgom- In: Loewy AD, Spyer KM, eds. Central Regula-
ery LC, Loriaux DL. Hormonal and metabolic tion of Autonomic Functions. New York: Oxford
responses of untrained, moderately trained, and University Press; 1990: 189-207.
highly trained men to three exercise intensities. Rowell LB. Human Cardiovascular Control. New
Metabolism. 1989; 38: 141-148. York: Oxford University Press; 1993.
d’Uscio LV, Shaw S, Barton M, Luscher TF. Stampler JS, Meissner G. Physiology of nitric oxide
Losartan but not verapamil inhibits angiotensin in skeletal muscle. Physiol Rev. 2001; 81: 209-237.
II-induced tissue endothelin-1 increase: Role of Steckelings U, Lebrun C, Qadri F, Veltmar A,
blood pressure and endothelial function. Hyper- Unger T. Role of brain angiotensin in cardio-
tension. 1998; 31: 1305-1310. vascular regulation. J Cardiovasc Cardiol. 1992;
Freund BJ, Shizuru EM, Hashiro GM, Claybaugh 19(suppl 6): S72-S79.
JB. Hormonal, electrolytic, and renal responses Tidgren B, Hjemdahl P, Theodorsson E, Nuss-
to exercise are intensity dependent. J Appl berger J. Renal neurohumoral and vascular
Physiol. 1991; 70: 900-906. responses to dynamic exercise in humans. J Appl
Freund BJ, Wade, CE, Claybaugh JR. Effects of Physiol. 1991; 70: 2279-2286.
exercise on atrial natriuretic factor: Release Trice I, Haymes EM. Effects of caffeine ingestion
mechanisms and implications for fluid homeo- on exercise-induced changes during high-
stasis. Sports Med. 1988; 6: 364-377. intensity, intermittent exercise. Int J Sport Nutr.
Gebber GL. Central determinants of sympathetic 1995; 5: 37-44.
nerve discharge. In: Loewy AD, Spyer KM, eds. Tschakovsky ME, Joyner MJ. Nitric oxide and
Central Regulation of Autonomic Functions. New muscle blood flow in exercise. Appl Physiol Nutr
York: Oxford University Press; 1990: 126-144. Metab. 2008; 33: 151-161.
Graham TE, Spriet LL. Performance and meta- Vander A, Sherman J, Luciano D. Human Physiol-
bolic responses to a high caffeine dose during ogy: The Mechanisms of Body Function. 7th ed.
prolonged exercise. J Appl Physiol. 1991; 71: Boston: McGraw-Hill; 1998.
2292-2298. Wisler PL, Green FJ, Watanabe AM. Cardiovas-
Joyner MJ, Halliwill JR. Sympathetic vasodila- cular adrenergic and muscarinic cholinergic
tation in human limbs. J Physiol. 2000; 526: receptors. In: Parmley WW, Chatterjee K, eds.
471-480. Cardiology. Vol 1. Philadelphia: J.B. Lippincott;
Loewy AD, Spyer KM. Central Regulation of Auto- 1992: 1-34.
nomic Functions. New York: Oxford University
Press; 1990. Chapter 4
Lugar A, Watschinger B, Duester P, Svobodo T, Asplund CA, O’Connor FG, Noakes TD. Exercise-
Clodi M, Chrousos GP. Plasma growth hormone associated collapse: An evidence-based review
and prolactin responses to graded levels of acute and primer for clinicians. Br J Sports Med. 2011;
exercise and to lactate infusion. Neuroendocri- 45: 1157-1162.
nol. 1992; 56: 112-117. Bangsbo J, Krustrup P, Gonzalez-Alonso J, Saltin
Maeda S, Miyauchi T, Iemitsu M, Tanabe T, Goto B. ATP production and efficiency of human
K, Yamaguchi I, Matsuda M. Endothelin recep- skeletal muscle during intense exercise: Effect
tor antagonist reverses decreased NO system in of previous exercise. Am J Physiol. 2001; 280:
the kidney in vivo during exercise. Am J Physiol. E956-E964.
2004; 286: E609-E614. Buono MJ, Ball KD, Kolkhorst FW. Sodium ion
Maiorana A, O’Driscoll G, Taylor R, Green D. concentration vs. sweat rate relationship in
Exercise and the nitric oxide vasodilator system. humans. J Appl Physiol. 2007; 103: 990-994.
Sports Med. 2003; 33: 1013-1035. Buono MJ, Claros R, DeBoer T, Wong J. Na+ secre-
Papelier Y, Escourrou P, Gauthier JP, Rowell LB. tion rate increases proportionally more than the
Carotid baroreflex control of blood pressure Na+ reabsorption rate with increases in sweat
and heart rate in man during dynamic exercise. rate. J Appl Physiol. 2008; 105: 1044-1048.
J Appl Physiol. 1994; 77: 502-506. Bürge J, Knechtle B, Knechtle P, Gnädinger M,
Peters S, Kreulen DL. Vasopressin-mediated EPSPs Rüst AC, Rosemann T. Maintained serum
in mammalian sympathetic ganglion. Brain Res. sodium in male ultra-marathoners—The role
1985; 339: 126-129. of fluid intake, vasopressin, and aldosterone in
References 233

fluid and electrolyte regulation. Horm Metab Montain SJ, Coyle EF. Fluid ingestion during
Res. 2011; 43: 646-652. exercise increases skin blood flow independent
Carter R III. Exertional heat illness and hypona- of increases in blood volume. J Appl Physiol.
tremia: An epidemiological perspective. Curr 1992a; 73: 903-910.
Sports Med Rep. 2008; 7: S20-S27. Montain SJ, Coyle EF. Influence of graded dehy-
Costill DL, Cote R, Fink W. Muscle water and elec- dration on hyperthermia and cardiovascular
trolytes following varied levels of dehydration in drift during exercise. J Appl Physiol. 1992b; 73:
man. J Appl Physiol. 1976; 40: 6-11. 1340-1350.
Coyle EF. Fluid and fuel intake during exercise. J Montain SJ, Laird JE, Latzka WA, Sawka MN.
Sports Sci. 2004; 22: 39-55. Aldosterone and vasopressin responses in the
Coyle EF, Gonzalez-Alonzo J. Cardiovascular drift heat: Hydration level and exercise intensity
during prolonged exercise: New perspectives. effects. Med Sci Sports Exer. 1997; 29: 661-668.
Exerc Sports Sci Rev. 2001; 29: 88-92. Murray B, Eichner ER. Hyponatremia of exercise.
Fitzsimmons JT. Angiotensin, thirst, and sodium Curr Sports Med Rep. 2004; 3: 117-118.
appetite. Physiol Rev. 1998; 78: 583-686. Nadel ER, Bullard RW, Stolwojk JAJ. Importance
Fortney SM, Nadel ER, Wenger CB, Bove JR. of skin temperature in the regulation of sweat-
Effect of blood volume on sweating rate and ing. J Appl Physiol. 1971; 31: 80-87.
body fluids in exercising humans. J Appl Physiol. Nielsen M. Die Regulation der Körpertemperatur
1981; 51: 1594-1600. bei Muskelarbeit. Skand Arch Physiol. 1938;
Freund BJ, Shizuru FM, Hashiro GM, Claybaugh 79: 193-230.
JR. Hormonal, electrolyte, and renal responses Noakes T. Fluid replacement during marathon
to exercise are intensity dependent. J Appl. running. Clin J Sport Med. 2003; 13: 309-319.
Physiol. 1991; 70: 900-906. Noakes TD, Sharwood K, Speedy D, Hew T, Reid
Gonzalez-Alonso J, Teller C, Andersen SL, Jensen S, Dugas J, Almond C, Wharam P, Weschler
FB, Hyldig T, Nielsen B. Influence of body tem- L. Three independent biological mechanisms
perature on the development of fatigue during cause exercise-associated hyponatremia: Evi-
prolonged exercise in the heat. J Appl Physiol. dence from 2,135 weighed competitive athletic
1999; 86: 1032-1039. performances. Proc Natl Acad Sci USA. 2005;
Huffman EA, Yard EE, Fields SK, Collins CL, 102: 18550-18555.
Comstock RD. 1. Epidemiology of rare injuries Remick D, Chancellor K, Pederson J, Zambraski
and conditions among United States high school EJ, Sawka MN, Wenger CB. Hyperthermia
athletes during the 2005-2006 and 2006-2007 and dehydration-related deaths associated with
school years. Athl Train. 2008; 43: 624-630. intentional rapid weight loss in three collegiate
Joyner MJ, Halliwill JR. Sympathetic vasodilatation wrestlers—North Carolina, Wisconsin, and
in human limbs. J Physiol. 2000; 526: 471-480. Michigan, November-December 1997. JAMA.
Kellogg DL Jr, Pérgola PE, Piest KL, Kosiba WA, 1998; 279(11): 824-825.
Crandall CG, Grossmann M, Johnson JM. Roddie IC. Circulation to skin and adipose tissue.
Cutaneous active vasodilation in humans in In: Shepherd JT, Abboud FM, eds. Handbook
mediated by cholinergic nerve transmission. of Physiology. Section 2: The Cardiovascular
Circulat Res. 1995; 77: 1222-1228. System. Vol III: Peripheral Circulation and
Maeda S, Miyauchi T, Iemitsu M, Tanabe T, Goto Organ Blood Flow. Bethesda, MD: American
K, Yamaguchi I, Matsuda M. Endothelin recep- Physiologic Society; 1983: 285-317.
tor antagonist reverses decreased NO system in Sato K. The physiology, pharmacology, and bio-
the kidney in vivo during exercise. Am J Physiol. chemistry of the eccrine sweat gland. Rev Physiol
2004; 286: E609-E614. Biochem Pharmacol. 1977; 79: 51-131.
Maughan RJ, Leiper JB, Shirreffs SM. Restoration Sawka MN, Burke LM, Eichner ER, Maughan RJ,
of fluid balance after exercise-induced dehydra- Montain SJ, Stachenfeld NS. American College
tion: Effects of food and fluid intake. Eur J Appl of Sports Medicine position stand. Exercise and
Physiol. 1996; 73: 317-325. fluid replacement. Med Sci Sports Exerc. 2007;
Maughan RJ, Shirreffs SM. Dehydration and 39: 377-390.
rehydration in competitive sport. Scand J Med Shamsuddin AKM, Reddy MM, Quinton PM. Ion-
Sci Sports. 2010; 20(suppl 3): 40-47. tophoretic β-adrenergic stimulation of human
Maughan RJ, Shirreffs SM, Leiper JB. Errors in the sweat glands: Possible assay for cystic fibrosis
estimation of sweat loss and changes in hydra- transmembrane conductance regulator activity
tion status from changes in body mass during in vivo. Exp Physiol. 2008; 93: 969-981.
exercise. J Sports Sci. 2007; 25: 797-804. Shibasaki M, Wilson TE, Crandall CG. Neural
Merson SJ, Maughan RJ, Shirreffs SM. Rehydra- control and mechanisms of eccrine sweating
tion with drinks differing in sodium concen- during heat stress and exercise. J Appl Physiol.
tration and recovery from moderate exercise- 2006; 100: 1692-1701.
induced hypohydration in man. Eur J Appl Shirreffs SM, Maughan RJ. Whole body sweat
Physiol. 2008; 103: 585-594. collection in humans: An improved method with
234 References

preliminary data on electrolyte content. J Appl and skeletal muscle androgen receptor and
Physiol. 1997; 82: 336-341. insulin-like growth factor-I isoform expression
Shirreffs SM, Taylor A, Leiper JB, Maughan RJ. in strength trained men. J Strength Cond Res.
Post-exercise rehydration in man: Effects of 2011; 25: 767-777.
volume consumed and drink sodium content. Arner P, Hellstrom L, Wahrenberg H, Bronnegard
Med Sci Sports Exerc. 1996; 28: 1260-1271. M. Beta-adrenoceptor expression in human fat
Takamata A, Mack GW, Gillen CM, Jozsi AC, cells from different regions. J Clin Invest. 1990;
Nadel ER. Osmoregulatory modulation of 86: 1595-1600.
thermal sweating in humans: Ref lex effects Asada N, Takahashi Y, Wada M, Naito N, Uchida
of drinking. Am J Physiol Regul Integr Comp H, Ikeda M, Honjo M. GH induced lipolysis
Physiol. 1995; 268: R414-R422. stimulation in 3T3-L1 adipocytes stably express-
Takamata A, Mack GW, Gillen CM, Nadel ER. ing hGHR: Analysis on signaling pathway and
Sodium appetite, thirst, and body fluid regu- activity of 20K hGH. Mol Cell Endocrinol. 2000;
lation in humans during rehydration without 162: 121-129.
sodium replacement. Am J Physiol. 1994; 266: Baker JR, Bemben MG, Anderson MA, Bemben
R1493-R1502. DA. Effects of age on testosterone responses
Tarnopolsky MA, Cipriano N, Woodcroft C, to resistance exercise and musculoskeletal
Pulkkinen WJ, Robinson DC, Henderson JM, variables in men. J Strength Con Res. 2006; 20:
MacDougall JD. Effects of rapid weight loss and 874-881.
wrestling on muscle glycogen concentration. Bartness TJ, Song CK. Sympathetic and sensory
Clin J Sport Med. 1996; 6: 78-84. innervation of white adipose tissue. J Lipid Res.
Webster S, Rutt R, Weltman A. Physiological 2007; 48: 1655-1672.
effects of a weight loss regimen practiced by Beaven CM, Hopkins WG, Hansen KT, Wood MR,
college wrestlers. Med Sci Sports Exerc. 1990; Cronin JB, Lowe TE. Dose effect of caffeine on
22: 229-234. testosterone and cortisol responses to resistance
Wenger CB. Heat evaporation of sweat. Thermo- exercise. Int J Sport Nutr Exerc Metab. 2008;
dynamic considerations. J Appl Physiol. 1972; 18: 131-141.
32: 456-459. Birkenfeld AL, Boschmann M, Moro C, Adams
Wenos DL, Amato HK. Weight cycling alters F, Heusser K, Franke G, Berlan M, Luft FC,
muscular strength and endurance, ratings of per- Lafontan M, Jordan J. Lipid mobilization with
ceived exertion, and total body water in college physiological atrial natriuretic peptide concen-
wrestlers. Percept Mot Skills. 1998; 87: 975-978. trations in humans. J Clin Endocrinol Metab.
Wilkerson JE, Gutin B, Horvath SM. Exercise- 2005; 90: 3622-3628.
induced changes in blood, red cell, and plasma Borer KT, Wuorinen EC, Lukos JR, Denver JW,
volumes in men. Med Science Sports Exerc. Porges SW, Burant CF. Two bouts of exercise
1977; 9: 155-158. before meals, but not after meals, lower fasting
Wilson TE, Cui J, Crandall CG. Mean body blood glucose. Med Sci Sports Exerc. 2009; 41:
temperature does not modulate eccrine sweat 1606-1614.
rate during upright tilt. J Appl Physiol. 2005; Campbell PJ, Carlson MG, Hill JO, Nurjhan N.
98: 1207-1212. Regulation of free fatty acid metabolism by
insulin in humans: Role of lipolysis and reesteri-
Chapter 5 fication. Am J Physiol. 1992; 263: E1063-E1069.
Ahlborg G, Felig P. Influence of glucose ingestion Cartee GD, Funai K. Exercise and insulin: Con-
on fuel-hormone response during prolonged vergence or divergence at AS160 and TBC1D1?
exercise. J Appl Physiol. 1976; 41: 683-688. Exer Sport Sci Rev. 2009; 37: 188-195.
Ahlborg G, Felig P. Substrate utilization during Casanueva FF, Villanueva L, Dieguez C, Diaz Y,
prolonged exercise preceded by ingestion of Cabranes JA, Szoke B, Scanlon MF, Schally
glucose. Am J Physiol. 1977; 233: E188-E194. AV, Frenandez-Cruz A. Free fatty acids block
Ahlborg G, Felig P. Lactate and glucose exchange growth hormone (GH) releasing hormone-
across the forearm, legs, and splanchnic bed stimulated GH secretion in man directly at the
during and after prolonged leg exercise. J Clin pituitary. J Clin Endocr Metab. 1987; 64: 634-642.
Invest. 1982; 69: 45-54. Casey A, Short AH, Hultman E, Greenhaff PL.
Ahlborg G, Felig P, Hagenfeldt L, Hendler R, Glycogen resynthesis in human muscle fibre
Wahren J. Substrate turnover during prolonged types following exercise-induced glycogen deple-
exercise in man: Splanchnic leg metabolism of tion. J Physiol. 1995; 483: 265-271.
glucose, free fatty acid and amino acids. J Clin Chalmers RJ, Bloom SR, Duncan G, Johnson RH,
Invest. 1974; 53: 1080-1090. Sulaiman WR. The effect of somatostatin on
Ahtiainen JP, Lehti M, Hulmi JJ, Kraemer WJ, metabolic and hormonal changes during and
Alen M, Nyman K, Selänne H, Pakarinen A, after exercise. Clin Endocrinol. 1979; 10: 451-458.
Komulainen J, Kovanen V, Mero AA, Häkki- Chiasson JL, Liljenquist JE, Finger JE, Lacy WW.
nen K. Recovery after heavy resistance exercise Differential sensitivity of glycogenolysis and
References 235

gluconeogenesis to insulin infusions in dogs. Fatouros I, Chatzinikolaous A, Tournis S, Niko-


Diabetes. 1976; 25: 283. laidis MG, Jamurtaqs AZ, Douroudos II,
Clark C, Newgard CB. Hepatic regulation of Papassotiriou I, Thomakos PM, Taxildaris
fuel metabolism. In: Saltiel AR, Pessin JE, K, Mastorakos G, Mitrakou A. Intensity of
eds. Mechanisms of Insulin Action. New York: resistance exercise determines adipokine and
Springer Science and Business Media; 2007: resting energy expenditure responses on over-
90-109. weight elderly individuals. Diabetes Care. 2009;
Coggan AR, Coyle EF. Metabolism and perfor- 32: 2161-2167.
mance following carbohydrate ingestion late in Febbraio MA, Hiscock N, Saccherri M, Fischer
exercise. Med Sci Sports Exerc. 1989; 21: 59-65. CP, Pedersen BK. Interleukin-6 is a novel
Coyle EF, Coggan AR, Hammert MK, Ivy JL. factor mediating glucose homeostasis during
Muscle glycogen utilization during prolonged skeletal muscle contraction. Diabetes. 2004; 53:
strenuous exercise when fed carbohydrate. J 1643-1648.
Appl Physiol. 1986; 61: 165-172. Felig P. The glucose-alanine cycle. Metabolism.
Dodt C, Lonnroth P, Wellhoner JP, Fehm HL, Elam 1973; 22: 179-207.
M. Sympathetic control of white adipose tissue Fell RD, McLane JA, Winder WW, Holloszy JO.
in lean and obese women. Acta Physiol Scand. Preferential resynthesis of muscle glycogen
2003; 177: 351-357. in fasting rats after exhausting exercise. Am J
Donsmark M, Langfort J, Holm C, Ploug T, Galbo Physiol. 1980; 238: R328-R332.
H. Hormone-sensitive lipase as mediator of Fleck SJ, Kraemer WJ. Designing Resistance Train-
lipolysis in contracting skeletal muscle.  Exerc ing Programs. 3rd ed. Champaign, IL: Human
Sport Sci Rev. 2005; 33: 127-133. Kinetics; 2004.
Dovio A, Roveda E, Sciolla C, Montarrulli A, Fluckey JD, Hickey MS, Brambrink JK, Hart KK,
Raffaelli A, Saba A, Calogiuri G, De Francia S, Alexander K, Craig BW. Effects of resistance
Borrione P, Salvadori P, Carandente F, Angeli exercise on glucose tolerance in normal and
A. Intense physical exercise increases systemic glucose-intolerant subjects. J Appl Physiol. 1994;
11beta-hydroxysteroid dehydrogenase type 1 77: 1087-1092.
activity in healthy adult subjects. Eur J Appl Franz IW, Lohman FW, Koch G, Quabbe HJ.
Physiol. 2010; 108: 681-687. Aspects of hormonal regulation of lipolysis
Dreyer HC, Drummond MJ, Glynn EL, Fujita S, during exercise: Effects of chronic beta-receptor
Chionkes DL, Volpi E, Rasmussen BB. Resis- blockade. Int J Sport Med. 1983; 4: 14-20.
tance exercise increases human skeletal muscle French DN, Kraemer WJ, Volek JS, Spiering BA,
AS160/TBC1D4 phosphorylation in associa- Judelson DA, Hoffman JR, Maresh CM. Antici-
tion with enhanced leg glucose uptake during patory responses of catecholamines on muscle
postexercise recovery. J Appl Physiol. 2008; force production. J Appl Physiol. 2007; 102: 94-102.
105: 1967-1974. Funai K, Cartee GD. Inhibition of contraction-
Enevoldsen LH, Polak J, Simonsen L, Hammer T, stimulated AMP-activated protein kinase
Macdonald I, Crampes F, de Glisezinski I, Stich inhibits contraction-stimulated increases in
V, Bülow J. Post-exercise abdominal subcutane- PAS-TBC1D1 and glucose transport without
ous adipose tissue lipolysis in fasting subjects altering PAS-AS160 in rat skeletal muscle. Dia-
is inhibited by infusion of the somatostatin betes. 2009; 58: 1096-1104.
analogue octreotide. Clin Physiol Funct Imaging. Galbo H, Christensen NJ, Holst JJ. Catecholamines
2007; 27: 320-326. and pancreatic hormones during autonomic
Eriksson H, Ridderstrale M, Degerman E, Ekholm blockade in exercising man. Acta Physiol Scand.
D, Smith CJ, Manganiello VC, Belfrage P, 1977a; 101: 428-437.
Tornqvist H. Evidence for the key role of the Galbo H, Christensen NJ, Holst JJ. Glucose-
adipocyte cGMP phosphodiesterase in the induced decrease in glucagon and epinephrine
antilipolytic action of insulin. Biochim Biophys responses to exercise in man. J Appl Physiol.
Acta. 1995; 1266: 101-107. 1977b; 42: 525-530.
Evans SA, Doblado M, Chi MM, Corbett JA, Moley Galster AD, Clutter WE, Cryer PE, Collins JA,
KH. Facilitative glucose transporter 9 expres- Bier DM. Epinephrine plasma thresholds for
sion affects glucose sensing in pancreatic βcells. lipolytic effects in man: Measurements of fatty
Endocrinol. 2009; 150: 5302-5310. acid transport with [1-13C] palmitic acid. J Clin
Fatouros I, Chatzinikolaous A, Paltoglou G, Petri- Invest. 1981; 67: 1729-1738.
dou A, Avloniti A, Jamurtas A, Goussetis E, Gerich JE. Glucose counterregulation and its
Mitrakou A, Mougios V, Lazaropoulou C, Mar- impact on diabetes mellitus. Diabetes. 1988; 37:
geli A, Papassotiriou I, Mastorakos G. Acute 1608-1617.
resistance exercise results in catecholaminergic Goto K, Maemura H, Takamatsu K, Ishii N.
rather than hypothalamic-pituitary-adrenal axis Hormonal responses to resistance exercise after
stimulation during exercise in young men. Stress. ingestion of carnosine and anserine. J Strength
2010; 13: 461-468. Cond Res. 2011; 25: 398-405.
236 References

