Autonomic Nervous System - Basic and Clinical Aspects

Download as pdf or txt
Download as pdf or txt
You are on page 1of 393

Daniel Pedro Cardinali

Autonomic
Nervous System

Basic and Clinical Aspects

123
Autonomic Nervous System
Daniel Pedro Cardinali

Autonomic Nervous
System
Basic and Clinical Aspects
Daniel Pedro Cardinali
Facultad de Ciencias Médicas
Pontificia Universidad Católica Argentina
Buenos Aires, Capital Federal
Argentina

ISBN 978-3-319-57570-4    ISBN 978-3-319-57571-1 (eBook)


DOI 10.1007/978-3-319-57571-1

Library of Congress Control Number: 2017949889

© Springer International Publishing AG 2018


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recita-
tion, broadcasting, reproduction on microfilms or in any other physical way, and transmission or infor-
mation storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publica-
tion does not imply, even in the absence of a specific statement, that such names are exempt from the
relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, express or implied, with respect to the material contained herein or for any errors
or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims
in published maps and institutional affiliations.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer International Publishing AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Prologue

The autonomic nervous system (ANS) is an important component of the nervous


system consisting of a complex set of neurons and neural pathways that control the
function of the various visceral organ systems. The overall function of the ANS is to
maintain the body homeostasis and to react adaptively to changes in the external and
internal milieu.
The ANS innervates the heart, the smooth muscle in all the organs, the abdomi-
nal viscera, the exocrine and endocrine glands, and the immune system. Thus, the
ANS participates in the regulation of breathing, circulation, digestion, metabolism,
and the internal milieu, exocrine and endocrine gland secretion, immune responses,
body temperature, and reproduction [1, 2]. Unfortunately, such medical importance
tends to be underscored in many books on physiology or neural sciences in which
the subject takes up much less space than that accorded, for example, to somatosen-
sory or cognitive functions.
The basic structure and operation of the ANS were defined at the beginning of
the last century, primarily by Gaskell and Langley, who recognized its two main
divisions: the sympathetic and the parasympathetic [3]. Furthermore, Langley des-
ignated the enteric nervous system as a third division based on the submucous
plexus of Meissner and the myenteric plexus of Auerbach located in the wall of the
gastrointestinal tract, albeit controlled by the sympathetic and parasympathetic divi-
sions. Overcoming the classical concept of a purely efferent system, it is presently
accepted that the ANS is composed of visceral afferents, integration centers, par-
ticularly in the brainstem, hypothalamus, and limbic cortex, and visceral sympa-
thetic and parasympathetic efferents; thus, the ANS extends both into the central
nervous system (CNS) and to the periphery.
Conceptually, the bio-psycho-social-ecological nature of the individual is truly
expressed by the function of his or her ANS. Its name is misleading because none of
the components shows “autonomy” in an integrated body. Nor are they solely “pas-
sive” or generated “without elaboration by mind.” All body systems are dependent
and affected by the action of others in a multicellular organization.
These dynamic relationships are the core of homeostasis, a key concept in physiol-
ogy. “Homeostasis” is used today to define not only the strategies that allow the body’s
proper response to changes in the environment (reactive homeostasis), but also the
remarkably developed, temporal mechanisms that allow the body to predict the timing
of environmental stimuli (predictive homeostasis based on biological rhythms).

v
vi Prologue

Autonomic reflexes are mediated by neural pathways in the brainstem and spinal
cord and generally regulate organ and system performance very rapidly (in millisec-
onds). Autonomic control is also mediated by specific brain regions, such as the
hypothalamus, which is responsible for medium-term (minutes) and long-term
(hours/days) regulation of internal organ systems. Importantly, autonomic reflexes
are dynamic, where adaptations can alter rapid homeostatic control over longer time
scales [4].
This book discusses the ANS from both an enlarged and a timed perspective.
First, it presents how the organization of the ANS is built in four different hierarchi-
cal levels. Next, it discusses how the ANS function changes in the three body con-
figurations (wakefulness, slow-­­wave sleep, and rapid eye movement, REM, sleep)
found during a 24-h cycle. Finally, the most important clinical implications for this
enlarged and timed vision of the ANS are discussed.
The Autonomic Nervous System – Basic and Clinical Aspects is designed as a
comprehensive textbook for advanced medical students and health professionals. It
primes for a detailed and complete understanding of the neuroscience behind the
ANS and a proper clinical applicability of this knowledge. ANS dysfunction and
clinical manifestations involve multiple variables, which are often undervalued in
clinical practice. However, symptoms and signs of ANS disturbances should always
be considered according to their diagnostic implication, their impact on the quality
of life of patients, and their prognostic value for life expectancy.

References
1. Saper CB. The central autonomic nervous system: conscious visceral perception and auto-
nomic pattern generation. Annu Rev Neurosci. 2002;25:433–69.
2. Benarroch EE. The central autonomic network: functional organization, dysfunction, and per-
spective. Mayo Clin Proc. 1993;68:988–1001.
3. Oakes PC, Fisahn C, Iwanaga J, DiLorenzo D, Oskouian RJ, Tubbs RS. A history of the auto-
nomic nervous system: part II: from Reil to the modern era. Childs Nerv Syst. 2016;32:2309–15.
4. McDougall SJ, Munzberg H, Derbenev AV, Zsombok A. Central control of autonomic func-
tions in health and disease. Front Neurosci. 2014;8:440.
Contents

1 The Enlarged Autonomic Nervous System��������������������������������������������    1


The Control of Homeostasis Is the Most Important
Function of the ANS����������������������������������������������������������������������������������    2
There Are Functional Similarities in the Hierarchical Organization
of the Autonomic and Somatic Motor Pathways ��������������������������������������    6
The Autonomic Posture ����������������������������������������������������������������������������   10
Basic Neuronal Organization of ANS��������������������������������������������������������   14
References��������������������������������������������������������������������������������������������������   17
2 The Timed Autonomic Nervous System������������������������������������������������   19
Biological Rhythms Are the Basis of Predictive Homeostasis������������������   20
The Sleep/Wake Cycle as the Major 24-h Rhythm������������������������������������   31
Neurophysiology of Sleep��������������������������������������������������������������������������   36
Three Different ANS Programs (“Body Configurations”)
Occur in a 24-h Day/Night Cycle��������������������������������������������������������������   45
The Meaning of Dreaming������������������������������������������������������������������������   51
The Glymphatic System and Sleep������������������������������������������������������������   52
References��������������������������������������������������������������������������������������������������   54
3 First Level: Peripheral Sympathetic
and Parasympathetic Nervous System��������������������������������������������������   57
The Organization of the ANS at a Peripheral
Level Comprises Two Neurons in a Series������������������������������������������������   58
Cholinergic and Adrenergic Neurotransmission
and Peptidergic Co-transmission Are the Basis
of the Peripheral Motor Constituents of the ANS��������������������������������������   63
Sensory Autonomic Neurons ��������������������������������������������������������������������   79
Spinal Autonomic Reflexes Have Homologies
and Differences with Spinal Motor Reflexes ��������������������������������������������   87
Urination, Defecation, and Pupillary and Sexual Responses
Are Examples of Spinal Autonomic Reflexes��������������������������������������������   89
The Enteric ANS as an Individual Entity��������������������������������������������������   95