Granneman JG. Why do adipocytes make β3 of one year recombinant human growth hor-
adrenergic receptor? Cell Signaling. 1995; 7: 9-15. mone therapy. J Endocrinol. 2001; 171: 285-292.
Gravholt CH, Schmitz O, Simonsen L, Bulow J, Kietzman T, Porwol T, Zierold K, Jungermann
Christiansen JS, Moller N. Effects of a physi- K, Acker H. Involvement of a local fenton
ological GH pulse on interstitial glycerol in reaction in the reciprocal modulation by O2
abdominal and femoral adipose tissue. Am J of the glucagon-dependent activation of the
Physiol. 1999; 277: E848-E854. phosphoenolpyruvate carboxykinase gene and
Groop LC, Bonadonna RC, Simonson DC, Petrides the insulin-dependent activation of the glucoki-
AS, Shank M, DeFronzo RA. Effect of insulin nase gene in rat hepatocytes. Biochem J. 1998;
on oxidative and nonoxidative pathways of free 335: 425-432.
fatty acid metabolism in human obesity. Am J Kjaer M, Howlett K, Langfort J, Zimmerman-
Physiol. 1992; 263: E79-E84. Belsing T, Lorentsen J, Bulow J, Ihlemann J,
Hagberg JM, Hickson RC, McLane JA, Ehsani AA, Feldt-Rasmussen U, Galbo H. Adrenaline and
Winder WW. Disappearance of norepinephrine glycogenolysis in skeletal muscle during exercise:
from circulation following strenuous exercise. J A study in adrenalectomised humans. J Physiol.
Appl Physiol. 1979; 47: 1311-1314. 2000; 528: 371-378.
Henderson GC, Fattor JA, Horning MA, Faghih- Koeslag JH. Post-exercise ketosis and the hormone
nia M, Johnson ML, Mau TL, Luke-Zeitoun M, response to exercise: A review. Med Sci Sports
Brooks GA. Lipolysis and fatty acid metabolism Exerc. 1982; 14: 327-334.
in men and women during the postexercise Kon M, Ikeda T, Homma T, Akimoto T, Suzuki Y,
recovery period. J Physiol. 2007; 584: 963-981. Kawahara T. Effects of acute hypoxia on meta-
Horowitz J, Klein S. Lipid metabolism during bolic and hormonal responses to resistance exer-
endurance exercise. Am J Clin Nutr. 2000; cise. Med Sci Sports Exerc. 2010; 42: 1279-1285.
72(suppl): 558S-563S. Koopman R, Manders RJ, Zorenc AH, Hul GB,
Horowitz JF, Coyle EF. Metabolic responses to Kuipers H, Keizer HA, van Loon LJ. A single
preexercise meals containing various carbohy- session of resistance exercise enhances insulin
drates and fat. Am J Clin Nutr. 1993; 58: 235-241. sensitivity for at least 24 h in healthy men. Eur
Howlett KF, Sakamoto K, Garnham A, Cameron- J Appl Physiol. 2005; 94: 180-187.
Smith D, Hargreaves M. Resistance exercise Koziris LP, Hickson RC, Chatterton RT Jr, Groseth
and insulin regulate AS160 and interaction with RT, Christie JM, Goldflies DG, Unterman TG.
14-3-3 in human skeletal muscle. Diabetes. 2007; Serum levels of total and free IGF-I and IGFBP-3
56: 1608-1614. are increased and maintained in long-term train-
Howlett RA, Parolin ML, Dyck DJ, Hultman E, ing. J Appl Physiol. 1999; 86: 1436-1442.
Jones NL, Heigenhauser GJF, Spriet LL. Regu- Kraemer WJ, Gordon SE, Fleck SJ, Marchitelli
lation of skeletal muscle glycogen phosphory- LJ, Mello R, Dziados JE, Friedl K, Harman
lase and pyruvate dehydrogenase at different E, Maresh C, Fry AC. Endogenous anabolic
exercise power outputs. Am J Physiol. 1998; hormonal and growth factor responses to heavy
275: R418-R425. resistance exercise in males and females. Int J
Iizuka K, Horikawa Y. ChREBP: A glucose- Sports Med. 1991; 12: 228-235.
activated transcription factors involved in the Kraemer WJ, Häkkinen K, Newton RU, Nindl
development of metabolic syndrome. Endocr J. BC, Volek JS, McCormick M, Gotshalk LA,
2008; 55: 617-624. Gordon SE, Fleck SJ, Campbell WW, Putukian
Jenkins PJ. Growth hormone and exercise. Clin M, Evans WJ. Effects of heavy-resistance train-
Endocrinol. 1999; 50: 683-689. ing on hormonal response patterns in younger
Jensen MD. Role of body fat distribution and vs. older men. J Appl Physiol. 1999; 87: 982-992.
the metabolic complications of obesity. J Clin Kraemer WJ, Marchitelli L, Gordon SE, Harman
Endocr Metab. 2008; 93: S57-S63. E, Dziados JE, Mello R, Frykman P, McCurry
Jimenez C, Santiago M, Sitler M, Boden G, Homko D, Fleck SJ. Hormonal and growth factor
C. Insulin-sensitivity response to a single bout responses to heavy resistance exercise protocols.
of resistive exercise in type 1 diabetes mellitus. J Appl Physiol. 1990; 69: 1442-1450.
J Sport Rehabil. 2009; 18: 564-571. Kraemer WJ, Nindl BC, Volek JS, Marx JO,
Judelson DA, Maresh CM, Yamamoto LM, Far- Gotschalk LA, Bush JA, Welsch JR, Vingren JL,
rell MJ, Armstrong LE, Kraemer WJ, Volek JS, Spiering BA, Fragala MS, Hatfield DL, Ho JY,
Spiering BA, Casa DJ, Anderson JM. Effect of Maresh CM, Mastro AM, Hymer WC. Influence
hydration state on resistance exercise-induced of oral contraceptive use on growth hormone in
endocrine markers of anabolism, catabolism, and vivo bioactivity following resistance exercise:
metabolism. J Appl Physiol. 2008; 105: 816-824. Responses of molecular mass variants. Growth
Khalfallah Y, Sassolas G, Borson-Chazot F, Vega Horm IGF Res. 2008; 18: 238-244.
N, Vidal H. Expression of insulin target genes Kraemer WJ, Ratamess NA. Hormonal responses
in skeletal muscle and adipose tissue in adult and adaptations to resistance exercise and train-
patients with growth hormone deficiency: Effect ing. Sports Med. 2005; 35: 339-361.
References 237

Krag MB, Gormsen LC, Guo Z, Christiansen JS, McCall GE, Goulet RE, Roy RR, Grindeland
Jensen MD, Nielsen S, Jørgensen JO. Growth JE, Boorman GI, Bigbee AJ, Hodgson JA,
hormone-induced insulin resistance is associ- Greenisen MC, Edgerton VR. Spacef light
ated with increased intramyocellular triglyc- suppresses exercise-induced release of bioas-
eride content but unaltered VLDL-triglyceride sayable growth hormone. J Appl Physiol. 1999;
kinetics. Am J Physiol Endocrinol Metab. 2007; 87: 1207-1212.
292: E920-E927. McCaulley GO, McBride JM, Cormie P, Hudson
Krief S, Lonnqvist F, Raimbault S, Baude B, Van MD, Nuzzo JL, Quindry JC, Travis Triplett N.
Spronsen A, Arner P, Strosberg AD, Ricquier Acute hormonal and neuromuscular responses
D, Emorine LJ. Tissue distribution of beta to hypertrophy, strength and power type resis-
3-adrenergic receptor mRNA in man. J Clin tance exercise. Eur J Appl Physiol. 2009; 105:
Invest. 1993; 91: 344-349. 695-704.
Lafontan M, Langin D. Lipolysis and lipid mobili- Mikines KJ, Sonne B, Farrell PA, Tronier B, Galbo
zation in human adipose tissue. Progress Lipid H. Effect of physical exercise on sensitivity
Res. 2009; 48: 275-297. and responsiveness to insulin in humans. Am J
Langfort J, Ploug T, Ihlemann J, Baranczuk E, Physiol. 1988; 254: E248-E259.
Donsmark M, Gorski J, Galbo H. Additivity of Miller AD, Ruby BC, Laskin JJ, Gaskill SE. Effects
adrenaline and contractions on hormone-sensi- of high intensity/low volume and low intensity/
tive lipase, but not on glycogen phosphorylase, in high volume isokinetic resistance exercise on
rat muscle. Acta Physiol Scand. 2003; 178: 51-60. postexercise glucose tolerance. J Strength Cond
Lavoie C, Ducros F, Bourque J, Langelier H, Chi- Res. 2007; 21: 330-335.
asson J-L. Glucose metabolism during exercise Møller N, Jorgenson JOL. Effects of growth hor-
in man: The role of insulin and glucagon in the mone on glucose, lipid, and protein metabolism
regulation of hepatic glucose production and in human subjects. Endocr Rev. 2009; 30: 152-177.
glucoeneogenesis. Can J Physiol Pharmacol. Mounier C, Posner BJ. Transcriptional regulation
1997; 75: 26-35. by insulin: From the receptor to the gene. Can J
Leite RD, Prestes J, Rosa C, De Salles BF, Maior Physiol Pharmacol. 2006; 84: 713-724.
A, Miranda H, Simão R. Acute effect of resis- Nielsen S, Guo ZK, Johnson CM, Hensrud DD,
tance training volume on hormonal responses Jensen MD. Splanchnic lipolysis in human obe-
in trained men. J Sports Med Phys Fitness. 2011; sity. J Clin Invest. 2004; 113: 1582-1588.
51: 322-328. Nindl BC, Pierce JR. Insulin-like growth factor I
Lins P, Wajngot A, Adamson U, Vranic M, Efendic as a biomarker of health, fitness, and training
S. Minimal increases in glucagon levels enhance status. Med Sci Sports Exerc. 2010; 42: 39-49.
glucose production in man with partial insulin Nurjhan N, Campbell PJ, Kennedy FP, Miles JM,
deficiency. Diabetes. 1983; 32: 633-636. Gerich JE. Insulin dose-response characteristics
Lithell H, Schele R, Vessby B, Jacobs I. Lipopro- for suppression of glycerol release and conver-
teins, lipoprotein lipase, and glycogen after sion to glucose in humans. Diabetes. 1986; 35:
prolonged physical activity. J Appl Physiol. 1326-1331.
1984; 57: 698-702. Pafili ZK, Bogdanis GC, Tsetsonis NV, Maridaki
Magkos F, Tsekouras YE, Prentzas KI, Basioukas M. Postprandial lipemia 16 and 40 hours after
KN, Matsama SG, Yanni AE, Kavouras SA, low-volume eccentric resistance exercise. Med
Sidossis LS. Acute exercise-induced changes Sci Sports Exerc. 2009; 41: 375-382.
in basal VLDL-triglyceride kinetics leading Pan DA, Lillioja S, Kriketos AD, Milner MR,
to hypotriglyceridemia manifest more readily Baur LA, Bogardus C, Jenkins AB, Storlien LH.
after resistance than endurance exercise. J Appl Skeletal muscle triglyceride levels are inversely
Physiol. 2008; 105: 1228-1236. related to insulin action. Diabetes. 1997; 46:
Manetta J, Brun JF, Maimoun L, Callis A, Prefaut 983-988.
C, Mercier J. Effect of training on the GH/IGF-I Pedersen BK, Febbraio MA. Muscle as an endo-
axis during exercise in middle-aged men: Rela- crine organ: Focus on muscle-derived interleu-
tionship to glucose homeostasis. Am J Physiol kin-6. Physiol Rev. 2008; 88: 1379-1406.
Endocrinol Metab. 2002; 283: E929-E936. Porte D, Graber AL, Kuzuya T, Williams RH.
Marliss EB, Simantirakis E, Purdon C, Gougeon The effects of epinephrine on immunoreactive
R, Field CJ, Halter JB, Vranic M. Glucoregula- insulin levels in man: 1965. J Clin Invest. 1966;
tory and hormonal responses to repeated bouts 45: 228-236.
of intense exercise in normal male subjects. J Price TB, Rothman DL, Taylor R, Avison MJ, Shul-
Appl Physiol. 1991; 71: 924-933. man GI, Shulman RG. Human muscle glycogen
McCall GE, Goulet RE, Grindeland JE, Hodgson resynthesis after exercise: Insulin dependent
JA, Bigbee AJ, Edgerton VR. Bed rest sup- and independent phases. J Appl Physiol. 1994;
presses bioassayable growth hormone release 76: 104-111.
in response to muscle activity. J Appl Physiol. Pritzlaff CJ, Wideman L, Blumer J, Jensen M,
1997; 83: 2086-2090. Abbott RD, Gaesser GA, Veldhuis JD, Weltman
238 References

A. Catecholamine release, growth hormone transporter GLUT3: 20 years of distinction. Am


secretion, and energy expenditure during exer- J Physiol. 2008; 295: E242-E253.
cise vs. recovery in men, J Appl Physiol. 2000; Singhal A, Trilk JL, Jenkins NT, Bigelman KA,
89: 937-946. Cureton KJ. Effect of intensity of resistance
Pritzlaff CJ, Wideman L, Weltman JY, Abbott exercise on postprandial lipemia. J Appl Physiol.
RD, Gutgesell ME, Hartman ML, Veldhuis JD, 2009; 106: 823-829.
Weltman A. Impact of acute exercise intensity Smith TR, Elmendorf JS, David TS, Turinsky J.
on pulsatile growth hormone release in men. J Growth-hormone-induced insulin resistance:
Appl Physiol. 1999; 87: 498-504. Role of insulin receptor, IRS-1, GLUT-1, and
Radenne A, Akpa M, Martel C, Sawadiogo S, GLUT-4. Am J Physiol. 1997; 272: E1071-E1079.
Mauvoisin D, Mounier C. Hepatic regulation of Sotsky MJ, Shilo S, Shamoon H. Regulation of
fatty acid synthase by insulin and T3: Evidence counterregulatory hormone secretion in man
for T3 genomic and nongenomic actions. Am J during exercise and hypoglycemia. J Clin Endo-
Physiol. 2008; 295: E884-E894. crinol Metab. 1989; 68: 9-16.
Randle PJ, Newsholme EA, Hales CN. The glucose Spiering BA, Kraemer WJ, Vingren JL, Ratamess
fatty acid cycle in obesity and maturity onset NA, Anderson JM, Armstrong LE, Nindl BC,
diabetes mellitus. Ann New York Acad Sci. 1965; Volek JS, Hakkinen K, Maresh CM. Elevated
131: 324-333. endogenous testosterone concentrations poten-
Rarick KR, Pikosky MA, Grediagin A, Smith TJ, tiate muscle androgen receptor responses to
Glickman EL, Alemany JA, Staab JS, Young AJ, resistance exercise. J Steroid Biochem Mol Biol.
Nindl BC. Energy flux, more so than energy bal- 2009; 114: 195-199.
ance, protein intake, or fitness level, influences Surya S, Horowitz JF, Goldenberg N, Sakharova
insulin-like growth factor-I system responses A, Harber M, Cornford AS, Symons K, Barkan
during 7 days of increased physical activity. J AL. The pattern of growth hormone delivery
Appl Physiol. 2007; 103: 1613-1621. to peripheral tissues determines insulin-like
Reeves GV, Kraemer RR, Hollander DB, Clavier growth factor-1 and lipolytic responses in obese
J, Thomas C, Francois M, Castracane VD. subjects.  J Clin Endocrinol Metab. 2009; 94:
Comparison of hormone responses following 2828-2834.
light resistance exercise with partial vascular Thamer C, Machann J, Tschritter O, Haap M,
occlusion and moderately difficult resistance Wietek B, Dahl D, Bachmann O, Fritsche A,
exercise without occlusion. J Appl Physiol. 2006; Jacob S, Stumvoll M, Schick F, Haring H-U.
101: 1616-1622. Relationship between serum adiponectin con-
Roberts MD, Dalbo VJ, Hassell SE, Kerksick CM. centration and intramyocellular lipid stores in
The expression of androgen-regulated genes humans. Horm Metab Res. 2002; 34: 646-649.
before and after a resistance exercise bout in Thirone ACP, Huang C, Klip A. Tissue-specific
younger and older men. J Strength Cond Res. roles of IRS proteins in insulin signaling and
2009; 23: 1060-1067. glucose transport. Trends Endocr Metab. 2006;
Romijn JA, Coyle EF, Sidossis LS, Gastaldelli A, 17: 725-727.
Horowitz JF, Endert E, Wolfe RR. Regulation of Tomas E, Kelly M, Xiang X, Tsao T-S, Keller C,
endogenous fat and carbohydrate metabolism in Keller P, Luo Z, Lodish H, Saha AK, Unger R,
relation to exercise intensity and duration. Am J Ruderman NB. Metabolic and hormonal inter-
Physiol Endocrinol Metab. 1993; 265: E380-E391. actions between muscle and adipose tissue. Proc
Romijn JA, Coyle EF, Sidossis LS, Rosenblatt J, Nutr Soc. 2004; 63: 381-385.
Wolfe RR. Substrate metabolism during dif- Toniguchi CM, Meanuelli B, Kahn CR. Critical
ferent exercise intensities in endurance-trained nodes in signaling pathways: Insights into insulin
women. J Appl Physiol. 2000; 88: 1707-1714. action. Nature Rev Mol Cell Biol. 2006; 7: 85-96.
Sasaki K, Cripe TP, Koch SR, Andreone TL, Tuckow AP, Rarick KR, Kraemer WJ, Marx JO,
Petersen DD, Beale EG, Granner DK. Mul- Hymer WC, Nindle BC. Nocturnal growth
tihormonal regulation of phosphoenolpyru- hormone secretory dynamics are altered after
vate carboxykinase gene transcription. The resistance exercise: Deconvolution analysis of
dominant role of insulin. J Bio Chem. 1984; 259: 12-hour immunofunctional and immunoreactive
15242-15251. isoforms. Am J Physiol. 2006; 291: R1749-R1755.
Sengenes C, Berlan M, De Glisezinski I, Lafontan Uchida MC, Crewther BT, Ugrinowitsch C,
M, Galitzky J. Natriuretic peptides: A new Bacurau RF, Moriscot AS, Aoki MS. Hor-
lipolytic pathway in human adipocytes. FASEB monal responses to different resistance exercise
J. 2000; 14: 1345-1351. schemes of similar total volume. J Strength Cond
Sharp CP, Pearson DR. Amino acid supplements Res. 2009; 23: 2003-2008.
and recovery from high-intensity resistance Van Schaftingen E, Gerin I. The glucose-6-phos-
training. J Strength Cond Res. 2010; 24: 1125-1130. phatase system. Biochem J. 2002; 362: 513-532.
Simpson IA, Dwyer D, Malide D, Moley KH, Vingren JL, Kraemer WJ, Hatfield DL, Volek
Travis A, Vannucci SJ. The facilitative glucose JS, Ratamess NA, Anderson JM, Häkkinen
References 239

K, Ahtiainen J, Fragala MS, Thomas GA, Ho Borer KT. Absence of weight regulation in exercis-
JY, Maresh CM. Effect of resistance exercise ing hamsters. Physiol Behav. 1974; 12: 589-597.
on muscle steroid receptor protein content in Borer KT. Nonhomeostatic control of human appe-
strength-trained men and women. Steroids. tite and physical activity in regulation of energy
2009; 74: 1033-1039. balance. Exer Sports Sci Rev. 2010; 38: 114-121.
Wahren J, Felig P, Hendler R, Ahlborg G. Glucose Borer KT, Bonna R, Kielb M. Hippocampal sero-
and amino acid metabolism during recovery tonin mediates hypoactivity in dietarily obese
after exercise. J Appl Physiol. 1973; 34: 838-845. hamsters: A possible manifestation of aging?
Wasserman DH, Cherrington AD. Regulation Pharm Biochem Behav. 1989; 31: 885-892.
of extramuscular fuel sources during exercise. Borer KT, Kaplan LR. Exercise-induced growth
In: Rowell LB, Shephard JT, eds. Handbook in golden hamsters: Effects of age, weight, and
of Physiology. Section 12: Exercise: Regulation activity level. Physiol Behav. 1977; 18: 29-34.
and Integration of Multiple Systems. New York: Borer KT, Kelch RP, White MP, Dolson L, Kuhns
Oxford University Press; 1996: 1036-1074. LR. The role of septal area in the neuroen-
Wee J, Charlton C, Simpson H, Jackson NC, docrine control of growth in the adult golden
Shojaee-Moradie F, Stolinski M, Pentecost C, hamster. Neuroendocrinology. 1977; 23: 133-150.
Umpleby AM. GH secretion in acute exercise Borer KT, Kuhns LR. Radiographic evidence for
may result in post-exercise lipolysis. Growth acceleration of skeletal growth in adult hamsters
Horm IGF Res. 2005; 15: 397-404. by exercise. Growth. 1977; 41: 1-13.
West DW, Burd NA, Tang JE, Moore DR, Staples Borer KT, Nicoski DR, Owens V. Alteration of
AW, Holwerda AM, Baker SK, Phillips SM. pulsatile growth hormone secretion by growth-
Elevations ion ostensibly anabolic hormones inducing exercise: Involvement of endogenous
with resistance exercise enhance neither train- opiates and somatostatin. Endocrinology. 1986;
ing-induced muscle hypertrophy nor strength of 118: 844-850.
the elbow flexors. J Appl Physiol. 2010; 108: 60-67. Borer KT, Peters NL, Kelch RP, Tsai AC, Holder
Zafeiridis A, Goloi E, Petridou A, Dipla K, Mou- S. Contribution of growth, fatness, and activity
gios V, Lellis S. Effects of low- and high-volume to weight disturbance after septohypothalamic
resistance exercise on postprandial lipaemia. Br cuts in adult hamsters. J Comp Physiol Psychol.
J Nutr. 2007; 97: 471-477. 1979; 93: 907-913.
Zang Y, Wang T, Xie W, Wang-Fischer Y, Getty Borer KT, Potter CD, Fileccia N. Basis for the
L, Han Y, Corkey BE, Guo W. Regulation of hypoactivity that accompanies rapid weight gain
acetyl CoA carboxylase and carnitine palmitoyl in hamsters. Physiol Behav. 1983; 30: 389-397.
transferase-1 in rat adipocytes. Obes Res. 2005; Borer KT, Wuorinen E, Ku K. Appetite responds
13: 1530-1539. to changes in meal content while ghrelin, leptin,
Zhao JT, Cowley MJ, Lee P, Birzniece V, Kaplan and insulin track changes in energy availability.
W, Ho KK. Identification of novel GH-regulated J Clin Endoc Metab. 2009; 94: 2290-2298.
pathway of lipid metabolism in adipose tissue: Browne SA, Borer KT. Basis for exercise-induced
A gene expression study in hypopituitary men. hyperphagia in adult hamsters. Physiol Behav.
J Clin Endocrinol Metab. 2011; 96: E1188-E1196. 1978; 20: 553-557.
Cabanac M, Duclos R. Olfactory-gustatory allies-
Chapter 6 thesia and food intake in rats. J Physiol (Paris).
Adolph EF. Urges to eat and drink in rats. Am J 1973; 66: 113-135.
Physiol. 1947; 151: 110-125. Cannon WB, Washburn AL. An explanation of
Anderson JW, Konz EC, Frederich RC, Wood hunger. Am J Physiol. 1912; 29: 441-454.
CL. Long-term weight-loss maintenance: A Casper RC. The “drive for activity” and “restless-
meta-analysis of US studies. Am J Clin Nutr. ness” in anorexia nervosa: Potential pathways.
2001; 74: 579-584. J Affect Disord. 2006; 92: 99-107.
Berridge KC. “Liking” and “wanting” food Catenacci VA, Wyatt HR. The role of physical activ-
rewards: Brain substrates and roles in eating ity in producing and maintaining weight loss.
disorders. Physiol Behav. 2009; 97: 537-550. Nature Clin Pract: Endocr Metab. 2007; 3: 518-529.
Berthoud HR. Vagal and hormonal gut-brain com- Chakravarthy MV, Booth FW. Eating, exercise, and
munication: From satiation to satisfaction. Neu- “thrifty” genotypes: Connecting the dots toward
rogastroenterol Motility. 2008; 20(suppl 1): 64-72. an evolutionary understanding of modern
Berthoud HR, Jeanrenaud B. Acute hyperin- chronic diseases. J Appl Physiol. 2004; 96: 3-10.
sulinemia and its reversal by vagotomy after Chen C-Y, Fujimiya M, Asakawa A, Chang F-Y,
lesions of the ventromedial hypothalamus in Cheng J-T, Lee S-D, Inui A. At the cutting edge:
anesthetized rats. Endocrinology. 1979; 105: Ghrelin gene products in food intake and gut
146-151. motility. Neuroendocrinology. 2009; 89: 9-17.
Borer KT. Control of food intake in Octopus bria- Cluskey M, Grobe D. College weight gain and
reus Robson. J Comp Physiol Psychol. 1971; 75: behavior transitions: Male and female differ-
171-185. ences. J Am Diet Assoc. 2009; 109: 325-329.
240 References