vii
viii Contents

The Overlooked Role of the Local Autonomic Projections


in Neuroendocrine Communication����������������������������������������������������������  102
The ANS Contributes to the Maintenance of Healthy Bone Tissue����������  106
References��������������������������������������������������������������������������������������������������  109
4 Second Level: The Brainstem ������������������������������������������������������������    113
At the Brainstem, Various Complex Autonomic Responses
Are Coordinated������������������������������������������������������������������������������������    114
Monoaminergic Systems in the Brainstem Modulate
24-h Rhythms in Physiological Function����������������������������������������������    120
24-h Rhythms in Cardiovascular Control������������������������������������������������   125
24-h Rhythms in Respiratory Control������������������������������������������������������   133
The Cerebellum and the Autonomic Posture ������������������������������������������   138
24-h Rhythms in the Immune Response��������������������������������������������������   144
24-h Rhythms in Gastrointestinal Function ��������������������������������������������   160
References������������������������������������������������������������������������������������������������   171
5 Third Level: The Hypothalamus������������������������������������������������������������  175
Hypothalamic Behaviors Comprise Coordinated
Mechanisms, Including Autonomic, Neuroendocrine,
and Motivational Components ������������������������������������������������������������������  176
24-h Rhythms in Neuroendocrine Function����������������������������������������������  180
Defense Behavior as a Paradigm of Reactive Homeostasis����������������������  190
24-h Rhythms in Food Intake, Energy Storage, and Metabolism��������������  203
24-h Rhythms in Plasma Osmolality and the Intravascular Volume����������  221
24-h Rhythms in Body Temperature Control��������������������������������������������  228
Sexual and Maternal Behavior������������������������������������������������������������������  235
References��������������������������������������������������������������������������������������������������  239
6 Fourth Level: The Limbic System������������������������������������������������������    245
The Limbic System Is Essential in Emotionality,
Motivation, Learning, and Memory ������������������������������������������������������    246
The Amygdala Is the Main “Motor Nucleus” of the Limbic System����    252
Emotions Comprise Feelings and Moods, and Their Expression
in Somatic and Autonomic Behaviors ����������������������������������������������������   257
Limbic Components of the Basal Ganglia ����������������������������������������������   262
Functional Neuroimaging of ANS ����������������������������������������������������������   268
Chronotypes, 24-h Rhythms and Emotion����������������������������������������������   271
24-h Rhythms and Learning and Memory ����������������������������������������������   274
References������������������������������������������������������������������������������������������������   283
7 Clinical Implications of the Enlarged Autonomic Nervous System    287
Semiological Aspects of ANS Disorders����������������������������������������������    288
Classification of ANS Disorders������������������������������������������������������������    292
Some Clinical Autonomic Entities��������������������������������������������������������    293
Peripheral Neuropathies with Dysautonomia������������������������������������    293
α-Synucleinopathies����������������������������������������������������������������������������   296
Contents ix

Diabetes Mellitus Autonomic Dysfunction ����������������������������������������   298


Autoimmune Autonomic Ganglionopathy������������������������������������������   299
Paraneoplastic Autonomic Dysfunction����������������������������������������������   299
Amyloidotic Autonomic Failure����������������������������������������������������������   299
Autonomic Dysfunction in Primary Sleep Disorders��������������������������   300
Autonomic Dysfunction in Cardiovascular Disorders ������������������������   303
Autonomic Dysfunction Associated with Aging ��������������������������������   305
Guillain–Barré Syndrome��������������������������������������������������������������������   306
Hereditary Autonomic Neuropathies ��������������������������������������������������   306
Autonomic Disturbances in Spinal Cord Injuries��������������������������������   307
Drug-Induced Autonomic Dysfunction ����������������������������������������������   307
Autonomic Disorder in Fibromyalgia, Chronic Fatigue
Syndrome and Chronic Pain����������������������������������������������������������������   307
Hyperthermia, Hypothermia����������������������������������������������������������������   309
References������������������������������������������������������������������������������������������������   310
8 Clinical Implications of the Timed Autonomic Nervous System ����    313
Due to the “24/7 Society,” the ANS Has Lost Adequate
Experience of Day and Night����������������������������������������������������������������    314
The Disruption of the Three ANS Physiological Programs
(“Body Configurations”) Is a Major Consequence
for the “24/7 Society”����������������������������������������������������������������������������    316
Jet Lag, Shift Work, and Chronodisruption as Examples
of a Desynchronized ANS ��������������������������������������������������������������������    319
Jet Lag��������������������������������������������������������������������������������������������������   321
Shift-Work Disorder����������������������������������������������������������������������������   324
Chronodisruption ��������������������������������������������������������������������������������   327
Chronobiological Treatment of a Desynchronized ANS ������������������������   328
Some Clinical Autonomic Entities Associated
with a Desynchronized ANS��������������������������������������������������������������������   332
Metabolic Syndrome����������������������������������������������������������������������������   332
Mental Illnesses ����������������������������������������������������������������������������������   339
Brain Aging������������������������������������������������������������������������������������������   349
Cancer��������������������������������������������������������������������������������������������������   361
References������������������������������������������������������������������������������������������������   365
Epilogue��������������������������������������������������������������������������������������������������������    375
References������������������������������������������������������������������������������������������������   376
Index��������������������������������������������������������������������������������������������������������������    377
Abbreviations