Coleman DL. Effects of parabiosis of obese with McCamish MA, O’Rahilly S. Effects of recom-
diabetes and normal mice. Diabetologia. 1973; binant leptin therapy in a child with congenital
9: 294-298. leptin deficiency. New Engl J Med. 1999; 341:
Coleman DL, Hummel KP. Effects of parabiosis 879-884.
of normal with genetically diabetic mice. Am J Figlewicz DP, Benoit SC. Insulin, leptin, and food
Physiol. 1969; 217: 1298-1304. reward: Update 2008. Am J Physiol. 2009; 296:
Considine RV, Sinha MK, Heiman ML, Kriauci- R9-R19.
unas A, Stephens TW, Nyce MR, Ohannesian Flegal KM, Carroll MD, Kit BK, Ogden CL.
JP, Marco CC, McKee LJ, Bauer TL, Caro JF. Prevalence of obesity and trends in body mass
Serum immunoreactive leptin concentrations in index among US adults, 1999-2010. JAMA. 2012;
normal-weight and obese humans. New Engl J 307: 491-497.
Med. 1996; 334: 221-229. Franz MJ, VanWormer JJ, Crain AL, Boucher JL,
Cordain L, Gotshall RW, Eaton SB, Eaton SB III. Histon T, Caplan W, Bowman JD, Pronk NP.
Physical activity, energy expenditure and fitness: Weight-loss outcomes: A systematic review and
An evolutionary perspective. Int J Sports Med. meta-analysis of weight-loss clinical trials with
1998; 19(5): 328-335. a minimum 1-year follow-up. Am Diet Assoc.
Dethier VG, Solomon RL, Turner LH. Sensory 2007; 107: 1755-1767.
input and central excitation and inhibition in Frederich RC, Hamann A, Anderson S, Löllmann
the blowfly. J Comp Physiol Psychol. 1965; 60: B, Lowell BB, Flier JS. Leptin levels reflect body
303-313. lipid content in mice: Evidence for diet-induced
Ducy P, Amling M, Takeda S, Priemel M, Schil- resistance to leptin action. Nat Med. 1995; 1:
ling AF, Beil FT, Shen J, Vinson C, Rueger JM, 1311-1314.
Karsenty G. Leptin inhibits bone formation Fulton S, Woodside B, Shizgal P. Modulation of
through a hypothalamic relay: A central control brain reward circuitry by leptin. Science. 2000;
of bone mass. Cell. 2000; 100: 197-207. 287: 125-128.
Eaton SB. The ancestral human diet: What was it Gao Q, Horvath TL. Neuronal control of energy
and should it be a paradigm for contemporary homeostasis. FEBS Letters. 2008; 582: 132-141.
nutrition? Proc Nutr Soc. 2006; 65: 1-6. Gautron L, Elmquist JK. Sixteen years and count-
Eaton SB, Cordain L, Sparling PB. Evolution, body ing: An update on leptin in energy balance. J
composition, insulin receptor competition, and Clin Invest. 2011; 121: 2087-2093.
insulin resistance. Prev Med. 2009; 49: 283-285. Guyenet SJ, Schwartz MW. Regulation of food
Eaton SB, Eaton SB III. An evolutionary perspec- intake, energy balance, and body fat mass: Impli-
tive on human physical activity: Implications cations for the pathogenesis and treatment of obe-
for health. Comp Biochem Physiol A Mol Integr sity. J Clin Endocrinol Metab. 2012; 97: 745-755.
Physiol. 2003; 136: 153-159. Hervey GR. The effects of lesions in the hypothala-
Ehrlich S, Burghardt R, Schneider N, Broecker- mus in parabiotic rats. J Physiol (Lond). 1959;
Preuss M, Weiss D, Merle JV, Craciun EM, 145: 336-352.
Pfeiffer E, Mann K, Lehmkuhl U, Hebebrand Heymsfield SB, Greenberg AS, Fujioka K, Dixon
J. The role of leptin and cortisol in hyperactiv- RM, Kushner R, Hunt T, Lubina JA, Patane
ity in patients with acute and weight-recovered J, Self B, Hunt P, McCamish M. Recombinant
anorexia nervosa. Prog Neuro-Psychopharm Biol leptin for weight loss in obese and lean adults:
Psychiat. 2009; 33: 658-662. A randomized, controlled, dose-escalation trial.
Eng R, Gold RM, Sawchenko PE. Hypothalamic JAMA. 1999; 282: 1568-1575.
hypoactivity prevented but not reversed by Jensen LB, Quaade F, Sorensen OH. Bone loss
subdiaphragmatic vagotomy. Physiol Behav. accompanying voluntary weight loss in obese
1978; 20: 637-641. humans. J Bone Miner Res. 1994; 9: 459-463.
Epstein LH, Wing RR. Aerobic exercise and weight. Johannsen DL, Knuth ND, Huizenga R, Rood
Addictive Behav. 1980; 5: 371-388. JC, Ravussin E, Hall KD. Metabolic slowing
Exner C, Hebebrand J, Remschmidt H, Wewetzer with massive weight loss despite preservation
C, Ziegler A, Herpertz S, Schweiger U, Blum of fat-free mass. J Clin Endocrinol Metab. 2012;
WF, Preibisch G, Heldmaier G, Klingenspor 97: 2489-2496.
M. Leptin suppresses semi-starvation induced Karsenty G, Oury F. The central regulation of bone
hyperactivity in rats: Implications for anorexia mass, the first link between bone remodeling and
nervosa. Mol Psychiatry. 2000; 5: 476-481. energy metabolism. J Clin Endocrinol Metab.
Fam BC, Morris MJ, Hansen MJ, Kebede M, 2010; 95: 4795-4801.
Andrikopoulos S, Proietto J, Thorburn AW. Kendall A, Levitsky DA, Strupp BJ, Lissner L.
Modulation of central leptin sensitivity and Weight loss on a low-fat diet: Consequence of
energy balance in a rat model of diet-induced the imprecision of the control of food intake
obesity. Diabetes Obes Metab. 2007; 9: 840-852. in humans. Am J Clin Nutr. 1991; 53: 1124-1129.
Farooqi IS, Jebb SA, Langmack G, Lawrence Kennedy GC. The hypothalamic control of food
E, Cheetham CH, Prentice AM, Hughes IA, intake in rats. Proc Roy Soc. 1950; 137: 535-549.
References 241

Kennedy GC. The role of depot fat in the hypotha- Rising R, Harper IT, Fontvielle AM, Ferraro RT,
lamic control of food intake in the rat. Proc R Spraul M, Ravussin E. Determinants of total
Soc B. 1953; 140: 578-592. daily energy expenditure: Variability in physi-
Kennedy GC. The development with age of hypo- cal activity. Am J Clin Nutr. 1994; 59: 800-804.
thalamic restraint upon the appetite of the rat. Routtenberg A, Kuznesof AW. Self-starvation
J Endocrinol. 1957; 16: 9-17. of rats living in activity wheels on a restricted
Kennedy GC, Mitra J. Body weight and food intake feeding schedule. J Comp Physiol Psych. 1967;
as initiating factors for puberty in the rat. J 64: 414-421.
Physiol. 1963a; 166: 408-418. Schulz LO, Schoeller DA. A compilation of total
Kennedy GC, Mitra J. Hypothalamic control of daily energy expenditures and body weights in
energy balance and the reproductive cycle in the healthy adults. Am J Clin Nutr. 1994; 60: 676-681.
rat. J Physiol. 1963b; 166: 396-407. Schwartz MW, Niswender KD. Adiposity signal-
Keys A, Brozek J, Henschel A, Mickelsen O, Taylor ing and biological defense against weight gain:
HL. The Biology of Human Starvation. Min- Absence of protection or central hormone
neapolis: University of Minnesota Press; 1950. resistance? J Clin Endocrinol Metab. 2004; 89:
King NA, Burley VJ, Blundell JE. Exercise-induced 5889-5897.
suppression of appetite: Effects on food intake Schwartz MW, Woods SC, Porte D Jr, Seeley RJ,
and implications for energy balance. Eur J Clin Baskin DG. Central nervous system control of
Nut. 1994; 48: 715-724. food intake. Nature. 2000; 404(6778): 661-671.
Levitsky DA. The non-regulation of food intake Sclafani A, Springer D. Dietary obesity in adult rats:
in humans: Hope for reversing the epidemic of Similarities to hypothalamic and human obesity
obesity. Physiol Behav. 2005; 86: 623-632. syndromes. Physiol Behav. 1976; 17: 461-471.
Levitsky DA, Halbmaier CA, Mrdjenovic G. The Seimon RV, Lange K, Little TJ, Brennan IM, Pilich-
freshman weight gain: A model for the study of iewicz AN, Feltrin KL, Smeets AJ, Horowitz M,
the epidemic of obesity. Int J Obes Relat Metab Feinle-Bisset C. Pooled-data analysis identifies
Disord. 2004; 28: 1435-1442. pyloric pressures and plasma cholecystokinin
Lissner L, Levitsky DA, Strupp BJ, Kalkwarf HJ, concentrations as major determinants of acute
Roe DA. Dietary fat and the regulation of energy energy intake in healthy, lean men. Am J Clin
intake in human subjects. Am J Clin Nutr. 1987; Nutr. 2010; 92: 61-68.
46: 886-892. Shepard TY, Weil KM, Sharp TA, Grunwald GK,
MacLean PS, Higgins JA, Wyatt HR, Melanson Bell ML, Hill JO, Eckel RH. Occasional physical
EL, Johnson GC, Jackson MR, Giles ED, Brown inactivity combined with a high-fat diet may be
IE, Hill JO. Regular exercise attenuates the important in the development and maintenance
metabolic drive to regain weight after long-term of obesity in human subjects. Am J Clin Nutr.
weight loss. Am J Physiol. 2009; 297: R793-R802. 2001; 73: 703-708.
McGuire MT, Wing RR, Klem ML, Deagle HM, Hill Speakman JR, Levitsky DA, Allison DB, Bray MS,
JO. Long-term maintenance of weight loss: Do de Castro JM, Clegg DJ, Clapham JC, Dulloo
people who lose weight through various weight loss AG, Grue L, Haw S, Hebebrand J, Hetherington
methods use different behaviors to maintain their MM, Higgs S, Jebb SA, Loos RJF, Luckman S,
weight? Int J Obesity. 1998; 22: 572-577. Luke A, Mohammed-Ali V, O’Rahilly S, Pereira
Neel JV. Diabetes mellitus a “thrifty” genotype M, Perusse L, Robinson TN, Rolls B, Symonds
rendered detrimental by “progress”? Am J Hum ME, Westerterp-Plantenga MS. Set points,
Genet. 1962; 14: 2323-2330. settling points and some alternative models:
O’Keefe JH, Vogel R, Lavie CJ, Cordain L. Organic Theoretical options to understand how genes
fitness: Physical activity consistent with our and environments combine to regulate body
hunter-gatherer heritage. Phys Sportsmed. 2010; adiposity. Dis Model Mech. 2011; 111: 1218-1224.
38: 11-18. Terjung RL, Baldwin KM, Winder WW, Holloszy
Pavlou KN, Krey S, Steffee WP. Exercise as an JO. Glycogen repletion in different types of
adjunct to weight loss and maintenance in muscle and liver after exhausting exercise. Am
moderately obese subjects. Am J Clin Nutr. 1989; J Physiol. 1974; 226: 1387-1391.
49: 1115-1123. Tsai AC, Rosenberg R, Borer KT. Metabolic
Pelleymounter MA, Cullen MJ, Baker MB, Hecht alterations induced by voluntary exercise and
R, Winters D, Boone T, Collins F. Effects of the discontinuation of exercise in hamsters. Am J
obese gene product on body weight regulation in Clin Nutr. 1982; 35: 943-949.
ob/ob mice. Science. 1995; 269: 540-543. Vanhecke TE, Frankin BA, Miller WM, deJong
Quarta D, Di Francesco C, Melotto S, Mangiarini AT, Coleman CJ, McCullough PA. Cardiore-
L, Heidbreder C, Hedou G. Systemic adminis- spiratory fitness and sedentary lifestyle in the
tration of ghrelin increases extracellular dopa- morbidly obese. Clin Cardiol. 2009; 32: 121-124.
mine in the shell but not the core subdivision of Vetter ML, Dumon KR, Williams NN. Surgical
the nucleus accumbens. Neurochem Int. 2009; treatments for obesity. Psychiatr Clin North Am.
54: 89-94. 2011; 34: 881-893.
242 References

Wing RR, Phelan S. Long-term weight loss main- A, Casaburi R. The effects of supraphysiologic
tenance. Am J Clin Nutr. 2005; 82(1 suppl): doses of testosterone on muscle size and strength
222S-225S. in normal men. New Engl J Med. 1996; 335: 1-7.
Zhang Y, Proenca R, Maffei M, Barone M, Leopold Bhasin S, Woodhouse L, Casaburi R, Singh AB,
L, Friedman JM. Positional cloning of the mouse Bhasin D, Berman N, Chen X, Yarasheski KE,
obese gene and its human homologue. Nature. Magliano L, Dzekov C, Dzekov J, Bross R, Phil-
1994; 372(6505): 425-432. lips J, Sinha-Hikim I, Shen R, Storer TW. Tes-
tosterone dose-response relationships in healthy
Chapter 7 young men. Am J Physiol. 2001; 281: E1172-E1181.
Adkisson EJ, Casey DP, Beck DT, Gurovich AN, Blouin K, Veilleux A, Luu-Thea V, Tchernof A.
Martin JS, Braith RW. Central, peripheral and Androgen metabolism in adipose tissue: Recent
resistance arterial reactivity: Fluctuates during advances. Mol Cell Endocr. 2009; 301: 97-103.
the phases of the menstrual cycle. Exp Biol Med Bonekat HW, Dombovy ML, Staats BA. Proges-
(Maywood). 2010; 235: 111-118. terone-induced changes in exercise performance
Ahima RS. Body fat, leptin, and hypothalamic and ventilatory response. Med Sci Sports Exerc.
amenorrhea. New Engl J Med. 2004; 351(9): 1987; 19: 118-123.
959-962. Borer KT. Physical activity in the prevention and
Ahmed ML, Ong KK, Dunger DB. Childhood obe- amelioration of osteoporosis in women: Inter-
sity and the timing of puberty. Trends Endocrinol action of mechanical, hormonal and dietary
Metab. 2009; 20: 237-242. factors. Sports Med. 2005; 35: 779-830.
Andreasen CH, Stender-Petersen KL, Mogensen Boyar RM, Katz J, Finkelstein JW, Kapen S,
MS, Torekov SS, Wegner L, Andersen G, Nielsen Weiner H, Weitzman ED, Hellman L. Anorexia
AL, Albrechtsen A, Borch-Johnsen K, Rasmus- nervosa: Immaturity of the 24-hour luteinizing
sen SS, Clausen JO, Sandbaek A, Lauritzen T, hormone secretory pattern. New Engl J Med.
Hansen L, Jørgensen T, Pedersen O, Hansen T. 1974; 291: 861-865.
Low physical activity accentuates the effect of Brook CG, Hindmarsh PC. The somatotropic axis
the FTO rs9939609 polymorphism on body fat in puberty. Endocrinol Metab Clin North Am.
accumulation. Diabetes. 2008; 57: 95-101. 1992; 21: 767-782.
Arends JC, Cheung MY, Barrack MT, Nattiv A. Burrows M, Bird SR. Velocity at VO2max and peak
Restoration of menses with nonpharmacologic treadmill velocity are not influenced within or
therapy in collegiate athletes with menstrual across the phases of menstrual cycle. Eur J Appl
disturbances: A 5 year retrospective study. Int J Physiol. 2005; 93: 575-580.
Sport Nutr Exerc Metab. 2012; 22: 98-108. Bushman B, Masterson G, Nelson J. Anaerobic
Arner P. Effects of testosterone on fat cell lipolysis. power performance and the menstrual cycle:
Species differences and possible role in polycys- Eumenorrheic and oral contraceptive users. J
tic ovarian syndrome. Biochimie. 2005; 87: 39-43. Sports Med Phys Fitness. 2006; 46: 132-137.
Bamman MM, Shipp JR, Jiang J, Gower BA, Cameron N, Tanner JM, Whitehouse RH. A
Hunter GR, Goodman A, McLafferty CL Jr, longitudinal analysis of the growth of limb seg-
Urban RJ. Mechanical load increases muscle ments in adolescence. Ann Human Biol. 1982;
IGF-I and androgen receptor mRNA concen- 9: 211-220.
trations in humans. Am J Physiol. 2001; 280: Campbell SE, Angus DJ, Febbraio MA. Glucose
E383-E390. kinetics and exercise performance during phases
Beitins IZ, McArthur JW, Turnbull BA, Skrinar of the menstrual cycle: Effect of glucose inges-
GS, Bullen BA. Exercise induces two types of tion. Am J Physiol. 2001; 281: E817-E825.
human luteal dysfunction: Confirmation by Carter A, Dobridge J, Hackney AC. Influence of
urinary free progesterone. J Clin Endocr Metab. estrogen on markers of muscle tissue damage
1991; 72: 1350-1358. following eccentric exercise. Fiziol Cheloveka.
Belanger MJ, Moore DC, Crisco JJ III, Fadale PD, 2001; 27: 133-137.
Hulstyn MJ, Ehrlich MG. Knee laxity does not Cary GA, La Spada AR. Androgen receptor func-
vary with the menstrual cycle, before or after tion in motor neuron survival and degeneration.
exercise. Am J Sports Med. 2004; 32: 1150-1157. Phys Med Rehab Clinics North Am. 2008; 19:
Bemben DA, Salm PC, Salm AJ. Ventilatory and 479-494.
blood lactate responses to maximal treadmill Casazza GA, Jacobs KA, Suh SH, Miller BF, Horn-
exercise during the menstrual cycle. J Sports ing MA, Brooks GA. Menstrual cycle phase
Med Phys Fitness. 1995; 35: 257-262. and oral contraceptive effects on triglyceride
Berga SL, Mortola JF, Yen SSC. Amplification of mobilization during exercise. J Appl Physiol.
nocturnal melatonin secretion in women with 2004; 97: 302-309.
functional hypothalamic amenorrhea. J Clin Casazza GA, Suh SH, Miller BF, Navazio FM,
Endocr Metab. 1988; 66: 242-244. Brooks GA. Effects of oral contraceptives on
Bhasin S, Storer TW, Berman N, Callegari C, Clev- peak exercise capacity. J Appl Physiol. 2002;
enger B, Phillips J, Bunnell TJ, Tricker R, Shirazi 93: 1698-1702.
References 243

Chagin AS, Sävenhahl L. Genes of importance in Dombovy ML, Bonekat HW, Williams TJ, Staats
the hormonal regulation of growth plate carti- BA. Exercise performance and ventilatory
lage. Horm Res. 2009; 71(suppl 2): 41-47. response in the menstrual cycle. Med Sci Sports
Chen J-Q, Brown TR, Russo J. Regulation of Exerc. 1987; 19: 111-117.
energy metabolism pathways by estrogens and Donoso MA, Muñoz-Calvo MT, Barrios V, Gar-
estrogenic chemicals and potential implications rido G, Hawkins F, Argente J. Increased circu-
in obesity associated with increased exposure lating adiponectin levels and decreased leptin/
to endocrine disruptors. Biochim Biophys Acta. soluble leptin receptor ratio throughout puberty
2009; 1793: 1128-1143. in female ballet dancers: Association with body
Christakos S, Dhawan P, Porta A, Mady LJ, Seth composition and the delay in puberty. Eur J
T. Vitamin D and intestinal calcium absorption. Endocrinol. 2010; 162: 905-911.
Mol Cell Endocrinol. 2011; 347: 25-29. Ducher G, Turner AI, Kukuljan S, Pantano
Clapp JF. Maternal carbohydrate intake and preg- KJ, Carlson JL, Williams NI, De Souza MJ.
nancy outcome. Proc Nutr Soc. 2002; 61: 45-50. Obstacles in the optimization of bone health
Clapp JF. Inf luence of endurance exercise and outcomes in the female athlete triad. Sports Med.
diet on human placental development and fetal 2011; 41: 587-607.
growth. Placenta. 2006; 27: 527-534. Ellingboe J, Veldhuis JD, Mendelson JH. Effect of
Confavreux CB, Levine RL, Karsenty G. A para- endogenous opioid blockade on the amplitude
digm of integrative physiology, the crosstalk and frequency of pulsatile luteinizing hormone
between bone and energy metabolisms. Mol Cell secretion in man. J Clin Endocr Metab. 1982;
Endocrinol. 2009; 310: 21-29. 54: 854-857.
Cook CJ, Crewther BT. Changes in salivary testos- Elliott KJ, Cable NT, Reilly T, Diver MJ. Effect of
terone concentrations and subsequent voluntary menstrual cycle phase on the concentration of
squat performance following the presentation of bioavailable 17-beta oestradiol and testosterone
short video clips. Horm Behav. 2012; 61: 17-22. and muscle strength. Clin Sci (Lond). 2003; 105:
Crewther BT, Cook C, Cardinale M, Weatherby 663-669.
RP, Lowe T. Two emerging concepts for elite Ferrando AA, Tipton KD, Doyle D, Phillips SM,
athletes: The short-term effects of testosterone Cortiella J, Wolfe RR. Testosterone injection stim-
and cortisol on the neuromuscular system and ulates net protein synthesis but not tissue amino
the dose-response training role of these endog- acid transport. Am J Physiol. 1998; 275: 864-871.
enous hormones. Sports Med. 2011; 41: 103-123. Fridén C, Hirschberg AL, Saartok T. Muscle
Cum m ing DC, Br unsting LA III, Strich G, strength and endurance do not significantly
Ries AL, Rebar RW. Reproductive hormone vary across 3 phases of the menstrual cycle in
increases in response to acute exercise in men. moderately active premenopausal women. Clin
Med Sci Sports Exerc. 1986; 18: 369-373. J Sport Med. 2003; 13: 238-241.
Cumming DC, Wall SR, Galbraith MA, Belcas- Friedman B, Kinderman W. Energy metabolism
tro AN. Reproductive hormone responses to and regulatory hormones in women and men
resistance exercise. Med Sci Sports Exerc. 1987; during endurance exercise. Eur J Appl Physiol
19: 234-238. Occup Physiol. 1989; 59: 1-9.
Cunningham MJ, Clifton DK, Steiner RA. Leptin’s Frisch RE. Weight at menarche: Similarity for well-
actions on the reproductive axis: Perspectives nourished and undernourished girls at differing
and mechanisms. Biol Rep. 1999; 60: 216-222. ages, and evidence for historical constancy.
Deschenes MR, Maresh CM, Armstrong LE, Pediatrics. 1972; 50: 445-450.
Covault J, Kraemer WJ, Crivello JF. Endurance Frisch RE, McArthur JW. Menstrual cycles: Fat-
and resistance exercise induce muscle fiber type ness as a determinant of minimum weight for
specific responses in androgen binding capacity. height necessary for their maintenance or onset.
J Steroid Biochem Mol Biol. 1994; 50: 175-179. Science. 1974; 185: 949-951.
De Souza MJ, Luciano AA, Arce JC, Demers Frisch RE, Revelle R. The height and weight of
LM, Loucks AB. Clinical tests explain blunted girls and boys at the time of initiation of the
cortisol responsiveness but not mild hypercorti- adolescent growth spurt in height and weight
solemia in amenorrheic runners. J Appl Physiol. and the relationship to menarche. Human Biol.
1994; 76: 1302-1309. 1971; 43: 140-159.
De Souza MJ, Toombs RJ, Scheid JL, O’Donnell E, Froberg K, Pedersen PK. Sex differences in
West SL, Williams NI. High prevalence of subtle endurance capacity and metabolic response to
and severe menstrual disturbances in exercising prolonged heavy exercise. Eur J Appl Physiol.
women: Confirmation using daily hormone mea- 1984; 52: 446-450.
sures. Hum Reprod. 2010; 25: 491-503. Fu MH, Maher AC, Hamadeh MJ, Ye C, Tarnop-
Dessypris K, Kuoppasalmi K, Alderkreutz H. olsky MA. Exercise, sex, menstrual cycle phase,
Plasma cortisol, testosterone, androstenedione and 17beta-estradiol influence on metabolism-
and luteinizing hormone (LH) in a non-competi- related genes in human skeletal muscle. Physiol
tive marathon run. J Ster Biochem. 1976; 7: 33-37. Genom. 2009; 40: 34-47.
244 References