3V Third ventricle
5-HT Serotonin
ACh Acetylcholine
ACTH Adrenocorticotropin
AD Alzheimer’s disease
AgRP Agouti protein-related peptide
AMPK AMP-activated protein kinase
AN Ambiguous nucleus
ANP Atrial natriuretic peptide
ANS Autonomic nervous system
AP Area postrema
APP Amyloid precursor protein
ARC Arcuate nucleus
AV3V Anterior ventral region of the third ventricle
AVP Arginine vasopressin
AVPV Anteroventral periventricular
Aβ Amyloid β
BAT Brown adipose tissue
BCRs B cell receptors
BF Basal forebrain
BMD Bone mineral density
BNP Brain natriuretic peptide
BP Blood pressure
BRAC Basic rest–activity cycle
BZD Benzodiazepine
CALB Calbindin
CALR Calretinin
CART Cocaine and amphetamine-regulated transcript
CCG Controlled clock genes
CCK Cholecystokinin
CGRP Calcitonin gene-related peptide
CNP C-type natriuretic peptide
CNS Central nervous system
CO Carbon monoxide

xi
xii Abbreviations

COMT Catechol-O-methyl transferase


CRH Corticotropin-releasing hormone
CSF Cerebrospinal fluid
CVC Cutaneous vasoconstriction
DA Dopamine
DDR DNA damage response
DLMO Dim light melatonin onset
DMH Dorsomedial hypothalamus
DRN Dorsal raphe nucleus
DSM Diagnostic and Statistical Manual of Mental Disorders
E Epinephrine
ECL Enterochromaffin-like cells
ECN External carotid nerve
EEG Electroencephalographic
EGFR Epidermal growth factor receptor
EKG Electrocardiogram
EMA European Medicines Agency
EPSP Excitatory postsynaptic potential
FDA Food and Drug Administration
FEO Food-entrained oscillators
fMRI Functional magnetic resonance imaging
FSH Follicle-stimulating hormone
GABA γ-Aminobutyric acid
GH Growth hormone
GHRH Growth hormone-releasing hormone
GLP-1 Glucagon-like peptide-1
Glu Glutamate
Gly Glycine
GnRH Gonadotropin-releasing hormone
GRP Gastrin-releasing peptide
His Histamine
HRV Heart rate variability
ICN Internal carotid nerve
ICSD International Classification of Sleep Disorders
IFN Interferon
Ig Immunoglobulin
IGF1 Insulin-like growth factor 1
IL Interleukin
IPSP Inhibitory postsynaptic potential
IRS Insulin R substrate
IRS-1 Insulin receptor substrate 1
Kiss1 R Kisspeptin receptor
Kiss1 Kisspeptin
LAN Light at night
LC Locus coeruleus
Abbreviations xiii

l-Dopa l-Dihydroxyphenylalanine
LDT Laterodorsal tegmental nucleus
LH Luteinizing hormone
LHA Lateral hypothalamic area
LPB Lateral parabrachial nucleus
MAO Monoamine oxidase
MCH Melanin concentrating hormone
MCI Mild cognitive impairment
MMC Migrating motor complex
MnPO Median preoptic area
MS Metabolic syndrome
MSH Melanocyte-stimulating hormone
MT Melatonin receptor
NE Norepinephrine
NF Nuclear factor
NK Natural killer
NMDA N-methyl-d-aspartate
NO Nitric oxide
NOS NO synthase
NPY Neuropeptide Y
NREM Non-REM
NTS Nucleus tractus solitarius
OC Optic chiasm
O-GlcN-Ac O-β-d-N-acetylglucosamine
OVLT Organum vasculosum laminae terminalis
PACAP Pituitary adenylate cyclase-activating peptide
PARP Poly ADP-ribose polymerase
PB/PC Parabrachial/precoeruleus
PBN Parabrachial nucleus
PET Positron emission tomography
PG Prostaglandin
PGO Ponto-geniculo-occipital
PH Posterior hypothalamus
POMC Pro-opiomelanocortin
PPARs Peroxisome proliferator-activated receptors
PPT Pedunculopontine tegmentum nucleus
PPT/LDT Pedunculopontine and laterodorsal tegmenti
PRL Prolactin
PSG Polysomnography
PTH Parathyroid hormone
PVH Paraventricular hypothalamus
PVN Paraventricular nucleus
PZ Parafacial zone
REM Rapid eye movement
RHT Retino-hypothalamic tract
xiv Abbreviations

RNS Reactive nitrogen species


ROS Reactive oxygen species
rRPa Rostral raphe pallidus
SCG Superior cervical ganglion
SCGx Superior cervical ganglionectomy
SCN Suprachiasmatic nuclei
SFO Subfornical organ
SIF Small, intensely fluorescent
SLD Sublaterodorsal
SN Substantia nigra
SON Supraoptic nucleus
sPVz Sub PVN zone
TCRs T cell receptors
Th T helper
TMN Tuberomammillary nucleus
TNF Tumor necrosis factor
T-reg T regulatory cells
TRH Thyrotropin-releasing hormone
TRP Transient receptor potential
TSH Thyroid-stimulating hormone
VIP Vasoactive intestinal peptide
VLPO Ventrolateral preoptic nucleus
VMH Ventromedial hypothalamus
VMN Ventromedial nucleus
VTA Ventral tegmental area
WAT White adipose tissue
WHO World Health Organization
The Enlarged Autonomic Nervous
System 1

Abstract
The traditional view of the autonomic nervous system (ANS) considers only its
peripheral part, the sympathetic and parasympathetic systems, to which some-
times the enteric ANS is added as an independent entity. However, this view is
insufficient to understand the most important function of the ANS: the mainte-
nance of homeostasis. This Chapter describes the hierarchical organization of the
autonomic motor pathways and their similarities with the somatic motor organi-
zation. It analyzes the types of neurons, neuronal circuits, and electric potentials
identified in autonomic neurons.

Keywords
Acetylcholine • Autonomic posture • Basic neuronal circuits • Circadian clock
• Fight or flight reaction • Golgi I and Golgi II neurons • Hierarchical organiza-
tion of the ANS • Homeostasis • Norepinephrine • Somatic and visceral motor
neurons

Objectives
After studying this chapter, you should be able to:

• Describe the control of reactive and predictive homeostasis as the most


important function of the ANS
• Describe the hierarchical organization of the autonomic and somatic motor
pathways and their similarities
• Name the components of the autonomic posture in comparison with the
mechanisms of body posture
• Describe the two types of neurons and the three neuronal circuits that give
rise to the basic neuronal organization of ANS
• Name the different types of electrical potentials identified in autonomic
neurons

© Springer International Publishing AG 2018 1


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1_1
2 1  The Enlarged Autonomic Nervous System