Gabory A, Attig L, Junien C. Sexual dimorphism cycle phase on skeletal muscle contractile
in environmental epigenetic programming. Mol characteristics in humans. J Physiol. 2001; 530:
Cell Endocrinol. 2009; 304: 8-18. 161-166.
Galbo H, Hummer L, Petersen IB, Christensen Jensen MD, Cryer PE, Johnson CM, Murray MJ.
NJ, Bie N. Thyroid and testicular hormone Effects of epinephrine on regional free fatty acid
responses to graded and prolonged exercise in and energy metabolism in men and women. Am
man. Eur J Appl Physiol. 1977; 36: 101-106. J Physiol. 1996; 270: E259-E264.
Garcia AM, Lacerda MG, Fonseca IA, Reis FM, Jezova D, Vigas M. Testosterone response to
Rodrigues LO, Silami-Garcia E. Luteal phase exercise during blockade and stimulation of
of the menstrual cycle increases sweating rate adrenergic receptors in man. Hormone Res.
during exercise. Braz J Med Biol Res. 2006; 39: 1981; 15: 141-147.
1255-1261. Jezova D, Vigas M, Tatar P, Kvetnansky R, Nazar
Gill JM, Malkova D, Hardman AE. Reproduc- K, Kaciuba-Uscilko H, Kozlowski S. Plasma tes-
ibility of an oral fat tolerance test is influenced tosterone and catecholamine responses to physi-
by phase of menstrual cycle. Horm Metab Res. cal exercise of different intensities in men. Eur
2005; 37: 336-341. J App Physiol Occupat Physiol. 1985; 54: 62-66.
Glickman-Weiss EL, Cheatham CC, Caine N, Kadi F. Cellular and molecular mechanisms
Blegen M, Marcinkiewicz J. Influence of gender responsible for the action of testosterone on
and menstrual cycle on a cold air tolerance test human skeletal muscle. A basis for illegal per-
and its relationship to thermosensitivity. Under- formance enhancement. Br J Pharmacol. 2008;
sea Hyperb Med. 2000; 27: 75-81. 154: 522-528.
Hackney AC, McCracken-Compton MA, Ain- Khosla S. Minireview: the OPG/RANKL/RANK
sworth B. Substrate responses to submaximal system. Endocrinology 2001, 142: 5050-5055.
exercise in the midfollicular and midluteal Kolka MA, Stephenson LA. Control of sweating
phases of the menstrual cycle. Int J Sport Nutr. during the human menstrual cycle. Eur J Appl
1994; 4: 299-308. Physiol Occup Physiol. 1989; 58: 890-895.
Hardman AE. The influence of exercise on post- Kolka MA, Stephenson LA. Effect of luteal phase
prandial triacylglycerol metabolism. Atheroscle- elevation in core temperature on forearm blood
rosis. 1998; 141(suppl 1): S93-S100. flow during exercise. J Appl Physiol. 1997; 82:
Hellstrom L, Blaak E, Hagstrom-Toft E. Gender 1079-1083.
differences in adrenergic regulation of lipid Kraemer WJ, Heleniak RJ, Tryniecki JL, Kraemer
mobilization during exercise. Int J Sports Med. GR, Okazaki NJ, Castracane VD. Follicular and
1996; 17: 439-447. luteal phase hormonal responses to low-volume
Hessemer V, Brück K. Influence of menstrual cycle resistive exercise. Med Sci Sports Exerc. 1995;
on thermoregulatory, metabolic, and heart rate 27: 809-817.
responses to exercise at night. J Appl Physiol. Kraemer WJ, Ratamess NA. Hormonal responses
1985; 59: 1911-1917. and adaptations to resistance exercise and train-
Hirata K, Nagasaka T, Hirai A, Hirashita M, ing. Sports Med. 2005; 35: 339-361.
Takahata T, Nunomura T. Effects of human Kuoppasalmi K, Naveri H, Rehunen S, Harkonen
menstrual cycle on thermoregulatory vasodila- M, Adlercreutz H. Effects of strenuous anaero-
tion during exercise. Eur J Appl Physiol Occup bic running exercise on plasma growth hormone,
Physiol. 1986; 54: 559-565. cortisol, luteinizing hormone, testosterone,
Horton TJ, Pagliassotti MJ, Hobbs K, Hill JO. Fuel androstenedione, estrone, and estradiol. J Ster
metabolism in men and women during and after Biochem. 1976; 7: 823-829.
long-duration exercise. J Appl Physiol. 1998; 85: Lawrence JE, Klein JC, Carter JR. Menstrual cycle
1823-1832. elicits divergent forearm vascular responses to
Imai Y, Nakamura T, Matsumoto T, Takaoka K, vestibular activation in humans. Auton Neurosci.
Kato S. Molecular mechanisms underlying the 2010; 154: 89-93.
effects of sex steroids on bone and mineral metab- Lewis-Wambi JSZ, Jordan VC. Estrogen regulation
olism. J Bone Miner Metab. 2009; 27: 127-130. of apoptosis: How can one hormone stimulate
Isacco L, Duché P, Boisseau N. Influence of hor- and inhibit? Breast Cancer Res. 2009; 11: 206.
monal status on substrate utilization at rest and Limberg JK, Eldridge MW, Proctor LT, Sebranek
during exercise in the female population. Sports JJ, Schrage WG. Alpha-adrenergic control of
Med. 2012; 42: 327-342. blood flow during exercise: Effect of sex and
James P, Rivier C, Lee S. Presence of corticotro- menstrual phase. J Appl Physiol. 2010; 109:
phin-releasing factor and/or tyrosine hydroxy- 1360-1368.
lase in cells of a neural brain-testicular pathway Loucks AB, Mortola JF, Girton L, Yen SSC.
that are labelled by a transganglionic tracer. J Alterations in the hypothalamic-pituitary-
Neuroendocrinol. 2008; 20: 173-181. ovarian and hypothalamic-pituitary-adrenal
Janse de Jonge XA, Boot CR, Thom JM, Ruell PA, axes in athletic women. J Clin Endocr Metab.
Thompson MW. The influence of menstrual 1989; 68: 402-411.
References 245

Loucks AB, Thuma JR. Luteinizing hormone cise of moderate intensity in untrained subjects.
pulsatility is disrupted at a threshold of energy Am J Physiol. 2002; 283: E58-E65.
availability in regularly menstruating women. J Nakayama T, Suzuki M, Ishizuka N. Action of pro-
Clin Endocr Metab. 2003; 88: 297-311. gesterone on preoptic thermosensitive neurones.
Loucks AB, Verdun M, Heath EM. Low energy Nature. 1975; 258: 80.
availability, not stress of exercise, alters LH Ogilvie K, Hales K, Roberts M, Hales D, Rivier C.
pulsatility in exercising women. J Appl Physiol. The inhibitory effect of intracerebroventricularly
1998; 84: 37-46. injected interleukin 1beta on testosterone secre-
Lussana F, Painter RC, Ocke MC, Buller HR, tion in the rat: Role of steroidogenic acute regu-
Bossuyt PM, Roseboom TJ. Prenatal exposure latory protein. Biol Reprod. 1999; 60: 527-533.
to the Dutch famine is associated with a prefer- O’Loughlin PD, Morris HA. Estrogen deficiency
ence for fatty foods and a more atherogenic lipid impairs intestinal calcium absorption in the rat.
profile. Am J Clin Nutr. 2008; 88: 1648-1652. J Physiol. 1998; 511: 313-322.
Lynn BM, McCord JL, Halliwill JR. Effects of the Olson D, Sikka RS, Hayman J, Novak M, Stavig
menstrual cycle and sex on postexercise hemody- C. Exercise in pregnancy. Curr Sports Med Rep.
namics. Am J Physiol. 2007; 292: R1260-R1270. 2009; 8: 147-153.
MacConnie SE, Barkan A, Lampman RM, Schork Oosthuyse T, Bosch AN. The effect of the menstrual
MA, Beitins IZ. Decreased hypothalamic cycle on exercise metabolism: Implications for
gonadotropin-releasing hormone secretion in exercise performance in eumenorrhoeic women.
male marathon runners. New Engl J Med. 1986; Sports Med. 2010; 40: 207-227.
315: 411-417. Oscai LB, Babirak SP, Dubach FB, McGarr JA,
Macnutt MJ, De Souza MJ, Tomczak SE, Homer Spirakis CN. Exercise and food restriction:
JL, Sheel AW. Resting and exercise ventilatory Effect on adipose tissue cellularity. Am J Physiol.
chemosensitivity across the menstrual cycle. J 1974; 227: 901-904.
Appl Physiol. 2012; 112: 737-747. Oscai LB, Spirakis CN, Wolff CA, Beck RJ. Effects
Malina RM, Harper AB, Campbell DE. Age at of exercise and of food restriction on adipose
menarche in athletes and non-athletes. Med Sci tissue cellularity. J Lipid Res. 1972; 13: 588-592.
Sports. 1973; 5: 11-13. Park SK, Stefanyshyn DJ, Ramage B, Hart DA,
Marshall WA, Tanner JM. Variations in pattern Ronsky JL. Relationship between knee joint
of pubertal changes in girls. Arch Dis Children. laxity and knee joint mechanics during the men-
1969; 44: 291-303. strual cycle. Br J Sports Med. 2009; 43: 174-179.
Marshall WA, Tanner JM. Variations in pattern Pivarnik JM, Marichal CJ, Spillman T, Morrow JR
of pubertal changes in boys. Arch Dis Children. Jr. Menstrual cycle phase affects temperature
1970; 45: 13-23. regulation during endurance exercise. J Appl
Marshall WA, Tanner JM. Puberty. In: Falkner F, Physiol. 1992; 72: 543-548.
Tanner JM, eds. Human Growth: A Comprehen- Rencken ML, Chesnut CH III, Drinkwater BL.
sive Treatise. New York: Plenum Press; 1986: 171. Bone density at multiple skeletal sites in amenor-
Martos-Moreno GA, Chowen JA, Argente J. rheic athletes. JAMA. 1996; 276: 238-240.
Metabolic signals in human puberty: Effects of Rivier C, Rivest S. Effect of stress on the activity
over and undernutrition. Mol Cell Endocrinol. of the hypothalamic-pituitary-gonadal axis:
2010; 324: 70-81. Peripheral and central mechanisms. Biol Rep.
Mauras N, Rogol AD, Haymond MW, Veldhuis 1991; 45: 523-532.
JD. Sex steroids, growth hormone, insulin-like Safarinejad MR, Azma K, Kolahi AA. The effects of
growth factor-1: Neuroendocrine and metabolic intensive, long-term treadmill running on repro-
regulation of puberty. Horm Res. 1996; 45: 74-80. ductive hormones, hypothalamus-pituitary-testis
McCaffery JM, Papandonatos GM, Lyons MJ, Wing axis, and semen quality: A randomized controlled
RR. Gene x environment interaction of vigorous study. J Endocrinol. 2009; 200: 259-271.
exercise and body mass index among Vietnam-era Seeman E. Sexual dimorphism in skeletal size,
twins. Am J Clin Nutr. 2009; 89: 1011-1018. density, and strength. J Clin Endocrinol Metab.
Meyer MR, Haas E, Prossnitz ER, Barton M. 2001; 86: 4576-4584.
Non-genomic regulation of vascular cell func- Selvage DJ, Lee SY, Parsons LH, Seo DO, Rivier
tion and growth by estrogen. Mol Cell Endocrin. CL. A hypothalamic-testicular neural pathway
2009; 308: 9-16. is influenced by brain catecholamines, but not
Miller BF, Hansen M, Olesen JL, Flyvbjerg A, testicular blood flow. Endocrinology. 2004; 145:
Schwarz P, Babraj JA, Smith K, Rennie MJ, 1750-1759.
Kjaer M. No effect of menstrual cycle on myo- Selvage DJ, Rivier C. Importance of the para-
fibrillar and connective tissue protein synthesis ventricular nucleus of the hypothalamus as a
in contracting skeletal muscle. Am J Physiol. component of a neural pathway between the
2006; 290: E163-E168. brain and the testes that modulates testosterone
Mittendorfer B, Horowitz JF, Klein S. Effect of secretion independently of the pituitary. Endo-
gender on lipid kinetics during endurance exer- crinology. 2003; 144: 594-598.
246 References

Shultz SJ, Sander TC, Kirk SE, Perrin DH. Sex Tremblay MS, Copeland JL, Van Helder W. Influ-
differences in knee joint laxity change across ence of exercise duration on post-exercise steroid
the female menstrual cycle. J Sports Med Phys hormone responses in trained males. Eur J Appl
Fitness. 2005; 45: 594-603. Physiol. 2005; 94: 505-513.
Skaletsky H, Kuroda-Kawaguchi T, Minx PJ, Vaiksaar S, Jürimäe J, Mäestu J, Purge P, Kalytka
Cordum HS, Hillier L, Brown LG, Repping S, Shakhlina L, Jürimäe T. No effect of men-
S, Pyntikova T, Ali J, Bieri T, Chinwalla A, strual cycle phase and oral contraceptive use
Delehaunty A, Delehaunty K, Du H, Fewell on endurance performance in rowers. J Strength
G, Fulton L, Fulton R, Graves T, Hou S-H, Cond Res. 2011; 25: 1571-1578.
Latrielle P, Leonard S, Mardis E, Maupin R, Vandenput L, Ohlsson C. Estrogens as regulators
McPherson J, Miner T, Nash W, Nguyen C, of bone health in men. Nature Rev Endocrin.
Ozersky P, Pepin K, Rock S, Rohlfing T, Scott 2009; 5: 437-443.
K, Schultz B, Strong C, Tin-Wollam A, Yang S-P, Van Lunen BL, Roberts J, Branch JD, Dowling
Waterston RH, Wilson RK, Rozen S, Page DC. EA. Association of menstrual-cycle hormone
The male-specific region of human Y chromo- changes with anterior cruciate ligament laxity
some is a mosaic of discrete sequence classes. measurements. J Athl Train. 2003; 38: 298-303.
Nature. 2003; 423: 825-837. von Schnurbein J, Moss A, Nagel SA, Muehleder H,
Smekal G, von Duvillard SP, Frigo P, Tegelhofer Debatin KM, Farooqi IS, Wabitsch M. Leptin
T, Pokan R, Hofmann P, Tschan H, Baron R, substitution results in the induction of menstrual
Wonisch M, Renezeder K, Bachl N. Menstrual cycles in an adolescent with leptin deficiency and
cycle: No effect on exercise cardiorespiratory hypogonadotropic hypogonadism. Horm Res
variables or blood lactate concentration. Med Paediatr. 2012; 77(2): 127-133.
Sci Sports Exerc. 2007; 39: 1098-1106. Warren MP. The effect of exercise on pubertal
Staron RS, Hagerman FC, Hikida RS, Murray progression and reproductive function in girls.
TF, Hostler DP, Crill MT, Ragg KE, Toma K. J Clin Endocr Metab. 1980; 51: 1150-1157.
Fiber type composition of the vastus lateralis Waxman DJ, O’Connor C. Growth hormone regu-
muscle of young men and women. J Histochem lation of sex-dependent liver gene expression.
Cytochem. 2000; 48: 623-629. Mol Endocrinol. 2006; 20: 2613-2629.
Sunderland C, Tunaley V, Horner F, Harmer D, Weiss LW, Cureton KJ, Thompson FN. Compari-
Stokes KA. Menstrual and oral contraceptives’ son of serum testosterone and androstenedione
effect on growth hormone response to sprinting. responses to weight lifting in men and women.
Appl Physiol Nutr Metab. 2011; 36: 495-502. Eur J App Physiol Occupat Physiol. 1983; 50:
Tanner JM. Sequence and tempo in the somatic 413-419.
changes in puberty. In: Grumbach MM, Grave Wenz M, Berend JZ, Lynch NA, Chappell S, Hack-
GD, Mayer FE, eds. Control of the Onset of ney AC. Substrate oxidation at rest and during
Puberty. Philadelphia: Lippincott Williams & exercise: Effects of menstrual cycle phase and
Wilkins; 1974. diet composition. J Physiol Pharmacol. 1997;
Tanner JM. A History of the Study of Human 48: 851-860.
Growth. Cambridge: Cambridge University White PC. Genetics of steroid 21-hydroxylase defi-
Press; 1981. ciency. Recent Prog Horm Res. 1987; 43: 305-336.
Tarnopolsky LJ, MacDougal JD, Atkinson SA, Williams NI, Berga SL, Cameron JL. Synergism
Tarnopolsky MA, Sutton JR. Gender differ- between psychosocial and metabolic stressors:
ences in substrate for endurance exercise. J Appl Impact on reproductive function in cynomolgus
Physiol. 1990; 68: 302-308. monkeys. Am J Physiol Endocrinol Metab. 2007;
Tenaglia SA, McLellan TM, Klentrou PP. Influence 293: E270-E276.
of menstrual cycle and oral contraceptives on Williams NI, McArthur JW, Turnbull BA, Bullen
tolerance to uncompensable heat stress. Eur BA, Skrinar GS, Beitins IZ, Besser GM, Rees
J Appl Physiol Occup Physiol. 1999; 80: 76-83. LH, Gilbert I, Cramer D, Perry L, Tunstal Pedoe
Thomas DM, Clapp JF, Shernce S. A foetal energy DS. Effects of follicular phase exercise on lutein-
balance equation based on maternal exercise izing hormone pulse characteristics in sedentary
and diet. J Roy Soc Interface. 2008; 5: 449-455. eumenorrhoeic women. Clin Endocrinol. 1994;
Thompson HS, Hyatt JP, De Souza MJ, Clark- 41: 787-794.
son PM. The effects of oral contraceptives on Zachman M. Interrelations between growth
delayed onset muscle soreness following exercise. hormone and sex hormones: Physiological and
Contraception. 1997; 56: 59-65. therapeutic consequences. Horm Res. 1992;
Ting HJ, Chang C. Actin associated proteins 38(suppl 1): 1-8.
function as androgen receptor coregulators: Zderic TW, Coggan AR, Ruby BC. Glucose kinet-
An implication of androgen receptor’s roles in ics and substrate oxidation during exercise in
skeletal muscle. J Steroid Biochem Mol Biol. the follicular and luteal phases. J Appl Physiol.
2008; 111: 157-163. 2001; 90: 447-453.
References 247

Chapter 8 Brahm H, Piehl-Aulin K, Saltin B, Ljunghall S. Net


fluxes over working thigh of hormones, growth
Adams GR. Invited review: Autocrine/paracrine
factors and biomarkers of bone metabolism
IGF-I and skeletal muscle adaptation. J Appl
during short lasting dynamic exercise. Calcif
Physiol. 2002; 93: 1159-1167.
Tissue Int. 1997; 60: 175-180.
Adamson JW, Vapnek D. Recombinant erythro- Canalis E, Delany AM. Mechanisms of glucocor-
poietin to improve athletic performance. New ticoid action in bone. Ann NY Acad Sci. 2002;
Engl J Med. 1991; 324: 698-699. 966: 73-81.
Adi S, Bin-Abbas B, Wu NY, Rosenthal SM. Early Caralis DG, Edwards L, Davis PJ. Serum total
stimulation and late inhibition of extracellular and free thyroxine and triiodothyronine during
signal-regulated kinase 1/2 phosphorylation by dynamic muscular exercise in man. Am J Physiol.
IGF-I: A potential mechanism mediating the 1977; 233: E115-E118.
switch in IGF-I action on skeletal muscle cell dif- Chalmers RJ, Bloom SR, Duncan G, Johnson RH,
ferentiation. Endocrinology. 2002; 143: 511-516. Sulaiman WR. The effect of somatostatin on
Ahima RS, Prabakaran D, Mantzoros C, Qu D, metabolic and hormonal changes during and after
Lowell B, Maratos-Flier E, Flier JS. Role of exercise. Clin Endocrinol (Oxf). 1979; 10: 451-458.
leptin in the neuroendocrine response to fasting. Clemmons DR. Role of IGF-I in skeletal muscle
Nature. 1999; 382: 250-252. mass maintenance. Trends Endocrinol Metab.
Atherton PJ, Babraj J, Smith K, Singh J, Rennie MJ, 2009; 20: 349-356.
Wackerhage H. Selective activation of AMPK- Considine RV. Human leptin: An adipocyte hor-
PGC-1alpha or PKB-TSC2-mTOR signaling can mone with weight-regulatory and endocrine
explain specific adaptive responses to endurance functions. Sem Vasc Med. 2005; 5: 15-24.
or resistance training-like electrical muscle Coolican SA, Samuel DS, Ewton DZ, McWade
stimulation. FASEB J. 2005; 19: 786-788. FJ, Florini JR. The mitogenic and myogenic
Bassel-Duby R, Olson EN. Signaling pathways in actions of insulin-like growth factors utilize
skeletal muscle remodeling. Ann Review Bio- distinct signaling pathways. J Biol Chem. 1997;
chem. 2006; 75: 19-37. 272: 6653-6662.
Baylor LS, Hackney AC. Resting thyroid and leptin Cuneo RC, Salomon F, Wiles CM, Hesp R, Sönk-
hormone changes in women following intense, sen PH. Growth hormone treatment in growth
prolonged exercise training. Eur J Appl Physiol. hormone-deficient adults. I. Effects on muscle
2003; 88: 480-484. mass and strength. J Appl Physiol. 1991; 70: 688.
Bhasin S, Storer TW, Berman N, Callegari C, Clev- Deuster PA, Chrousos GP, Luger A, DeBolt JE,
enger B, Phillips J, Bunnell TJ, Tricker R, Shirazi Bernier LL, Trostmann UH, Kyle SB, Montgom-
A, Casaburi R. The effects of supraphysiologic ery LC, Loriaux DL. Hormonal and metabolic
doses of testosterone on muscle size and strength responses of untrained, moderately trained, and
in normal men. New Engl J Med. 1996; 335: 1-7. highly trained men to three exercise intensities.
Bhasin S, Taylor WE, Singh R, Artaza J, Sinha- Metab Clin Exp. 1989; 38: 141-148.
Hakim I, Jasuja R, Choi H, Gonzalez-Cadavid DeVol DL, Rothwein P, Sadow JL, Nowakowski
NF. The mechanism of androgen effect on body J, Bechtel PJ. Activation of insulin-like growth
composition: Mesenchymal pluripotent cell as factor gene expression during work-induced
the target of androgen action. J Gerontol A Biol skeletal muscle growth. Am J Physiol. 1990;
Sci Med Sci. 2003; 58: M1103-M1110. 259: E89-E95.
Bigard AX, Koulmann N, Bahi L, Sanchez H, Deyssig R, Frisch H, Blum WF, Waldhör T. Effect of
Ventura-Clapier R. Thyroid hormones and growth hormone treatment on hormonal param-
muscle phenotype: Involvement of new signaling eters, body composition and strength in athletes.
pathways. J Soc Biol. 2008; 202: 93-100. Acta Endocrinol (Copenh). 1993; 128: 313-318.
Birzniece V, Nelson AE, Ho KK. Growth hormone Doessing S, Kjaer M. Growth hormone and con-
and physical performance. Trends Endocrinol nective tissue in exercise. Scand J Med Sci Sports.
Metab. 2011; 22: 171-178. 2005; 15: 202-210.
Borer KT, Wuorinen E, Ku K, Burant C. Appetite Dorn GW II, Force T. Protein cascades in the
responds to changes in meal content, whereas ghre- regulation of cardiac hypertrophy. J Clin Invest.
lin, leptin, and insulin track changes in energy avail- 2005; 115: 527-537.
ability. J Clin Endocr Metab. 2009; 94: 2290-2298. Eliakim A, Brasel JA, Mohan S, Barstow TJ,
Boström P, Wu J, Jedrychowski MP, Korde A, Ye Berman N, Cooper DM. Physical fitness, endur-
L, Lo JC, Rasbach KA, Boström EA, Choi JH, ance training, and the growth hormone-insulin-
Long JZ, Kajimura S, Zingaretti MC, Vind BF, like growth factor I system in adolescent females.
Tu H, Cinti S, Højlund K, Gygi SP, Spiegelman J Clin Endocrinol Metab. 1996; 81: 3986-3992.
BM. A PGC1-α-dependent myokine that drives Force T, Bonventre JV. Growth factors and mito-
brown-fat-like development of white fat and gen-activated protein kinases. Hypertension.
thermogenesis. Nature. 2012; 481(7382): 463-468. 1998; 31: 152-161.
248 References

Frank SJ, Fuchs SY. Modulation of growth hor- Herrington J, Carter-Su C. Signaling pathways
mone receptor abundance and function: Roles activated by the growth hormone receptor.
for the ubiquitin-proteasome system. Biochim TRENDS Endocr Metab. 2001; 12: 252-257.
Biophys Acta. 2008; 1782: 785-794. Hickey MS, Considine RV, Israel RG, Mahar TL,
Fryburg DA, Gelfand RA, Barrett EJ. Growth McCammon MR, Tyndall GL, Houmard JA,
hormone stimulates skeletal muscle protein Caro JF. Leptin is related to body fat content
synthesis and antagonizes insulin’s antiproteo- in male distance runners. Am J Physiol. 1996;
lytic action in humans. Am J Physiol. 1991; 260: 271: E938-E940.
E499-E504. Hickson RC. Interference of strength development
Giannoulis MG, Martin FC, Nair KS, Umpleby by simultaneously training for strength and
AM, Sonksen P. Hormone replacement therapy endurance. Eur J Appl Physiol Occup Physiol.
and physical function in healthy older men. Time 1980; 45: 255-263.
to talk hormones? Endocr Rev. 2012; 33: 314-377. Hietakangas V, Cohen SM. Regulation of tissue
Gibbons JJ, Abraham RT, Yu K. Mammalian growth through nutrient sensing. Ann Rev Genet.
target of rapamycin: Discovery of rapamycin 2009; 43: 389-410.
reveals a signaling pathway important for Hill M, Goldspink G. Expression and splicing of
normal and cancer cell growth. Semin Oncol. the insulin-like growth factor gene in rodent
2009; 36(suppl 3): S3-S17. muscle is associated with muscle satellite (stem)
Gibney J, Healy ML, Sonksen PH. The growth hor- cell activation following local tissue damage. J
mone/insulin-like growth factor axis in exercise Physiol. 2003; 549: 409-418.
and sport. Endocr Rev. 2007; 28: 603-624. Ho PJ, Friberg RD, Barkan AL. Regulation of
Goldberg AI. Work-induced growth of skeletal pulsatile growth hormone secretion by fasting in
muscle in normal and hypophysecomized rats. normal subjects and patients with acromegaly. J
Am J Physiol. 1967; 213: 1193-1198. Clin Endocrinol Metab. 1992; 75: 812-819.
Goldspink G. Research on mechano growth factor: Hoffman DM, Crampton L, Sernia C, Ho KKY.
Its potential for optimising physical training as Short term growth hormone (GH) treatment of
well as misuse in doping. Br J Sports Med. 2005; GH deficient adults increases body sodium and
39: 787-788. extracellular water but not blood pressure. J Clin
Gupta MP. Factors controlling cardiac myosin- Endocrinol Metab. 1996; 81: 1123-1128.
isoform shift during hypertrophy and heart Jenkins PJ. Growth hormone and exercise. Clin
failure. J Mol Cell Cardiol. 2007; 43: 388-403. Endocrinol. 1999; 50: 683-689.
Hambrecht R, Schulze PC, Gielen S, Linke A, Jessen N, Djurhuus CB, Jørgensen JO, Jensen LS,
Mobius-Winkler S, Erbs S, Kratzsch J, Schubert Møller N, Lund S, Schmitz O. Evidence against
A, Adams V, Schuler G. Effects of exercise train- a role for insulin-signaling proteins PI 3-kinase
ing on insulin-like growth factor-I expression and Akt in insulin resistance in human skeletal
in the skeletal muscle of non-cachectic patients muscle induced by short-term GH infusion. Am J
with chronic heart failure. Eur J Cardiovasc Prev Physiol Endocrinol Metab. 2005; 288: E194-E199.
Rehab. 2005; 12: 401-406. Jilka RL. Molecular and cellular mechanisms of
Han SL, Wan SL. Effect of teriparatide on bone the anabolic effect of intermittent PTH. Bone.
mineral density and fracture in postmenopausal 2007; 40: 1434-1446.
osteoporosis: Meta-analysis of randomised con- Jørgensen JO, Jessen N, Pedersen SB, Vestergaard
trolled trials. Int J Clin Pract. 2012; 66: 199-209. E, Gormsen L, Lund SA, Billestrup N. GH
Hardy R, Cooper MS. Adrenal gland and bone. receptor signaling in skeletal muscle and adipose
Arch Biochem Biophys. 2010; 503: 137-145. tissue in human subjects following exposure to
Hawke TJ, Garry DJ. Myogenic satellite cells: an intravenous GH bolus. Am J Physiol Endo-
Physiology to molecular biology. J Appl Physiol. crinol Metab. 2006; 291: E899-E905.
2001; 91: 534-551. Joseph A-M, Pilegaard H, Litvintsev A, Leick L,
Healy ML, Gibney J, Pentecost C, Croos P, Russell- Hood DA. Control of gene expression and mito-
Jones DL, Sönksen PH, Umpleby AM. Effects chondrial biogenesis in the muscular adaptation
of high-dose growth hormone on glucose and to endurance exercise. Assays Biochem. 2006;
glycerol metabolism at rest and during exercise 42: 13-29.
in endurance-trained athletes. J Clin Endocrinol Kadi F, Bonnerud P, Eriksson A, Thornell LE. The
Metab. 2006; 91: 320-327. expression of androgen receptors in human neck
Healy ML, Gibney J, Russell-Jones DL, Pentecost C, and limb muscles: Effects of training and self-
Croos P, Sönksen PH, Umpleby AM. High dose administration of androgenic-anabolic steroids.
growth hormone exerts an anabolic effect at rest Histochem Cell Biol. 2000; 113: 25-29.
and during exercise in endurance-trained ath- Kelly GS. Peripheral metabolism of thyroid
letes. J Clin Endocrinol Metab. 2003; 88: 5221-5226. hormones: A review. Altern Med Rev. 2000; 5:
Heineke J, Molkentin JD. Regulation of cardiac 306-333.
hypertrophy by intracellular signaling pathways. Kicman AT. Pharmacology of anabolic steroids.
Nat Rev Mol Cell Biol. 2006; 7: 589-600. Br J Pharmacol. 2008; 154: 502-521.
References 249

Kino T, Su YA, Chrousos GP. Human glucocorti- Lowell BB, Bassel-Duby R, Spiegelman BM.
coid receptor isoform β: Recent understanding Transcriptional co-activator PGC-1 α drives the
of its potential implications in physiology and formation of slow-twitch muscle fibres. Nature.
pathophysiology. Cell Mol Life Sci. 2009; 66: 2002; 418: 797-801.
3435-3448. Liu H, Bravata DM, Olkin I, Friedlander A, Liu V,
Klover P, Chen W, Zhu B-M, Hennighausen L. Roberts B, Bendavid E, Saynina O, Salpeter SR,
Skeletal muscle growth and fiber composi- Garber AM, Hoffman AR. Systematic review:
tion are regulated through the transcription The effects of growth hormone on athletic per-
factors STAT5a/b: Linking growth hormone formance. Ann Intern Med. 2008; 148: 747-758.
to the androgen receptor. FASEB J. 2009; 23: Long YC, Zierath JR. Influence of AMP-activated
3140-3148. protein kinase and calcineurin on metabolic
Koziris LP, Hickson RC, Chatterton RT Jr, Groseth networks in skeletal muscle. Am J Physiol. 2008;
RT, Christie JM, Goldflies DG, Unterman TG. 295: E545-E552.
Serum levels of total and free IGF-I and IGFBP-3 Loucks AB. The endocrine system: Integrated influ-
are increased and maintained in long-term train- ences on metabolism, growth, and reproduction.
ing. J Appl Physiol. 1999; 86: 1436-1442. In: Tipton CM, Sawka MN, Tate CA, Terjung
Kraemer RR, Blair MS, McCaferty R, Castracane RL, eds. ACSM’s Advanced Exercise Physiology.
VD. Running-induced alterations in growth hor- Philadelphia: Lippincott Williams & Wilkins;
mone, prolactin, triiodothyronine, and thyrox- 2006: 453-481.
ine concentrations in trained and untrained men Loucks AB, Thuma JR. Luteinizing hormone
and women. Res Q Exerc Sport. 2003; 64: 69-74. pulsatility is disrupted at a threshold of energy
Kraemer WJ, Marchitelli L, Gordon SE, Harman availability in regularly menstruating women. J
E, Dziados JE, Mello R, Frykman P, McCurry Clin Endocr Metab. 2003; 88: 297-311.
D, Fleck SJ. Hormonal and growth factor Ma XM, Blenis J. Molecular mechanisms of mTOR-
responses to heavy resistance exercise protocols. mediated translational control. Nature Rev Mol
J Appl Physiol. 1990; 69: 1442-1450. Cell Biol. 2009; 10: 307-318.
Kraemer WJ, Ratamess NA. Hormonal responses Manetta J, Brun JF, Maimoun L, Callis A, Préfaut
and adaptations to resistance exercise training. C, Mercier J. Effect of training on the GH/IGF-I
Sports Med. 2005; 35: 339-361. axis during exercise in middle-aged men: Rela-
Kvorning T, Andersen M, Brixen K, Madsen K. tionship to glucose homeostasis. Am J Physiol
Suppression of endogenous testosterone produc- Endocrinol Metab. 2002; 283: E929-E936.
tion attenuates the response to strength training: Marín-Garcia J. Thyroid hormone and myocardial
A randomized, placebo-controlled, and blinded mitochondrial biogenesis. Vasc Pharmacol. 2010;
intervention study. Am J Physiol Endocrinol 52: 120-130.
Metab. 2006; 291: E1325-E1332. Mason S, Johnson RS. The role of HIF-1 in hypoxic
Kvorning T, Andersen M, Brixen K, Schjerling P, response in skeletal muscle. Adv Exp Med Biol.
Suetta C, Madsen K. Suppression of testoster- 2007; 618: 229-244.
one does not blunt mRNA expression of myoD, Matheny RW Jr, Nindl, BC, Adamo ML. Mini-
myogenin, IGF, myostatin or androgen receptor review: Mechano-growth factor: A putative
post strength training in humans. J Physiol. product of IGF-I gene expression involved in
2007; 578: 579-593. tissue repair and regeneration. Endocrinology.
Lange KH, Andersen JL, Beyer N, Isaksson F, 2010; 151: 865-875.
Larsson B, Rasmussen MH, Juul A, Bülow J, Matheny W, Merritt E, Zannikos SV, Farrar RP,
Kjaer M. GH administration changes myosin Adamo ML. Serum IGF-I deficiency does not
heavy chain isoforms in skeletal muscle but does prevent compensatory hypertrophy in resistance
not augment muscle strength or hypertrophy, exercise. Exp Biol Med. 2009; 234: 164-170.
either alone or combined with resistance exercise Mauro A. Satellite cell of skeletal muscle fibers. J
training in healthy elderly men. J Clin Endocrinol Biophys Biochem Cytol. 1961; 9: 493-495.
Metab. 2002; 87: 513-523. Møller N, Jørgensen JO. Effects of growth hormone
Lange KH, Larsson B, Flyvbjerg A, Dall R, Ben- on glucose, lipid, and protein metabolism in
nekou M, Rasmussen MH, Orskov H, Kjaer M. human subjects. Endocr Rev. 2009; 30: 152-177.
Acute growth hormone administration causes Murgia M, Serrano AL, Calabria E, Pallafacchina
exaggerated increases in plasma lactate and G, Lomo T, Schiaffino S. Ras is involved in
glycerol during moderate to high intensity bicy- nerve-activity-dependent regulation of muscle
cling in trained young men. J Clin Endocrinol genes. Nature Cell Biol. 2000; 2: 142-147.
Metab. 2002; 87: 4966-4975. Navegantes LC, Baviera AM, Kettelhut IC. The
Lanning NJ, Carter-Su C. Recent advances in inhibitory role of sympathetic nervous system
growth hormone signaling. Rev Endocr Metab in the Ca2+-dependent proteolysis of skeletal
Disord. 2006; 7: 225-235. muscle. Braz J Med Biol Res. 2009; 42: 21-28.
Lin J, Wu H, Tarr PT, Zhang CY, Wu Z, Boss O, Navegantes LC, Migliorini RH, do Carmo Kettel-
Michael LF, Puigserver P, Isotani E, Olson EN, hut I. Adrenergic control of protein metabolism
250 References

in skeletal muscle. Curr Opin Clin Nutr Metab Rennie MJ, Bohé J, Smith K, Wackerhage H,
Care. 2002; 5: 281-286. Greenhaff P. Branched-chain amino acids as
Neri M, Bello S, Bonsignore A, Cantatore S, Riezzo fuels and anabolic signals in human muscle. J
I, Turillazzi E, Fineschi V. Anabolic androgenic Nutr. 2006; 136: 264S-268S.
steroids abuse and liver toxicity. Mini Rev Med Revollo JR, Cidlowski JA. Mechanisms generating
Chem. 2011; 11: 430-437. diversity in glucocorticoid receptor signaling.
Nindl BC, Pierce JR. Insulin-like growth factor I Glucocorticoids and mood. Ann NY Acad Sci.
as a biomarker of health, fitness, and training 2009; 1179: 167-178.
status. Med Sci Sports Exerc. 2010; 42: 39-49. Robling AG, Burr DB, Turner CH. Recovery
Nørrelund H, Møller N, Nair KS, Christiansen JS, periods restore mechanosensitivity to dynami-
Jørgensen JO. Continuation of growth hormone cally loaded bone. J Exp Biol. 2001; 204(part
(GH) substitution during fasting in GH-deficient 19): 3389-3399.
patients decreases urea excretion and conserves Roelen CA, de Vries WR, Koppeschaar HP, Ver-
protein synthesis. J Clin Endocrinol Metab. 2001; voorn C, Thijssen JH, Blankenstein MA. Plasma
86: 3120-3129. insulin-like growth factor-I and high affinity
O’Connell M, Robbins DC, Horton ES, Sims EA, growth hormone-binding protein levels increase
Danforth E Jr. Changes in serum concentrations after two weeks of strenuous physical training.
of 3,3',5'-triiodothyronine and 3,5,3'-triiodothy- Int J Sports Med. 1997; 18: 238-241.
ronine during prolonged moderate exercise. J Rudman D, Feller AG, Nagraj HS, Gergans GA,
Clin Endocr Metab. 1979; 49: 242-246. Lalitha PY, Goldberg AF, Schlenker RA, Cohn
Ojmaa K. Signaling mechanisms in thyroid L, Rudman IW, Mattson DE. Effects of human
hormone-induced cardiac hypertrophy. Vasc growth hormone in men over 60 years old. New
Pharmacol. 2010; 52: 113-119. Engl J Med. 1990; 323: 1-6.
Paddon-Jones D, Sheffield-Moore M, Urban RJ, Sakamoto K, Goodyear LJ. Exercise effects on
Aarsland A, Wolfe RR, Ferrando AA. The muscle insulin signaling and action. Invited
catabolic effects of prolonged inactivity and review: Intracellular signaling in contracting
acute hypercortisolemia are offset by dietary skeletal muscle. J Appl Physiol. 2002; 93: 369-383.
supplementation. J Clin Endocr Metab. 2005; Schakman O, Gilson H, Kalista S, Thissen JP.
90: 1453-1459. Mechanisms of muscle atrophy induced by glu-
Pappas M, Mourouzis K, Karageorgious H, Tes- cocorticoids. Horm Res. 2009; 72(suppl 1): 36-41.
seromatis C, Mourouzis I, Kostopanagiotou Schuenke MD, Kopchick JJ, Hikida RS, Kraemer
G, Pantos C, Cokkinos DV. Thyroid hormone WJ, Staron RS. Effects of growth hormone
modulates the responsiveness of rat aorta to overexpression vs. growth hormone receptor
alpha1-adrenergic stimulation: An effect due to gene disruption on mouse hindlimb muscle fiber
increased activation of beta2-adrenergic signal- type composition. Growth Hormone IGF Res.
ing. Int Angiol. 2009; 28: 474-478. 2008; 18: 479-486.
Pritzlaff CJ, Wideman L, Weltman JY, Abbott Sheffield-Moore M, Yeckel CW, Volpi E, Wolf
RD, Gutgesell ME, Hartman ML, Veldhuis JD, SE, Morio B, Chinkes DL, Paddon-Jones D,
Weltman A. Impact of acute exercise intensity Wolfe RR. Postexercise protein metabolism in
on pulsatile growth hormone release in men. J older and younger men following moderate-
Appl Physiol. 1999; 87: 498-504. intensity aerobic exercise. Am J Physiol. 2004;
Rabinowitz D, Zierler KL. A metabolic regulating 287: E513-E522.
device based on the actions of human growth Short KR, Møller N, Bigelow ML, Coenen-
hormone and of insulin, singly and together, on Schimke J, Nair KS. Enhancement of muscle
the human forearm. Nature. 1963; 199: 913-915. mitochondrial function by growth hormone. J
Rarick KR, Pikosky MA, Grediagin A, Smith TJ, Clin Endocrinol Metab. 2008; 93: 597-604.
Glickman EL, Alemany JA, Staab JS, Young AJ, Short KR, Vittone JL, Bigelow ML, Proctor
Nindl BC. Energy flux, more so than energy bal- DN, Nair KS. Age and aerobic exercise train-
ance, protein intake, or fitness level, influences ing effects on whole body and muscle protein
insulin-like growth factor-I system responses metabolism. Am J Physiol. 2004; 286: E92-E101.
during 7 days of increased physical activity. J Sinha-Hikim I, Cornford M, Gaytan H, Lee ML,
Appl Physiol. 2007; 103: 1613-1621. Bhasin S. Effects of testosterone supplementa-
Rasmussen BB, Phillips SM. Contractile and tion on skeletal muscle fiber hypertrophy and
nutritional regulation of human muscle growth. satellite cells in community-dwelling older men.
Exerc Sport Sci Rev. 2003; 31: 127-131. J Clin Endocrinol Metab. 2006; 91: 3024-3033.
Ratamess NA, Kraemer WJ, Volek JS, Maresh CM, Sinha-Hikim I, Taylor WE, Gonzalez-Cadavid
Vanheest JL, Sharman MJ, Rubin MR, French NF, Zheng W, Bhasin S. Androgen receptor
DN, Vescovi JD, Silvestre R, Hatfield DL, Fleck in human skeletal muscle and cultured muscle
SJ, Deschenes MR. Androgen receptor content satellite cells: Up-regulation by androgen
following heavy resistance exercise in men. J treatment. J Clin Endocrinol Metab. 2004; 89:
Steroid Biochem Mol Biol. 2005; 93: 35-42. 5245-5255.
References 251

Spangenburg EE, Le Roith D, Ward CW, Bodine tionship between age, percentage body fat,
SC. A functional insulin-like growth factor fitness, and 24-hour growth hormone release in
receptor is not necessary for load-induced healthy young adults: Effects of gender. J Clin
skeletal muscle hypertrophy. J Physiol. 2006; Endocrin Metab. 1994; 78: 543-548.
586: 283-291. Weltman A, Weltman JY, Schurrer R, Evans WS,
Steinacker JM, Brkic M, Simsch C, Nething Veldhuis JD, Rogol AD. Endurance train-
K, Kresz A, Prokopchuk O, Liu Y. Thyroid ing amplifies the pulsatile release of growth
hormones, cytokines, physical training and hormone: Effects of training intensity. J Appl
metabolic control. Horm Metab Res. 2005; 37: Physiol. 1992; 72: 2188-2196.
538-544. Widdowson WM, Healy M-L, Sönksen PH, Gibney
Sutton JR. Effect of acute hypoxia on the hormonal J. The physiology of growth hormone and sport.
response to exercise. J Appl Physiol. 1977; 42: Growth Hormone IGF Res. 2009; 19: 308-319.
587-592. Willoughby DS, Taylor L. Effects of sequential
Taniguchi CM, Emanuelli B, Kahn CR. Critical bouts of resistance exercise on androgen recep-
nodes in signaling pathways: Insights into insu- tor expression. Med Sci Sports Exerc. 2004; 36:
lin action. Nat Rev Mol Cell Biol. 2006; 7: 85-96. 1499-1506.
Ten Broek RW, Grefte S, Von den Hoff JW. Regu- Wilson JD. Androgen abuse by athletes. Endocr
latory factors and cell populations involved in Rev. 1988; 9: 181-199.
skeletal muscle regeneration. J Cell Physiol. Winder WW, Hagberg JM, Hickson RC, Ehsani
2010; 224: 7-16. AA, McLane JA. Time course of sympathoad-
Terjung RL, Tipton CM. Plasma thyroxine and renal adaptation to endurance exercise training
thyroid-stimulating hormone levels during in man. J Appl Physiol. 1978; 45: 370-374.
submaximal exercise in humans. Am J Physiol. Wu H, Naya FJ, McKinsey TA, Mercer B, Shel-
1971; 220: 1840-1845. ton JM, Chin ER, Simard AR, Michel RN,
Van Cauter E, Shapiro ET, Tillil H, Polonsky KS. Bassel-Duby R, Olson EN, Williams RS. MEF2
Circadian modulation of glucose and insulin responds to multiple calcium-regulated signals
responses to meals: relationship to cortisol in the control of skeletal muscle fiber type.
rhythm. Am J Physiol 1992, 2G2: E467-E475. EMBO J. 2000; 19: 1963-1973.
Vance ML, Mauras N. Growth hormone therapy Yarasheski KE, Campbell JA, Smith K, Rennie
in adults and children. New Engl J Med. 1999; MJ, Holloszy JO, Bier DM. Effect of growth
341: 1206-1216. hormone and resistance exercise on muscle
VanHelder WP, Casey K, Radomski MW. Regu- growth in young men. Am J Physiol. 1992; 262:
lation of growth hormone during exercise by E261-E267.
oxygen demand and availability. Eur J Appl Yarasheski KE, Zachweija JJ, Angelopoulos TJ,
Physiol. 1987; 56: 628-632. Bier DM. Short-term growth hormone treatment
Vescovo G, Ravara B, Gobbo V, Angelini A, Dalla does not increase muscle protein synthesis in
Libera L. Skeletal muscle fibres synthesis in experienced weight lifters. J Appl Physiol. 1993;
heart failure: Role of PGC-1, calcineurin and 74: 3073-3076.
GH. Int J Cardiol. 2005; 104: 298-306. Zhao JT, Cowley MJ, Lee P, Birzniece V, Kaplan
Viguerie N, Clement K, Barbe P, Courtine M, Benis W, Ho KK. Identification of novel GH-regulated
A, Larrouy D, Hanczar B, Pelloux V, Poitou C, pathway of lipid metabolism in adipose tissue:
Khalfallah Y, Barsh GS, Thalamas C, Zucker A gene expression study in hypopituitary men.
JD, Langin D. In vivo epinephrine-mediated J Clin Endocrinol Metab. 2011; 96: E1188-E1196.
regulation of gene expression in human skel-
etal muscle. J Clin Endocr Metab. 2004; 89: Chapter 9
2000-2014. Ahima RS, Prabakaran D, Flier JS. Postnatal
Vingren JL, Kraemer WJ, Hatfield DL, Volek leptin surge and regulation of circadian rhythm
JS, Ratamess NA, Anderson JM, Häkkinen of leptin by feeding. Implications for energy
K, Ahtiainen J, Fragala MS, Thomas GA, Ho homeostasis and neuroendocrine function. J
JY, Maresh CM. Effect of resistance exercise Clin Invest. 1998; 101: 1020-1027.
on muscle steroid receptor protein content in Almon RR, Yang E, Lai W, Androulakis IP, Ghim-
strength-trained men and women. Steroids. bovschi S, Hoffman EP, Jusko WJ, Dubois DC.
2009; 74(13-14): 1033-1039. Relationship between circadian rhythms and
Waxman DJ, O’Connor C. Growth hormone regu- modulation of gene expression by glucocorti-
lation of sex-dependent liver gene expression. coids in skeletal muscle. Am J Physiol. 2008;
Mol Endocrinol. 2006; 20: 2613-2629. 295: R1031-R1047.
Weiner RB, Baggish AL. Exercise-induced car- Angeles-Castellanos M, Mendoza J, Escobar C.
diac remodeling. Prog Cardiovasc Dis. 2012; Restricted feeding schedules phase shift daily
54: 380-386. rhythm of cFos and protein Per1 immunoreactiv-
Weltman A, Weltman JY, Hartman MI, Abbott ity in corticolimbic regions in rats. Neuroscience.
RD, Rogol AD, Evans WS, Veldhuis JD. Rela- 2007; 144: 344-355.
252 References

Anisimov SV, Boheler KP, Anisimov VN. Microar- Borer KT, Kuhns LR. Radiographic evidence for
ray technology in studying the effect of melato- acceleration of skeletal growth in adult hamsters
nin on gene expression in the mouse heart. Dokl by exercise. Growth. 1977; 41: 1-13.
Biol Sci. 2002; 383: 90-95. Borer KT, Nicoski D, Owens V. Alteration of the
Anisimov SV, Khavinson VK, Anisimov VN. pulsatile growth hormone secretion by growth-
Effect of melatonin and tetrapeptide on gene inducing exercise: Involvement of endogenous
expression in mouse brain. Bull Exp Biol Med. opiates and somatostatin. Endocrinology. 1986;
2004; 138: 504-509. 118: 844-850.
Anson RM, Guo Z, deCabo R, Iyun T, Rios M, Borer KT, Wuorinen E, Lukos J, Denver J, Porges
Hagepanos A, Ingram DK, Lane MA, Mattson S, Burant C. Two bouts of exercise before meals,
MP. Intermittent fasting dissociates beneficial but not after meals, lower fasting blood glucose.
effects of dietary restriction on glucose metabo- Med Sci Sports Exerc. 2009; 41: 1606-1614.
lism and neuronal resistance to injury from Brandenberger G, Follenius M. Influence of timing
calorie intake. Proc Natl Acad Sci USA. 2003; and intensity of muscular exercise on temporal
100: 6216-6220. patterns of plasma cortisol levels. J Clin Endocr
Antle MC, Steen NM, Mistlberger RE. Adenos- Metab. 1975; 40: 845-849.
ine and caffeine modulate circadian rhythms Bucchi A, Barbuti A, Difrancesco D, Baruscotti
in the Syrian hamster. Neuroreport. 2001; 12: M. Funny current and cardiac rhythm: Insights
2901-2905. from HCN knockout and transgenic mouse
Bamshad M, Aoki VT, Adkison MG, Warren WS, models. Front. Physiol. 2012, 3:240. doi: 10.3389/
Bartness TJ. Central nervous system origins of the fphys.2012.00240.
sympathetic nervous system outflow to white adi- Cailotto C, La Fleur SE, Van Heijningen C, Wortel
pose tissue. Am J Physiol. 1998; 275: R291-R299. J, Kalsbeek A, Feenstra M, Pevet P, Buijs RM.
Baruscotti M, Bucchi A, Viscomi C, Mandelli G, The suprachiasmatic nucleus controls the daily
Consalez G, Gnecchi-Rusconi T, Montano N, variation of plasma glucose via the autonomic
Casali KR, Micheloni S, Barbuti A, DiFran- output to the liver: Are the clock genes involved?
cesco D. Deep bradycardia and heart block Eur J Neurosci. 2005; 22: 2531-2540.
caused by inducible cardiac-specific knockout Chappell PE, White RS, Mellon PL. Circadian gene
of the pacemaker channel gene Hcn4. Proc Natl expression regulates pulsatile gonadotropin-
Acad Sci USA. 2011; 108: 1705-1710. releasing hormone (GnRH) secretory patterns
Belchetz PE, Plant TM, Nakai Y, Krogh KJ, in the hypothalamic GnRH secreting GT1-7 cell
Knobil E. Hypophysial responses to continu- line. J Neurosci. 2003; 23: 11202-11213.
ous and intermittent delivery of hypothalamic Cornélissen G, Halberg F, Otsuka K, Singh RB,
gonadotropin-releasing hormone. Science. 1978; Chen CH. Chronobiology predicts actual and
202: 631-633. proxy outcomes when dipping fails. Hyperten-
Bellido T, Ali AA, Plotkin LI, Fu Q, Gubrij I, sion. 2007; 49: 237-239.
Roberson PK, Weinstein RS, O’Brien CA, Escobar C, Diaz-Munoz M, Encinas F, Aguilar-
Manolagas SC, Jilka RL. Proteasomal degrada- Roblero R. Persistence of metabolic rhythmicity
tion of Runx2 shortens parathyroid hormone- during fasting and its entrainment by restricted
induced anti-apoptotic signaling in osteoblasts: feeding schedules in rats. Am J Physiol. 1998;
A putative explanation for why intermittent 274: R1309-R1316.
administration is needed for bone anabolism. J Frøberg J, Karlsson C-G, Levi L, Lidberg L. Cir-
Biol Chem. 2003; 278: 50259-50272. cadian variation in performance, psychological
Boden G, Ruiz J, Urbain JL, Chen X. Evidence for ratings, catecholamine excretion and diuresis
a circadian rhythm of insulin secretion. Am J during prolonged sleep deprivation. Int J Psy-
Physiol. 1996; 271: E246-E252. chobiol. 1972; 2: 23-36.
Borer KT, Campbell CS, Tabor J, Jorgenson K, Froy O. Metabolism and circadian rhythms—Impli-
Kandarian S, Gordon L. Exercise reverses cations for obesity. Endocr Rev. 2010; 31: 1-24.
photoperiodic anestrus in golden hamsters. Biol Giustina A, Veldhuis JD. Pathophysiology of the
Reprod. 1983; 29: 38-47. neuroregulation of growth hormone secretion
Borer KT, Cornelissen G, Halberg F, Brook R, in experimental animals and the human. Endocr
Rajagopalan S, Fay W. Circadian blood pressure Rev. 1998; 19: 717-797.
overswinging in a physically fit, normotensive Gronfier C, Brandenberger G. Ultradian rhythms
African American woman. Am J Hypertens. in pituitary and adrenal hormones: Their rela-
2002; 15: 827-830. tions to sleep. Sleep Med Rev. 1998; 2: 17-29.
Borer KT, Kaplan LR. Exercise-induced growth Gupta AK, Cornelissen G, Greenway FL, Dhoo-
in golden hamsters: Effects of age, weight, and pati V, Halberg F, Johnson WD. Abnormalities
activity level. Physiol Behav. 1977; 18: 29-34. in circadian blood pressure variability and endo-
Borer KT, Kelch RP. Increased serum growth thelial function: Pragmatic markers for adverse
hormone and somatic growth in exercising adult cardiometabolic profiles in asymptomatic obese
hamsters. Amer J Physiol. 1978; 234: E611-E616. adults. Cardiovasc Diabetol. 2010; 24(9): 58-70.
References 253

Halberg F. Circadian (about twenty-four-hour) Kalsbeek A, Yi CX, Cailotto C, La Fleur SE,


rhythms in experimental medicine. Proc R Soc Fliers E, Buijs RM. Mammalian clock output
Med. 1963; 56: 253-257. mechanisms. Essays Biochem. 2011; 49: 137-151.
Halberg F, Prikryl P, Danková E, Sigelová J, Dusek Kanaley JA, Weltman JY, Pieper KS, Weltman A,
J. Prophylactic aspirin treatment: The merits of Hartman ML. Cortisol and growth hormone
timing. JAMA. 1991; 266: 3128-3129. responses to exercise at different times of day. J
Hara R, Wan K, Wakamatsu K, Aida R, Moriya Clin Endocr Metab. 2001; 86: 2881-2889.
T, Akiyama M, Shibata S. Restricted feeding Kanaley JA, Weltman JY, Veldhuis JD, Rogol
entrains liver clock without participation of AD, Hartman ML, Weltman A. Human growth
the suprachiasmatic nucleus. Genes Cells. 2001; hormone response to repeated bouts of aerobic
6: 269-278. exercise. J Appl Physiol. 1997; 83: 1756-1761.
Hayes LD, Bickerstaff GF, Baker JS. Interactions Karatsoreos IN, Silver R. Minireview: The neuro-
of cortisol, testosterone, and resistance training: endocrinology of the suprachiasmatic nucleus as
Influence of circadian rhythms. Chronobiol Int. a conductor of body time in mammals. Endocri-
2010; 27: 675-705. nology. 2007; 148: 5640-5647.
Homolka P, Cornélissen G, Homolka A, Siegelova J, Kriegsfeld LJ, LeSauter J, Silver R. Targeted
Halberg F. Exercise-associated transient circadian microlesions reveal novel organization of the
hypertension (CHAT)? [Abstract] In: Proceedings hamster suprachiasmatic nucleus. J Neurosci.
of the 3rd International Conference, Civilization 2004; 24: 2449-2457.
Diseases in the Spirit of VI Vernadsky October La Fleur SE, Kalsbeek A, Wortel J, Buijs RM. A
10-12, 2005; Moscow, Russia. 2005: 419-421. suprachiasmatic nucleus generated rhythm in
Honma K, von Goetz C, Aschoff J. Effects of basal glucose concentrations. J Neuroendocrinol.
restricted daily feeding on freerunning circadian 1999; 11: 643-652.
rhythms in rats. Physiol Behav. 1983; 30: 905-913. Lamia KA, Sachdeva UM, DiTacchio L, Williams
Horne JA, Staff LH. Exercise and sleep: Body- EC, Alvarez JG, Egan DF, Vasquez DS, Juguilon
heating effects. Sleep. 1983; 6: 36-46. H, Panda S, Shaw RJ, Thompson CD, Evans
Isgaard J, Carlsson L, Isaksson OGP, Jansson J-O. RM. AMPK regulates the circadian clock by
Pulsatile intravenous growth hormone (GH) cryptochrome phosphorylation and degrada-
infusion to hypophysectomized rat increases tion. Science. 2009; 326: 437-440.
insulin-like growth factor I messenger ribo- Landolt HP. Genetic determination of sleep EEG
nucleic acid in skeletal tissues more effectively profiles in healthy humans. Prog Brain Res.
than continuous GH infusion. Endocrinology. 2011; 193: 51-61.
1988; 123: 2605-2610. Langlais PJ, Hall T. Thiamine deficiency-induced
Isgaard J, Møller C, Isaksson OG, Nilsson A, disruptions in the diurnal rhythm and regulation
Mathews CS, Norsted TC. Regulation of insulin- of body temperature in the rat. Metab Brain Dis.
like growth factor ribonucleic acid in rat growth 1998; 13: 225-239.
plate by growth hormone. Endocrinology. 1988; Lefta M, Wolff G, Esser KA. Circadian rhythms,
122: 1515-1520. the molecular clock, and skeletal muscle. Curr
Iwanaga H, Yano M, Miki H, Okada K, Azama T, Top Dev Biol. 2011; 96: 231-271.
Takiguchi S, Fujiwara Y, Yasuda T, Nakayama Levine H, Saltzman W, Yankaskas J, Halberg F.
M, Kobayashi M, Oishi K, Ishida N, Nagai Circadian state dependent effect of exercise
K, Monden M. Per2 gene expressions in the upon blood pressure in clinically healthy men.
suprachiasmatic nucleus and liver differentially Chronobiologia. 1977; 4: 129-130 (abstract).
respond to nutrition factors in rats. J Parenter Liu C, Li S, Liu T, Borjigin J, Lin JD. Transcrip-
Enteral Nutr. 2005; 29: 157-161. tional coactivator PGC-1alpha integrates the
Jilka RL. Molecular and cellular mechanisms of mammalian clock and energy metabolism.
the anabolic effect of intermittent PTH. Bone. Nature. 2007; 447: 477-481.
2007; 40: 1434-1446. Lloyd D, Murray DB. Ultradian metronome: Time-
Kakolewski JW, Deaux E, Christensen J, Case B. keeper for orchestration of cellular coherence.
Diurnal patterns in water and food intake and Trends Biochem Sci. 2005; 30: 373-377.
body weight changes in rats with hypothalamic Loucks AB, Thuma JR. Luteinizing hormone
lesions. Am J Physiol. 1971; 221: 711-718. pulsatility is disrupted at a threshold of energy
Kalsbeek A, Fliers E, Romijn JA, La Fleur SE, availability in regularly menstruating women. J
Wortel J, Bakker O, Endert E, Buijs RM. The Clin Endocr Metab. 2003; 88: 297-311.
suprachiasmatic nucleus generates the diurnal Macchi MM, Bruce JN. Human pineal physiology
changes in plasma leptin levels. Endocrinology. and functional significance of melatonin. Front
2001; 142: 2677-2685. Neuroendocrinol. 2004; 25: 177-195.
Kalsbeek A, Ruiter M, La Fleur SE, Cailotto C, Martino T, Arab S, Straume M, Belsham DD, Tata N,
Kreier F, Buijs RM. The hypothalamic clock and Cai F, Liu P, Triveri M, Ralph M, Sole MJ. Day/
its control of glucose homeostasis. Prog Brain night rhythms in gene expression of the normal
Res. 2006; 153: 283-307. murine heart. J Mol Med. 2004; 82: 256-264.
254 References

Maywood ES, O’Neill JS, Chesham JE, Hastings signs, norepinephrine, epinephrine, thyroid hor-
MH. Minireview: The circadian clockwork of mones, and cortisol in schizophrenia. Psychiatry
the suprachiasmatic nuclei—Analysis of a cel- Res. 1995; 57: 21-39.
lular oscillator that drives endocrine rhythms. Reilly T, Atkinson G, Waterhouse J. Biological
Endocrinology. 2007; 148: 5624-5634. Rhythms and Exercise. New York: Oxford Uni-
McCarthy JJ, Andrews JL, McDearmon EL, versity Press; 1997a.
Campbell KS, Barber BK, Miller BH, Walker Reilly T, Atkinson G, Waterhouse J. Travel fatigue
JR, Hogenesch JB, Takahashi JS, Esser KA. and jet lag. J Sport Sci. 1997b; 15: 365-369.
Identification of the circadian transcriptome in Reppert SM, Weaver DR. Coordination of cir-
adult mouse skeletal muscle. Physiol Genomics. cadian timing in mammals. Nature. 2002; 418:
2007; 31: 86-95. 935-941.
Mellani L, Wolff G, Esser KA. Circadian rhythms, Robling AG, Burr DB, Turner CH. Recovery peri-
the molecular clock, and skeletal muscle. Curr ods restore mechanosensitivity to dynamically
Topics Dev Biol. 2011; 96: 231-271. loaded bone. J Exp Biol. 2001; 204: 3389-3399.
Mendoza J, Drevet K, Pevet P, Challet E. Daily Sakamoto K, Ishida N. Light-induced phase-shifts
meal timing is not necessary for resetting the in the circadian expression rhythm of mam-
main circadian clock by calorie restriction. J malian period genes in the mouse heart. Eur J
Neuroendocrinol. 2008; 20: 251-260. Neurosci. 2000; 12: 4003-4006.
Mieda M, Williams SC, Richardson A, Tanaka K, Shen J, Tanida M, Niijima A, Nagai K. In vivo
Yanagisawa M. The dorsomedial hypothalamic effects of leptin on autonomic nerve activity and
nucleus as a putative food-entrainable circadian lipolysis in rats. Neurosci Lett. 2007; 416: 193-197.
pacemaker. Proc Natl Acad Sci USA. 2006; 103: Shephard RJ. Sleep, biorhythms, and human per-
12150-12155. formance. Sports Med. 1984; 1: 11-34.
Minors D, Waterhouse J. Circadian Rhythms and Shirai H, Oishi K, Ishida N. Bidirectional CLOCK/
the Human. Bristol: Wright; 1981. BMAL1-dependent circadian gene regulation
Mistlberger RE, Mumby DG. The limbic system by retinoic acid in vitro. Biochem Biophys Res
and food-anticipatory circadian rhythms in the Commun. 2006; 351: 387-391.
rat: Ablation and dopamine blocking studies. Sikon A, Batur P. Profile of teriparatide in the
Behav Brain Res. 1992; 47: 159-168. management of postmenopausal osteoporosis.
Mohri T, Emoto N, Nonaka H, Fukuya H, Yagita Int J Wom Health. 2010; 2: 37-44.
K, Okamura H, Yokoyama M. Alterations of Singh RB, Pella D, Otsuka K, Halberg F, Corne-
circadian expressions of clock genes in Dahl salt- lissen G. New insights into circadian aspects
sensitive rats fed a high-salt diet. Hypertension. of health and disease. J Assoc Physicians India.
2003; 42: 189-194. 2002; 50: 1416-1425.
Monteleone P, Maj M, Fusco M, Orazzo C, Kemali Spanagel R, Rosenwasser AM, Schumann G,
D. Physical exercise at night blunts the nocturnal Sarkar DK. Alcohol consumption and the body’s
increase in plasma melatonin levels in healthy biological clock. Alcohol Clin Exp Res. 2005;
humans. Life Sci. 1990; 47: 1989-1995. 29: 1550-1557.
Morgan PJ, Ross AW, Mercer JG, Barrett P. Spiegel K, Leproult R, Van Cauter E. Impact of
Photoperiodic programming of body weight sleep debt on metabolic and endocrine function.
through the neuroendocrine hypothalamus. J Lancet. 1999; 354: 1435-1439.
Endocrinol. 2003; 177: 27-34. Staels B. When the clock stops ticking, metabolic
Nakahata Y, Kaluzova M, Griomaldi B, Sahar S, syndrome explodes. Nat Med. 2006; 12: 54-55.
Hirayama J, Chen D, Guarente LP, Sassone- Stephan FK. The “other” circadian system: Food
Corsi P. The NAD+-dependent deacetylase as a Zeitgeber. J Biol Rhythms. 2002; 17: 284-292.
SIRT1 modulates CLOCK-mediated chromatin Stephan FK, Davidson AJ. Glucose, but not fat,
remodeling and circadian control. Cell. 2008; phase shifts the feeding-entrained circadian
134: 329-340. clock. Physiol Behav. 1998; 65: 277-288.
Pritzlaff CJ, Wideman L, Weltman JY, Abbott Stephan FK, Zucker I. Circadian rhythms in drink-
RD, Gutgesell ME, Hartman ML, Veldhuis JD, ing behavior and locomotor activity of rats are
Weltman A. Impact of acute exercise intensity eliminated by hypothalamic lesions. Proc Natl
on pulsatile growth hormone release in men. J Acad Sci USA. 1972; 69: 1583-1586.
Appl Physiol. 1999; 87: 498-504. Surya S, Horowitz JF, Goldenberg N, Sakharova
Quinaglia T, Martins LC, Figueiredo VN, Santos A, Harber M, Cornford AS, Symons K, Barkan
RC, Yugar-Toledo JC, Martin JF, Demacq C, AL. The pattern of growth hormone delivery
Pimenta E, Calhoun DA, Moreno H Jr. Non-dip- to peripheral tissues determines insulin-like
ping pattern relates to endothelial dysfunction growth factor-1 and lipolytic responses in obese
in patients with uncontrolled resistant hyperten- subjects. J Clin Endocrinol Metab. 2009; 94:
sion. J Hum Hypertens. 2011; 25: 656-664. 2828-2834.
Rao ML, Strebel B, Halaris A, Gross G, Braunig P, Thayer JF, Yamamoto SS, Brosschot JF. The
Huber G, Marler M. Circadian rhythm of vital relationship of autonomic imbalance, heart
References 255

rate variability and cardiovascular disease risk ing non-22-kilodalton growth hormone isoforms
factors. Int J Cardiol. 2010; 141: 122-131. in short children: A possible mechanism for
Thompson DL, Weltman JY, Rogol AD, Metzger growth failure. J Clin Endocr Metab. 1997; 82:
DL, Veldhuis JD, Weltman A. Cholinergic and 2944-2949.
opioid involvement in release of growth hormone Borer KT, Kelch RP, Hayashida T. Hamster growth
during exercise and recovery. J Appl Physiol. hormone: Species specificity and physiological
1993; 75: 870-878. changes in blood and pituitary concentrations as
Van Cauter E, Kerkhofs M, Caufriez A, Van Onder- measured by a homologous radioimmunoassay.
bergen A, Thorner MO, Copinschi G. A quanti- Neuroendocrinology. 1982; 35: 349-358.
tative estimation of growth hormone secretion Borer KT, Wuorinen E, Ku K, Burant C. Appetite
in normal man: Reproducibility and relation to responds to changes in meal content, whereas
sleep and time of day. J Clin Endocrinol Metab. ghrelin, leptin, and insulin track changes in
1992; 74: 1441-1450. energy availability. J Clin Endocr Metab. 2009;
Van Reeth O, Sturis J, Byrne MM, Blackman JD, 94: 2290-2298.
L’Hermite-Baleriaux M, Leproult R, Oliner C, Convertino VA, Keil LC, Greenleaf JE. Plasma
Refetoff S, Turek FW, Van Cauter E. Nocturnal volume, renin, and vasopressin responses to
exercise phase delays circadian rhythms of mela- graded exercise after training. J Appl Physiol.
tonin and thyrotrophin secretion in normal men. 1983; 54: 508-514.
Am J Physiol. 1994; 266: E964-E974. Corwin RL, Gibbs J, Smith GP. Increased food
Veldhuis JD, Patrie J, Wideman L, Patterson M, intake after type A but not type B cholecysto-
Weltman JY, Weltman A. Contrasting negative- kinin receptor blockade. Physiol Behav. 1991;
feedback control of endogenously driven and 50: 255-258.
exercise-stimulated pulsatile growth hormone Davies CTM, Few JD. Effects of exercise on
secretion in women and men. J Clin Endo Metab. adrenocortical function. J Appl Physiol. 1973;
2004; 89: 840-846. 35: 887-891.
Yamanaka Y, Honma S, Honma K. Scheduled Durnam DM, Palmiter RD. A practical approach
exposures to a novel environment with a running for quantifying specific mRNAs by solution
wheel differentially accelerate re-entrainment of hybridization. Anal Biochem. 1983; 131: 385-393.
mice peripheral clocks to new light-dark cycles. Evans WS, Faria AC, Christiansen E, Ho KY,
Genes Cells. 2008; 13: 497-507. Weiss J, Rogol AD, Johnson ML, Blizzard
Young ME. The circadian clock within the heart: RM, Veldhuis JD, Thorner MO. Impact of
Potential influence on myocardial gene expres- intensive venous sampling on characterization
sion, metabolism, and function. Am J Physiol. of pulsatile GH release. Am J Physiol. 1987;
2006; 290: H1-H16. 252: 549-556.
Foley JE, Chen YD, Lardinois CK, Hollenbeck CB,
Chapter 10 Liu GC, Reaven GM. Estimates of in vivo insu-
Aizawa K, Iemitsu M, Maeda S, Otsuki T, Sato K, lin action in humans: Comparison of the insulin
Ushida T, Mesaki N, Akimoto T. Acute exercise clamp and the minimal model techniques. Horm
activates local bioactive androgen metabolism Metab Res. 1985; 17: 406-409.
in skeletal muscle. Steroids. 2010; 75: 219-223. Frøberg J, Karlsson C-G, Levi L, Lidberg L. Cir-
Arbogast LA, Voogt JL. Progesterone induces cadian variation in performance, psychological
dephosphorylation and inactivation of tyrosine ratings, catecholamine excretion and dieresis
hydroxylase in rat hypothalamic dopaminergic during prolonged sleep deprivation. Int J Psy-
neurons. Neuroendocrinology. 2002; 75: 273-281. chobiol. 1972; 2: 23-36.
Belchetz PE, Plant TM, Nakai Y, Krogh KJ, Isgaard J, Carlsson L, Isaksson OGP, Jansson J-O.
Knobil E. Hypophyseal responses to continu- Pulsatile intravenous growth hormone (GH)
ous and intermittent delivery of hypothalamic infusion to hypophysectomized rat increases
gonadotropin-releasing hormone. Science. 1978; insulin-like growth factor I messenger ribo-
202: 631-633. nucleic acid in skeletal tissues more effectively
Bergman RN. The minimal model of glucose than continuous GH infusion. Endocrinology.
regulation: A biography.  Adv Exp Med Biol. 1988; 123: 2605-2610.
2003; 537: 1-19. Isgaard J, Møller C, Isaksson OG, Nilsson A,
Bjarnason R, Banerjee K, Rose SJ, Rosberg S, Mathews CS, Norsted TC. Regulation of insulin-
Metherell L, Clark AJ, Albertsson-Wikland like growth factor ribonucleic acid in rat growth
K, Savage MO. Spontaneous growth hormone plate by growth hormone. Endocrinology. 1988;
secretory characteristics in children with partial 122: 1515-1520.
growth hormone insensitivity. Clin Endocrinol. Ives SJ, Blegen M, Coughlin MA, Redmond J,
2002; 57: 357-361. Matthews T, Paolone V. Salivary estradiol,
Boguszewski CL, Jansson C, Boguszewski MCS, interleukin-6 production, and the relationship to
Rosberg S, Carlsson B, Albertsson-Wikland K, substrate metabolism during exercise in females.
Carlsson LMS. Increased proportion of circulat- Eur J Appl Physiol. 2011; 111: 1649-1658.
256 References

Jessup SK, Malow BA, Symons KV, Barkan AL. Polonsky KS, Given BD, Hirsch LJ, Tillil H, Sha-
Blockade of endogenous growth hormone- piro ET, Beebe C, Frank BH, Galloway JA, Van
releasing hormone receptors dissociates noc- Cauter E. Abnormal patterns of insulin secre-
turnal growth hormone secretion and slow-wave tion in non-insulin-dependent diabetes mellitus.
sleep. Eur J Endocrinol. 2004; 151: 561-566. New Engl J Med. 1988; 318: 1231-1239.
Jilka RL. Molecular and cellular mechanisms of Pørksen N. The in vivo regulation of pulsatile
the anabolic effect of intermittent PTH. Bone. insulin secretion. Diabetologia. 2002; 45: 3-20.
2007; 40: 1434-1446. Schölin A, Nyström L, Arnqvist H, Bolinder J,
Johnson ML, Virostko A, Veldhuis JD, Evans WS. Björk E, Berne C, Karlsson FA, the Diabetes
Deconvolution analysis as a hormone pulse- Incidence Study Group in Sweden (DISS).
detection algorithm.  Methods Enzymol. 2004; Proinsulin/C-peptide ratio, glucagon and remis-
384: 40-54. sion in new-onset Type 1 diabetes mellitus in
Kitaoka R, Fujikawa T, Miyaki T, Matsumura S, young adults. Diabet Med. 2011; 28: 156-161.
Fushiki T, Inoue K. Increased noradrenergic Siegrist-Kaiser CA, Pauli V, Juge-Aubry CE, Boss
activity in the ventromedial hypothalamus O, Pernin A, Chin WW, Cusin I, Rohner-Jean-
during treadmill running in rats. J Nutr Sci renaud F, Burger AG, Zapf J, Meier CA. Direct
Vitaminol. 2010; 56: 185-190. effects of leptin on brown and white adipose
Kozaki T, Lee S, Nishimura T, Katsuura T, Yasuk- tissue. J Clin Invest. 1997; 100: 2858-2864.
ouchi A. Effects of saliva collection using cotton Surya S, Horowitz JF, Goldenberg N, Sakharova
swabs on melatonin enzyme immunoassay. J A, Harber M, Cornford AS, Symons K, Barkan
Circadian Rhythms. 2011; 9: 1. AL. The pattern of growth hormone delivery
Kraemer WJ, Nindl BC, Marx JO, Gotshalk LA, to peripheral tissues determines insulin-like
Bush JA, Welsch JR, Volek JS, Spiering BA, growth factor-1 and lipolytic responses in
Maresh CM, Mastro AM, Hymer WC. Chronic obese subjects. J Clin Endocr Metab. 2009; 94:
resistance training in women potentiates growth 2828-2834.
hormone in vivo bioactivity: Characterization Thomas GA, Kraemer WJ, Kennett MJ, Comstock
of molecular mass variants. Am J Physiol. 2006; BA, Maresh CM, Denegar CR, Volek JS, Hymer
291: E1177-E1187. WC. Immunoreactive and bioactive growth
Lange KH, Lorentsen J, Isaksson F, Simonsen L, hormone responses to resistance exercise in
Bülow J, Kjaer M. Lipolysis in human adipose men who are lean or obese. J Appl Physiol. 2011;
tissue during exercise: Comparison of micro- 111: 465-472.
dialysis and a-v measurements. J Appl Physiol. Thuma JR, Gilders R, Verdun M, Loucks AB.
2002; 92: 1310-1316. Circadian rhythm of cortisol confounds cortisol
Li CH, Lostroh AJ. Bioassay of hypophyseal responses to exercise: Implications for future
growth hormone in hypophysectomized mice research. J Appl Physiol. 1995; 78: 1657-1664.
by the tibia test. Endocrinology. 1957; 60: 308-317. Veldhuis JD, Johnson ML. Cluster analysis: A
Loucks AB, Mortola JF, Girton L, Yen SS. Altera- simple, versatile, and robust algorithm for
tions in the hypothalamic-pituitary-ovarian endocrine pulse detection. Am J Physiol. 1986;
and the hypothalamic-pituitary-adrenal axes 250: E486-E493.
in athletic women. J Clin Endocrinol Metab. Veldhuis JD, Straume M, Iranmanesh A, Mulligan
1989; 68: 402-411. T, Jaffe C, Barkan A, Johnson ML, Pincus S.
Mano-Otagiri A, Ohata H, Iwasaki-Sekino A, Secretory process regularity monitors neuro-
Nemoto T, Shibasaki T. Ghrelin suppresses nor- endocrine feedback and feedforward signaling
adrenaline release in the brown adipose tissue strength in humans. Am J Physiol. 2001; 280:
of rats. J Endocr. 2009; 201: 341-349. R721-R729.
Payne AH, Youngblood GL. Regulation of expres- Weltman A, Pritzlaff CJ, Wideman L, Considine
sion of steroidogenic enzymes in Leydig cells. RV, Fryburg DA, Gutgesell ME, Hartman ML,
Biol Reprod. 1995; 52: 217-225. Veldhuis JD. Intensity of acute exercise does
Pencek RR, James FD, Lacy DB, Jabbour K, Wil- not affect serum leptin concentrations in young
liams PE, Fueger PT, Wasserman DH. Exercise- men. Med Sci Sports Exerc. 2000; 32: 1556-1561.
induced changes in insulin and glucagon are not Wilkerson JE, Gutin B, Horvath SM. Exercise-
required for enhanced hepatic glucose uptake induced changes in blood, red cell, and plasma
after exercise but influence the fate of glucose volumes in man. Med Sci Sports. 1977; 9: 155-158.
within the liver. Diabetes. 2004; 53: 3041-3047.
Index
Note: Page numbers followed by an italicized f or t refer to the figure or table on that page, respectively.

A activation 44f, 45, 50, 114 B


acetyl CoA 103, 105 effect on muscle 183 Bad gene 189, 190f
acetyl-coenzyme synthetase (ACSS) gene transcription with 184 baroreceptors 56, 57f
109, 110f isoforms 44 baroreflex 66f, 70f, 73f, 81, 95
Achilles tendon 160 response to energy depletion 184, Barr body 175
actin cytoskeleton 36, 37, 41 194 β-catenin 41, 42, 192
activator protein-1 (AP-1) 23, 46, 47, 185 amphetamines 75 Bcl-2 gene 42, 48, 189, 190f, 214
acute exercise 46, 162-164, 168, 186. See anabolic steroids 196-198 beige adipocytes 36, 182, 183
also exercise intensity androgen receptors (ARs) 123, 150, Bernard, Claude 56
acyl ghrelin 134 153f, 159, 186-187. See also BH4 (H4 biopterin cofactor) 34f, 35
adaptive thermogenesis 35-36 reproductive hormones binding proteins (BPs) 15, 18, 39
adenosine 67 androgens. See reproductive hormones bioassay 224, 227
adenyl cyclase (C) 20, 34f, 35, 104, 108 angiogenesis 183, 188 biological rhythms 201-202, 202-203.
adipogenesis 43, 44 angiotensin-converting enzyme (ACE) See also circadian rhythms;
adiponectin 50, 114, 124 84 ultradian rhythms
adipose hormones 7t angiotensin II 2, 64-65, 66f, 84, 90, 91 control by food-entrainable oscilla-
adipose tissue. See body fat anorexia nervosa 134, 138 tor 209
adipose triglyceride lipase (ATGL) 108 anterior cruciate ligament (ACL) inju- effects on training adaptations 214-
adrenal cortex 11 ries 160 215
adrenal hormones 7t, 12f antidiuretic hormone (ADH). See argi- exercise as rhythm modulator 211,
adrenal medulla 4, 5 nine vasopressin (AVP) 212-214
adrenergic receptors (AdRs) 2, 21f, apoptosis 40, 99f and heart health 216-217
59, 89 actin related 160 and hormone secretion measure-
alpha 23, 24f effect of ROS on 48, 49f ment 226
beta 20, 23, 59, 61 genes 157, 189, 192 nutrient ingestion effects 212, 215
desensitization 60 of osteoclasts 157, 158f, 192 parameters of 202-203
in fat tissue 108, 111, 161 aquaporin 87 zeitgebers 202, 203, 209, 212, 213
adrenocorticotropic hormone (ACTH) archvillin 159 blood cell hormones 8t
12, 16, 17, 73, 74f, 89, 116, 190 arcuate nucleus (ARC) 128, 129, blood flow
aerobic capacity 158 143, 144 cutaneous 81, 83
aerobic exercise. See endurance train- arginine vasopressin (AVP) 65f, 66 to muscle 65, 67
ing; exercise intensity action in kidney tubule 87f, 88f regional by exercise intensity 62,
afferent signals 56, 69 role in fluid balance 85-87, 89, 90, 63f
aging 39, 61, 152, 161, 192 91, 92 blood pressure
Akt (protein kinase B) 33f, 35, 40, 41, 45 secretion by hypovolemia and angiotensin II effects 65
–mTOR pathway 38f, 98, 99, 182, plasma osmolarity 86f, 88f change with postural changes 69,
189, 195 arterial gases 56, 71 70
P13K activation 38, 39f, 154, 180, AS160 protein 45 circadian rhythm 62, 63f, 216, 217
193 ATF-4 191 during exercise 64-65, 67, 73
substrates 100 ATP 44, 80, 105, 107, 184 and plasma volume 83, 84
translocation of GLUT-4 100 atrial natriuretic peptide (ANP) 65f, role of central command 72-73
alanine 105, 190 91-92, 107 blood vessels 62. See also vasocon-
aldo-keto reductase 1C (AKR1C) 161 atrogin 1 39, 40 striction
aldosterone 65f, 66f, 84, 89, 90, 91, 92 autocrine messengers 5 dilatation 34, 67, 80, 81
amenorrhea 156, 167, 168, 170 autonomic nervous system (ANS) 1, hormones 8t
amino acids 191-192, 199 55. See also neurotransmitters innervation 5, 58
dietary supplementation 123, 192- afferent signals to 56, 58 respiratory 69
193 efferent components 3, 58 body fat 36, 143-144, 167, 170, 218
gluconeogenesis from 105 and enteric system control 5 AdRs by location of fat depot 108,
hormones 10-11 functions of 2 111, 161
5-aminoimidazole-4-carboxamide- hormones of 1, 3-4 in energy equation 128
1-rosiglitazone (AICAR) 45 organization 3f energy storage capacity 107
AMP-activated protein kinase (AMPK) parasympathetic division 5 homeostatic lipostatic regulation
32, 38, 40f, 44 sympathetic division 4-5 128-130

257
258 Index

body fat (continued) coordination of functions 71, 72f core temperature (Tc). See thermo-
hormones from adipose tissue 7t ergogenic aids 74-76, 76t regulation
in relation to physical activity 137f, cardiovascular disease 50, 171 Cori cycle 105, 106
184 cardiovascular drift 83 corticosteroids 12f, 61
visceral versus subcutaneous 111, carnitine palmitoyltransferase I (CPT I) corticotropin-releasing hormone (CRH)
112, 161 45, 103, 104, 105 16, 135, 164, 170, 190, 207
body mass 127, 128, 145-146. See also casein kinase 1ϵ (CK1ϵ) 204 cortisol 166f, 168, 170, 190-191
body fat; weight loss caspases 49f circadian rhythms 203f, 211f, 226-
compartmental regulation 143-144 catecholamines 1, 50, 61, 189. See also 227
homeostatic versus nonhomeostatic epinephrine (E); norepineph- effects on bone 192
regulation 139, 142 rine (NE) effects on skeletal muscle 121, 191-
increases 128 biosynthesis 10, 11f 192
role of physical activity in regulation concentrations and exercise inten- exercise effects 115f, 116-117, 211f,
137-138, 146-147 sity 61f, 110, 112, 115f, 181f 212-213
setpoint 128, 143, 146-147 as ergogenic aids 74 genomic action of 191f
body mass index (BMI) 145, 170, 218 measurement 225 positive feedback loop 73, 74f
bone receptor sensitivity to 60-61 postprandial levels 193f
actions of sex hormones 157-158 role in training adaptations 180-181, Cortisol-glucocorticoid receptor alpha
cortisol effects 192 189 complex (CGRα) 190, 191
formation 43, 44, 157, 189-190 S-innervation 4f cranial nerves 56, 72f
leptin effects 144, 145f Cd36 32 cryptochrome (Cry) genes 203, 204f,
loss 157, 170 C/ EBP (CCAAT/enhancer-binding 210
mechanical loading 32, 40-43, 214 protein) 43 Cx43 41
mineral density 157, 170, 192 cell adhesion 41 cycl ic adenosi ne monophosphate
PTH effects 189-190, 214 cell growth cycle 48, 49f (cAMP) 97, 104, 154, 210, 214
resorption 189-190, 214 cellular hypoxia 182, 183 cyclin 48, 189
bone cells 7t, 40, 41, 157, 158f cellular proliferation 99f, 214, 221 cyclin-dependent kinases (CDKs) 48
bone morphogenetic protein (BMP) central nervous system (CNS) 1 Cyp genes 29, 155, 157, 185
42, 44 c-fos 47, 180f, 185 cytochrome c 49f
Bötzinger neurons 71 cGMP 24, 34f, 35, 104, 108 cytochrome oxidase (COX) enzymes
Boyar, R.M. 166 CHAT (circadian hyperamplitude ten- 183, 184
bradykinin 67, 84 sion) 216, 217 cytokine receptors 20
brain 85f chemical messengers 1 cytokines 35, 46, 114, 157, 214
blood-brain-barrier 84, 85, 137 cholecystokinin (CCK) 2, 222, 224
control circuits 58, 136f, 145f cholesterol 11, 14, 124 D
hormones 8t CIDEA 109, 110f deconvolution analysis 225, 228
medullary nuclei 72f ciliary neurotrophic factor (CNTF) dehydration 82
bronchoconstriction 69, 75 50, 188 cellular 81, 83, 84-85, 90
bronchodilatation 69 circadian rhythms 202, 204f, 215 distribution of water losses 85t
Bürge, J. 92 body temperature 208, 212 voluntary 90, 94-95
clock mechanism 203-204 dexamethasone (DEX) 17
C of cortisol 212-213, 226-227 diabetes 124, 218
Cabanac, Michel 134 effect on performance 212 prevalence 50
C (adenyl cyclase) 34f, 35, 104, 108 endocrine 206-208 prevention through exercise 50-53
caffeine 33, 74-75, 123, 210, 212 glucose control 208 diacylglycerol (DAG) 15, 23, 108, 109
calcineurin (Cn) 33, 35, 39f, 182, 184 heart 62, 63f digestion 56, 58
calcitriol 12, 14, 15 jet lag 212 diuresis 86f, 92, 93f
calcium 20, 23 lipid metabolism 208 DNA damage 46, 48, 49f
absorption 157 locomotion 208 doping 196-198
release for signal transduction metabolic events 208-209
33-36 nutrient effects 209, 212, 215 E
sensing enzymes 184 in peripheral tissues 205 eccentric exercise 37, 46
Calmodulin (CaM) 23, 33-35 of reproductive hormones 207 efferent neurons 1, 3
calpain 181, 192 sleep-wakefulness 203f, 208, 210, eicosanoids 15
calpastatin 180, 181 218 electrostatic fields 32, 33f
CaMKK signaling 33, 44, 45, 184 synchronization by suprachiasmatic Elk 185, 199
cAMP response element-binding protein nucleus 205-209, 206f Embden-Meyerhof glycolytic pathway
(CREBP) 22, 42f, 105, 185 circumventricular organs (CVOs) 84, 101f, 103
Cannon, Walter 58, 136 85f endocrine system 1, 2
Ca 2+ pumps 33 citrate synthase (Cs) 32 endothelial NO synthetase (eNOS)
carbohydrate metabolism 100-103, 101f, c-jun 47, 157, 185 34, 43, 67
107f. See also glucose c-jun terminal kinases 1 and 2 (JNK1/2) endothelin (ET-1) 23, 65, 66
carbon dioxide (pCO2) 56, 71 46, 47 endurance training 45, 194
cardiac muscle 21, 59, 64, 181, 182, 183 c-myc 48, 185 adaptations to 178
cardiac output 63f, 87, 178, 179 collagens 42 catecholamines role 180-181
cardiorespiratory function concentric exercise 37 intracellular signaling pathways
control 69 contraceptive effects 171 182-184
Index 259

role of growth hormone 179-180 and AMPK activation 44 exercise timing 215
sex hormones 164-167 cardiac output by 63f feast-famine cycle 142, 143f
systemic hormones in adaptations cardiovascular measures 64f finickiness 134, 136
178-179 catecholamine concentration 61f, gastric contractions 136
thyroid hormones 181-182 110, 112, 115f, 116, 181f GI hormone effects 137
energy availability fuel mobilization/use by 109-112, hedonic eating 134, 135f, 136
and food intake 130-132, 133f 114-117 homeostatic versus nonhomeostatic
hormone changes 168, 169f and GH secretion 110, 111f control 139, 142
restricted 166-167, 168, 170, 209, and hunger 130, 131f hunger sensation 132f, 133f, 134-
221 hydromineral hormone increases 136, 140
tissue sensing of abundance 194- 65f, 89 leptin effects 130, 134, 135, 144,
195 recovery from low- to moderate- 145f
tissue sensing of depletion 193-194, 113-114 opportunistic eating 139-142
199 recovery from moderate- to high- Paleolithic diet 142, 145-146
energy balance 97, 104-109, 128. See 118-119 postexercise supplementation 130,
also food intake; insulin on renal function 89f 131f, 193
energy equation as 128 extracellular signal-regulated kinases regulation of 130-132
fuel use by exercise intensity 109- (ERK1/2) 43, 46, 47, 98, 99, satiation 132f, 133f, 134, 140
112, 114-117 185, 199 “thrifty genes” 142
fuel use during resistance exercise Forkhead box TF FOXO 38
118-119 F FOXO 38, 39, 40, 192
hormonal mediation of 97-98, 114- fatigue 80, 92 Frank-Starling mechanism 59
115f fatty acids 15, 35, 103, 161. See also free free fatty acids (FFAs) 103, 104, 105,
enteric nervous system 5 fatty acids (FFAs) 107, 108
enzyme-linked immunosorbent assay fatty acid synthetase (FAS) 21, 103 female sex hormone effects on 173
(ELISA) 223-224 female sex hormones 12, 150-151. See fuel use with exercise intensity 115f,
ephedrine 75 also estrogen 117f
epinephrine (E) contraceptive use 171 free radicals. See reactive oxygen spe-
effects on gene expression 181, 189 effects on circulation 171 cies (ROS)
heart activation 59 effects on muscle function 172 frizzled transmembrane spanning pro-
and plasma glucose 104, 112 effects on substrate utilization 172- tein (Fz) 41f, 42
release by adrenal positive feedback 173 fuel utilization
73, 74f effects on thermoregulation during by exercise intensity 109-112, 114-
respiratory system activation 69 exercise 172 117
stimulation of lipolysis 107-108, 112 effects on ventilation 172 female sex hormone effects on 173
synthesis 73 secretion with exercise 168, 170 hormones acting through GPCR
ergogenic aids 74-76, 76t fibronectin 42 20-24
ergosterol 14, 15 fight or flight response 58 of muscle fibers 111, 112, 114, 117,
erythropoietin (EPO) 15, 20 filamin C 159 118
estradiol (E2) 7t, 151, 153f, 157, 170 flavine adenine dinucleotide (FAD) 34 during resistance exercise 118-119
estrogen. See also female sex hormones; fluid balance 80, 95-96. See also dehy-
reproductive hormones dration; plasma osmolarity; G
and bone turnover 157 sweating gamma-amino- butyric-acid (GABA)
conversion from androgens 12, 150, fluid intake 85, 90, 91, 92, 93 206, 207, 208
152 heat loss reflexes 80, 81 gastrin 16, 222
deficiency 157 hyperhydration 91-92 gastrin-releasing peptide (GRP) 206
effect on GHRH 154 hyponatremia 92, 96 gastrointestinal hormones 8t, 137
effect on growth hormone 121 renal reabsorptive function 87, 89 gene expression 32
during menstrual cycle 171, 226 strategies for 85, 90-91, 92-93 action of cortisol 191f
estrogen receptors 41, 42, 154 thirst 90-91 action of growth hormone 180f, 185
eukaryotic elongation factors (eEFs) urine output to NaCl ingestion 93f with AMP-activated protein kinase
33f, 39f, 40f fndc5 36, 36f, 183 (AMPK) 184
eukaryotic translation initiation factors focal adhesion kinase (FAK) 40f and biological rhythms 203, 204,
(eIFs) 33f, 38, 39f, 40f, 185 follicle-stimulating hormone (FSH) 12, 205, 210, 211
exercise. See also biological rhythms; 15, 17, 151, 152, 221 catecholamines 181, 189
endurance training; resistance food intake effects of resistance exercise on 123
exercise (RE) alliesthesia 134 insulin effects 103
body heat load 80 appetite suppression 98, 130, 131f oxidative genes 184
high-impact 43, 44 and biological rhythms 208, 209 with PGC-1α 183
during pregnancy 173-174 brain neuronal circuits 136f survival genes 42, 159, 189, 214
timing of 215 control mechanisms 134-137 thyroid effects on 182
exercise-associated collapse 94-95 descriptors of feeding behavior 134 and vibration 43-44
exercise hyperaemia 62, 66-67, 68f and energy availability 130-132, ghrelin 134, 135, 136, 140f, 141,
exercise-induced hemoconcentration 133f, 139-140 208, 218
89, 220 evolutionary insights into 139-142, glucagon 17, 103, 104, 106f
exercise intensity 143f exercise response 112, 114f
activation of PDH by 116f exercise effects 130 production 98
260 Index

glucagon-like peptide-1 (GLP-1) 5, 16, measurement of 221, 222, 224 hormone–receptor binding 6, 26,
137, 194, 220 and oxygen delivery to muscle 179 28f, 154
glucocorticoid (GR) gene 190-191 response to RE 120f, 121, 122, 123 affinity/sensitivity 27
glucocorticoid hormones 12f, 25, 84, role in endurance training adapta- specificity 27, 29
105, 190, 207. See also cortisol tions 179-180 hormone receptors. See also membrane
glucokinase (GK) 52, 98, 100-101, 102f, role in muscle hypertrophy 121, receptors
103, 215 184-186, 214 classification 18
gluconeogenesis 35, 105-106, 112, 113, self-administered 198 effect of resistance exercise on 123
116, 211 signaling pathways 185 intracellular and nuclear 25-26
glucose 50, 117f. See also glucose variants 121, 227 membrane 18-24
uptake growth hormone receptors (GHRs) 18, structure 19f
lowering blood levels 53 20, 180f hormones
sensing cells 98 growth hormone-releasing hormone actions 6f, 177-178
setpoint 215 (GHRH) 17, 154, 155, 209, 224 of autonomic nervous system 3-4
storage 32 guanine-protein coupled receptors chemical structure 6
transport 51-52, 98, 100 (GPCR) 20, 22-24 classification 5-6
glucose-alanine cycle 105, 106 guanylyl cyclase (GC) 34f, 35 glandular sources 7t
glucose-dependent insulinotropic pep- half-life 9, 221, 228
tide (GIP) 5, 137, 141f, 142 H nonglandular sources 8-9t
glucose-6-phosphate (G6P) 51, 101, Had 32, 35 pulsatile fluctuation 221, 228
104, 105, 106 Halberg, Franz 202, 215 release 6f, 15-17
glucose-phosphorylating enzyme (GK) Hcn4 210 transport 18
98, 100-102, 113, 142, 215 heart. See also blood pressure; cardio- hormone-sensitive lipase (HSL) 21,
glucose tolerance test 123, 124, 225 respiratory function 108, 109
glucose uptake 32, 46f, 50-53, 98, 99f biological rhythms 62, 63f, 215, hormone supplementation 196-198
during exercise recovery 113, 118 216-217, 218 Hsp 47
insulin-dependent 100 hormones 8t hybridization assays 225
insulin-independent 45, 46f innervation 57f hydrogen (H+) concentration 71
stimulation by IL-6 114, 116f pacemaker depolarization 60f, 62, hydrogen peroxide (H2O2) 35, 44, 47
Glut4 32, 35, 45, 100, 184 210 hydrolase (H) 34f, 35
glutamate 106, 207 pain signals 56 21-hydroxylase deficiency 29, 175
glycerol 107, 115f, 117f PS control of 61-62 hyd r oxyst er oid d ehyd r ogena se
glycogen depletion 99, 105, 108, 112, S control of 58-61 (11βHSD) 84, 109, 123, 185
113-114, 184, 215 stroke volume 60f hyperglycemia 17, 103, 110, 117
glycogenolysis 104-105, 112, 116 ultradian rhythms 62, 189 hyperhydration 91-92
and Ca 2+ binding 34 ventricular hypertrophy 181, 182, 189 hyperinsulinemia 143
hepatic 104, 107f, 111, 112, 113 heart rate 216 hypertension 61, 76, 182, 216, 218
glycogen storage 53 orthostatic reflex 70 hyperthyroidism 61
glycogen synthase kinase 3 (GSK3) 33f, resting 62 hypoglycemia 104, 113, 142
38, 39f, 40f, 41, 98, 99, 103, 189 variability in supine versus upright hyponatremia 92, 96
glycogen synthetase 21 61, 62f hypotension 95
glycolytic pathways 101f, 102f, 103 heat acclimation 82. See also thermo- hypothalamic neurons 1, 65, 66, 81,
glyconeogenesis 98, 99f, 103, 106f regulation 84, 92, 128
glycoproteins 10 heat stress. See thermoregulation hypothalamic–pituitary–gonadal (HPG)
Goldberg, A. 36, 184, 224 heme biosynthesis 183 axis 164, 165, 166
gona dot ropi n-rele a si ng hor mone heme oxidase 1 (HO1) 47, 184 hypovolemia 81, 83-84, 87, 88f, 90
(GnRH) 8t, 150, 165, 170, hemoglobin 158 hypoxia 47, 61
209, 210 HK 101, 113, 142, 184, 215
gonadotropins 12, 15, 17, 151, 152, 221 Hk2 32, 45 I
G protein-coupled ER (GPER) 154 homeostasis 2, 15, 55 IκK kinase 47, 192
Grb-2–SOS complex 19, 46, 98, 99 hormone measurement 219 immune response 35, 211
G regulatory proteins 20, 21f, 23, 61 assays 222-224 incretin 5
growth factors 32, 33f, 37, 187, 188, biological actions 224-225 inositol triphosphate (IP 3) 23, 34,
192, 214. See also insulin-like blood samples 219-220 84, 189
growth factor-I (IGF-I) collection medium 219-221 insulin 168, 169f
growth hormone (GH) 154, 155, 156, concentration 219-224 antagonists 97, 109
168, 169f confidence in 227 biosynthesis of fuel stores 103
bone growth 157 production site 225 during exercise 112, 114f
circadian control 203f rate of secretion 225, 228 exercise recovery 113
daily pulsatile pattern 109f salivary 220-221 gene translation effect 103
estrogen effects 121 sampling frequency 221-222, 226- glucose uptake 100
exercise effects 110, 115f, 213 227 and hunger 134, 135
genomic actions of 180f, 185 synthesis 225 and lipolysis 104, 107, 108
and lipolysis 107, 108-109, 110f, tissue samples 220 metabolic actions of 98-100
112, 179-180 urinary 221 postprandial levels 193f
Index 261

release 98 L-citrulline 34f, 35 m itogen-activated protein k inase


signaling pathways 98, 99f leptin 50, 170, 207, 208, 222 (MEK) 19, 47f, 99, 185, 199
stimulation of carbohydrate metabo- and body fat regulation 129, 144 mitogen-activated protein kinases
lism 100-103 effect on bone mass 143, 144, 145f (MAPK) 37, 40f, 43, 46-47, 47f.
structure 9f exercise response 114, 168, 169f See also Ras-MAPK pathway
suppression of fuel mobilization 104 food intake effects 134, 135, 141f, monoacylglycerol lipase (MGL) 108
synthesis byproducts 225 144, 145f motoneurons 32
insulin-like growth factor-I (IGF-I) 157, resistance 130 mRNA 33f, 39, 225
168, 169f, 225 leptin receptors (LepRb) 129, 144 mTOR 37, 38, 39f, 40, 40f, 185, 195, 199
exercise response 121, 186 leukotrienes 15 MuRF1 39, 40
isoforms 38, 186, 188 ligaments 160 muscle
role in muscle hypertrophy 37f, 38, lipid messengers 11, 15 actions of reproductive hormones
39, 123, 185, 186, 187-188 lipid metabolic genes 35, 211 158-160
signaling 39f lipid metabolism 116, 117, 181, 190, cardiac 21, 59, 64, 181, 182, 183
upregulation by growth hormone 208, 211, 218 effects of cortisol 191-192
121, 185 lipogenesis 98, 101f, 103 hormones 8t
insulin-like growth factor-II (IGF- lipolysis 50, 99f, 106-109, 161, 224 mechanical loading 36-40
II) 37f control of 107 reproductive hormones 160
insulin receptors (IRs) 18-19, 98 by exercise intensity 111, 112, 114 muscle atrophy 39, 121, 190, 191f, 192
insulin receptor substrate (IRS) 18, 98, and growth hormone 107, 108-109, muscle contraction
99, 100, 188 110f, 179-180 and Ca 2+ binding 34, 46
insulin resistance 50, 53, 109, 179 and insulin 104, 105, 107, 108 effects on blood glucose 50, 53
insulin sensitivity 51, 53, 124, 184, 225 stimulation by IL-6 114, 116f muscle fibers 158, 159, 191
integrins 36, 42 lipoprotein lipase (LPL) 103, 114 conversion 32, 35, 45
intense exercise 46 lipoprotein receptor-related protein fuel use 111, 112, 114, 117, 118
interleukin-6 (IL-6) 50, 114, 116f, 157 (Lrp5) 41f, 42 sex hormones effects 158, 159
interleukins 8t, 20, 35, 46, 50, 188. See liver 155 muscle hypertrophy 36-37, 39f, 118-
also interleukin-6 (IL-6) enzyme genes 211 119, 159
intermediolateral (IML) 4, 56, 58, 71 glucose production capacity 106 androgens 123, 186-187
intracrine messengers 5 hormones 8t antagonists to 183
intramyocellular triglycerides (IMTGs) LKB1 44, 45 catecholamines 189
107, 114 Lpl 32 circadian influences 215
iodine 11 lungs. See respiratory system growth hormone (GH) 120f, 121,
irisin 36, 183 luteinizing hormone (LH) 7t, 12, 166f, 184-186
IRMA (immunoradiometric assay) 223 167f, 168f insulin-like growth factor-I role in
ischemia 47 in female athletes 166-167, 214-215 37f, 38, 39, 123, 185, 186
isometric exercise 65, 81 in males 164 RE protocols for 119-120, 123, 124
with resistance exercise 165f satellite cells 187-189
J through life span 151f muscle wasting 40
Janus activated kinase-2 (JAK-2) 20, LY294002 38 myocyte enhancer factor 2 (MEF2) 183
180f, 185 myogenic regulatory factors (MRFs) 35,
JNK 46, 47 M 38, 39, 40, 187, 188
joint mechanoreceptors 56 macrophage colony-stimulating factor myoglobin 35
jump training 43-44 (M-CSF) 157, 158 myosin heavy chain (MHC) 181,
juxtaglomerular apparatus (JGA) 84 malonyl-CoA 45 182, 184
m a m m a l ia n t a rget of r apa myci n myostatin 38, 39, 40, 187, 188
K (mTOR) 32, 33f
Kennedy, Gordon 128 mechano growth factor (MGFs) 38, N
ketone bodies 168, 169f 123, 186, 188 NADPH+ (nicotinamide adenine dinu-
ketosis 113 medulla oblongata (MO) 16, 56 cleotide phosphate) 34, 46
kidneys Mef2 35, 36f, 183, 184 National Weight Control Registry
angiotensin effects 66f, 84 melatonin 11 146, 147
AVP action in tubule 88f circadian control 203f, 207 negative feedback 17
effects of exercise intensity 89f exercise effects 213f neuroendocrine cells 1, 74
hormones 8t membrane receptors 18, 19f, 23-24 neurotransmitters 1, 2, 3-4
reabsorptive function 84, 85-87, 89, menstrual cycle 151, 152f, 170, 171, 226. neurotrophic factors (NTFs) 159, 188
91 See also female sex hormones Nfat 39f, 184
suppression 65 mesenchymal stem cells (MSCs) 42, nicotine 75
Krebs cycle 100, 101f, 103, 106f 43, 187 nicotinic receptors 20, 59
metabolic syndrome 50, 124, 218 nitric oxide (NO) 34, 67, 68f
L metformin (N,N-dimethylimidodicar- nonsteroidal anti-inflammatory drugs
lactate threshold (LT) 111f bonimidic diamide) 45 92
lactic acid (LA) 116, 118, 119, 121, MHC isoforms 181, 182, 184 norepinephrine (NE) 4, 59, 61, 61f,
123, 165f mitochondrial biogenesis 35, 182, 121-122
L-arginine 34f, 35, 67 183, 184 NO synthetase (NOS) 35, 67, 81
262 Index

Nrf1 35, 36f, 45 phosphatidic acid phosphatase (LPIN1) with testosterone 159, 186
nuclear factor-κB 40, 46, 47, 48f, 192 109, 110f through muscle loading 37, 38, 39,
nuclear respiratory factors (NRF1/2) phosphatidylinositol biphosphate (PIP2) 40, 194
183, 184 18 proteolysis 38, 39, 40, 191
nucleus of the tractus solitarius (NTS) phosphatidylinositol-3-kinase (PI3K) puberty 155-157, 159f
56, 57f, 58 32, 33f, 38, 39f, 40f, 41, 98, 100 pulmonary system. See respiratory
nutrition. See energy availability; food P13–Akt and MAPK 154, 185, 188 system
intake P13K–mTOR signaling pathway 32, pyruvate 100, 101f, 102, 105, 116, 130
33f, 38f, 188 pyruvate dehydrogenase (PDH) 100,
O phosphodiesterase (PDE) 22f, 23, 24, 103, 116
obesity 50, 145, 218. See also body fat; 97, 98, 104
diabetes; weight loss phosphoenolpyruvate carboxykinase R
and dietary restriction during preg- (PEPCK) 22, 104, 105, 106, 190 Rac 46, 160f
nancy 173 phosphofructokinase (PFK) 101, 102, radioimmunoassay methods (RIAs)
hedonic eating 136 105 219, 222-223
and hypoactivity 137-138 3-phosphoinositide-dependent protein Raf 99, 185
and leptin 129, 130 kinase-1 (PDK-1) 33f, 40f, 100 rapamycin 38
prevalence 127 phospholipases C (PLC) 23 Ras 20, 46, 47, 99, 185
oligomenorrhea 170 phosphorylase 21, 33 Ras-MAPK pathway 185, 195
orthostatic reflex 70 pineal gland 7t, 207 reactive oxygen species (ROS) 15, 183
osteoblasts. See bone cells pituitary hormones 1, 7t, 17, 73, 85, activation of MAPK 46, 47f
osteocalcin (OCN) 143, 144 86f, 150 activation of NF-κB 47, 48f
osteocytes. See bone cells PKC 23, 185 apoptosis influence 48, 49f
osteogenesis 43, 44, 157, 189-190 plasma osmolarity 65, 83, 84, 90, 92 effects on cell function 46
osteoporosis 170, 214 and AVP release 88f oxygen derivatives 46
osteopenia 157, 170 changes by exercise intensity and receptor activator of nuclear factor
overweight 50, 127 dehydration 85t (RANKL) 35, 43, 157, 192
oxidative phosphorylation (OXPHOS) and skin blood flow 81 REDD1 191
44, 182, 183 plasma volume (PV) 2, 79. See also renin 64, 65f, 84, 91
oxygen (pO2) 56, 71 hypovolemia reproductive hormones 7t, 149, 152,
oxytocin (OXY) 15, 16, 85 and AVP stimulation 65 174, 176. See also female sex
effect of fluid intake 91 hormones; luteinizing hormone
P and hemoconcentration 89, 220 (LH); testosterone
p38 46, 47 response to endurance training 178, actions on adipose tissue 161
p50 47, 48f 179 actions on bone 157-158
p53 48, 49f sodium concentration 81, 82 actions on muscle 158-160
p65 47, 48f, 192 pons 62 biosynthesis 12, 13f
p70S6 kinase 39, 40f, 191, 195 postprandial lipemia 124 circadian rhythms of 207
p90RSK 199 powerlifting 43 effects on performance 170, 171-173
P450 genes 150, 185 pregnancy and exercise 173-174 and energy availability 169f
pain 56, 65 proinflammatory cytokines 46, 157, exercise during pregnancy 173-174
pancreatic hormones 7t, 98 188, 192 genomic and nongenomic actions
paracrine messengers 5 proinsulin 9f 152, 153f, 154, 155
parasympathetic (PS) nerves 3, 5 prolactin 7t, 203f response to resistance training 161,
control of heart function 61-62 proliferator-activated receptor coacti- 164f, 165f, 186-187
hyperactivity 143 vator-1α (PGC-1α) 35-36, 38, secretion across life span 150-152
parathyroid hormone (PTH) 7t, 189- 40, 45, 182, 183 secretion in habitual exercise 161,
190, 190f, 214 pro-opiomelanocortin (POMC) 7t, 10f 164-168, 170
paraventricular hypothalamic nucleus prostacyclins 15 secretion with acute exercise 161,
(PVN) 16, 65, 135, 164, 207 prostaglandins 8t, 15, 41, 42, 43 162-164
paxillin 159 protein degradation 39-40, 105, 159, 189 sex dimorphism during puberty
peptide hormones 7, 9-10, 16, 225 cortisol-induced 191f, 192, 193 155-157
performance inhibition 98, 180-181 resistance exercise (RE) 40, 177, 215.
circadian rhythm effects 212 protein hormones 7, 9-10, 225 See also muscle hypertrophy
effects of female hormones 171 protein kinase A (PKA) 20, 21, 181, dietary supplementation effects 123
effects of testosterone 171 185, 214 effects on genes 123
with heat stress 80, 92 protein kinase B (PKB). See Akt effects on hormone receptors 123,
perilipins 108 protein kinase G (PKG) 34f, 35, 185 186-187
period (Per) genes 203, 204f, 205, 206 protein kinases 33, 49f, 102 effects on muscle enzymes 123
peroxisome proliferator-activated recep- protein synthesis 33f, 38f, 188f, 199. See endocrine responses to 119
tors (PPARs) 25, 26, 35, 43, also muscle hypertrophy fuel use during 118-119
44, 204 cortisol effects 178, 191, 192 hormone responses to 119-123,
PGE2 41, 42, 43 exercise and meal timing effects 164f, 165f, 184
pH 56 193, 215 metabolic effects of 123-124
phenylethanolamine-N-methyltransfer- with growth hormone 180, 185 during pregnancy 174
ase (PNMT) 11f, 17, 73 insulin action 98, 99f, 194, 195 protocols 119
Index 263

respiratory system 56, 62, 69. See also stress hormones 17, 73, 170, 220. See hydrolysis 109, 110f
cardiorespiratory function also cortisol intramuscular 114
resting energy expenditure (REE) stretch response 36, 38f resistance exercise effects on 124
123-124 sudomotor reflex 80, 81, 95 storage 43f, 103
Rheb (Ras homolog enriched in brain) superoxide 35 synthesis 98, 101f, 103
33f, 194, 195 suprachiasmatic nucleus (SCN) 205- troponin 33, 35
ribosomal p70S6 kinase 37 209 tryptophan 11
rosiglitazone 45 survival genes 42, 159, 189, 214 tuberous sclerosis complex (TSC) 33f,
rostral ventrolateral medulla (RVLM) sweating 80, 92-93, 180 38, 40f, 195, 199
62, 65, 71, 73 consequences of fluid loss 83-87 tyrosine 10, 11f, 98, 190
running 43, 92, 94-95, 164, 214 heat loss through 80f tyrosine kinase 20, 154, 187
runt-related TF 2 (Runx2) 42, 43, 44, Na concentration 81-82, 83f
189, 214 and Na hunger 29, 66f, 90-91, 92 U
neural control of 81, 82f ubiquitin proteasome system 40,
S sympathetic nerves 1, 3, 4-5, 89 171, 189
satellite cells 159, 187-189 control of heart function 58-61 ultradian rhythms 202, 212, 215
secretin 5, 16 during exercise 110-111, 117 control of 209-211
secretory vesicles 11 nerve ending 59f heart 62, 189
Seimon, R.V. 136 respiratory activation 69 ultraviolet (UV) light 14
sensory receptors 56, 58 and vasoconstriction 64-66 uncoupling protein (UCP) 36, 182,
serotonin 11, 75, 76, 144, 145f, 207 183, 184
sex hormones. See reproductive hor- T urine
mones TAK1 44, 45 AVP effects on output 85
sex identification in sport 175 Tanner stages 156 circadian control of volume 203f
Shc–Grb-2–SOS 46, 47f TBC1D1 protein 45, 100 flow by exercise intensity 89f
Shc–Ras–MEK–ERK1/2 20 tendons 160 for hormone measurement 221
shear stress 41, 42, 43 tension forces 43 hypertonicity of 86-87
signal transducers and activators of testosterone 150, 164, 166f, 167f, output with NaCl ingestion 93
transcription (STAT factors) 168f, 171 water loss through 90, 91, 92
20, 114, 155, 185, 191 and bone growth 157
signal transduction 31-32, 33f and muscle growth 159, 186, 187 V
by electrostatic events 32 neonatal exposure to 155 vagus nerve 5, 57f, 61
by intracellular calcium release production 150, 152 vascular resistance 70f, 171, 179, 182
33-36 receptor binding 26f vasculature 62
intracellular pathways in endurance with resistance exercise 121, 122f, vasoactive intestinal polypeptide (VIP)
training 182-184 123, 165f, 186, 187 206
motoneuronal 32 thermoregulation 79, 80-82, 83, 89. See vasoconstriction 62, 64-66, 87, 89
to muscle mechanical loading also fluid balance; sweating vasodilatation 34, 67, 80, 81
36-40 circadian control 203f, 212 vasomotor reflex 80, 95
in response to bone mechanical female sex hormone effects 172 ventilation 69, 172
loading 40-43 growth hormone effects 180 ventromedial hypothalamus (VMH)
in response to sensing energy need thiazolidinedione (TZD) drugs 45 128, 129, 143, 144
44-45, 46f threonine 32, 45 vibration 43-44
skin thromboxanes 15 vitamin D 12, 14-15, 14f
blood flow 81, 83 Thuma, J.R. 226
hormones 8t thymus hormones 8t W
slow myosin 35 thyroid hormones 7t, 11, 17 Washburn, A.L. 136
smooth endoplasmic reticulum 11 endurance training 181-182 weight gain 128, 139
sodium (Na). See also fluid balance; gene transcription activation 182 weight loss. See also body mass
sweating synthesis 11f body defenses against 128, 139,
concentration in sweat 81-82, 83f thyroid receptors (TR) 25, 26, 182 145-146
Na hunger 29, 66f, 90-91, 92 thyroid-stimulating hormone (TSH) 17, as index of body water loss 92
reabsorption in kidney 66f, 84 181, 182, 210, 213f maintenance of 128
somatrotropin-release inhibiting hor- thyroxine (T4) 2 methods 140, 147f
mone (SRIF) 5, 154, 180, tibial thickness 42-43 volitional dehydration 94
209, 224 TORC1 194, 195, 199 white adipose tissue (WAT) 36, 182,
SON neurons 16 training adaptations 177-178. See also 183, 185
sperm parameters 165, 167f endurance training; muscle Williams, N.I. 168
sport activities 42-43 hypertrophy Wing, Rena 146
Src 18, 20, 154, 188 training periodicity 215 Wnt 41, 42, 189
steroid hormones 11-12, 14-15. See also transcription factors (TF) expression
reproductive hormones 22, 43, 154, 155, 185, 204 Y
binding 18 transforming growth factor beta (TGFb) Yalow, Rosalyn 223
biosynthesis 13f 44, 109, 188, 192
measurement 225 triglycerides (TG)
storage 15 breakdown enzymes 108
About the Author
Katarina T. Borer, PhD, is a professor in the
School of Kinesiology at the University of Michi-
gan in Ann Arbor, where she has spent over 35
years teaching and researching the hormonal
control of metabolism, particularly in response
to exercise. She has spent 40 years researching
endocrine mechanisms operating in acceleration
of growth by exercise and regulation of energy
balance. Borer also developed and validated
radioimmunoassay for hamster growth hormone
and prolactin.
She is a member of the American College of
Sports Medicine, Endocrine Society, American
Diabetes Association, American Physiological
Society, and Society for the Study of Ingestive
Behavior. In 1991, Borer received a Fulbright scholarship to the University of Gothenburg,
Sweden, where she studied the expression of IGF-I mRNA in exercising hamsters. She has
been a visiting professor on the kinesiology faculty at the University of Zagreb in Croatia
since 2002. Borer was also awarded the title of Meritorious Professor in 2010 from the
University of Zagreb.
Borer and her husband, Paul Wenger, reside in Ann Arbor. Borer enjoys spending time
with her grandchildren, painting, studying art, listening to opera and classical music, and
devoting time to her environmental interests, especially the recycling of resources.

264

You might also like