 he Control of Homeostasis Is the Most Important Function


T
of the ANS

The traditional view of the autonomic nervous system (ANS) considers only its
peripheral part, the sympathetic and parasympathetic systems, to which sometimes
the enteric ANS is added as an independent entity. However, this view is insufficient
to understand the most important function of the ANS: the maintenance of
homeostasis.
The term “homeostasis” was coined by the distinguished American physiologist
Walter B. Cannon as a comprehensive concept to describe the physiological factors
that maintain the equilibrium state of the body, and therefore, life [1]. Following
Claude Bernard, Cannon perfected the idea of the constancy of the internal environ-
ment, giving it its present meaning. As explained by Cannon, he chose as a prefix
homeo- (“like”) instead of homo- (“the same”) to take into account the normal vari-
ations that any physiological variable exhibits, thus avoiding the misleading consid-
eration of a “constant (fixed) internal milieu.”
The development of chronobiology as an independent discipline added another
aspect to homeostasis. This term is used today to define not only the strategies that
allow the organism’s proper response to changes in the environment (“reactive
homeostasis”), but also the time-related mechanisms, remarkably developed, that
allow the body to predict the occurrence of a given environmental stimulus to initi-
ate the appropriate corrective responses beforehand (“predictive homeostasis”) [2].
Predictive homeostasis comprises the anticipatory mechanisms that precede a regu-
larly occurring environmental phenomenon, thus facilitating a better physiological
adaptation to it. For example, the increase in plasma cortisol that precedes waking
anticipates the energy demands of a changing posture; the increased gastrointestinal
secretion preceding the usual lunch time anticipates the changes in the content of
the digestive tract.
Let us suppose that a diurnally active mammal in search of food finds the food at
a place about 2 h from the shelter it uses to escape its predator at night. This situa-
tion makes it essential for the animal to predict nightfall about 2 h in advance. It do
this using the position of the sun or other environmental variables, such as tempera-
ture or humidity. However, the existence of clouds or a large daily variation in ambi-
ent temperature make the value of such reference parameters unreliable. It is then
extremely useful for survival to possess a system of time control integrated into its
own organism that allows a temporal prediction without having to depend on the
reading of external signals. This was probably the strategy selected as an advantage
by evolution.
The circadian clock is ideal to fulfill such a function: we could have a sufficiently
accurate idea of the time of day by simply analyzing our biological structure peri-
odically and without consulting our wristwatch. That is, a “day” and “night” have
been created within the organism to allow it to optimize adaptation [3].
Generally, the nervous system can produce a small number of actions: (a) the
contraction of a muscle or of a group of striated or smooth muscles; (b) the exo-
crine, endocrine, or paracrine secretion of a cell or cell group; (c) changes in cellular
The Control of Homeostasis Is the Most Important Function of the ANS 3

permeability. This is achieved via motor neurons, which can be classified per their
targets into three categories:

• Somatic motor neurons, which originate in the central nervous system (CNS),
project their axons to the skeletal muscles (such as the muscles of the limbs,
abdominal, and intercostal muscles) that are involved in locomotion.
• Special visceral motor neurons, also called branchial motor neurons, which
directly innervate the branchial muscles (that motorize the gills in fish and the
face and neck in land vertebrates).
• Visceral preganglionic neurons, which indirectly innervate the heart, the smooth
muscle in all organs, the abdominal viscera, the exocrine and endocrine glands,
and the immune system. They synapse onto neurons located in the autonomic
ganglia of the ANS (sympathetic and parasympathetic postganglionic neurons).

As a consequence: (a) the motor command of skeletal and branchial muscles is


monosynaptic (involving only one motor neuron, somatic and branchial respec-
tively, which synapses onto the muscle); (b) the command of the viscera is disynap-
tic, involving two neurons, the visceral preganglionic neuron located in the CNS,
which synapses onto a postganglionic neuron, which synapses onto the organ.
The ANS, which is in charge of the innervation of smooth muscles of all organs
and innervation of the viscera, immune tissue, and exocrine and endocrine glands,
can produce the three actions mentioned above (smooth muscle contraction; exo-
crine, endocrine, or paracrine secretion; permeability changes). It must be noted that
for the command of visceral muscles, the ganglionic neuron, parasympathetic or
sympathetic, is the real motor neuron, as it is the one that directly innervates the
muscle, whereas the general visceral motor neuron is, strictly speaking, the pregan-
glionic one [4].
All vertebrate motor neurons are cholinergic, that is, they release the neurotrans-
mitter acetylcholine (ACh). Parasympathetic ganglionic neurons are also choliner-
gic, whereas most sympathetic ganglionic neurons are noradrenergic, that is, they
release the neurotransmitter norepinephrine (NE).
It is not unusual to hear or read or to remain implicit that actions of the sympa-
thetic and parasympathetic divisions promote opposite effects in the organs they
innervate. However, for the precise aim of a homeostatic response requiring effec-
tiveness, energy economy, and sometimes a fast response, the sympathetic and para-
sympathetic neurons must collaborate for the final net effect.
In general, the sympathetic activity is designed to place the individual in a situa-
tion of defense in the face of danger, real or potential. Sympathetic stimulation leads
to variations in visceral functions designed to protect the integrity of the organism
and to ensure survival. In fact, a sympathectomized animal hardly survives if left
free and exposed in its natural environment. However, in addition to calm and rest-
ing states, our daily life involves many perturbations that induce active conditions
such as locomotion, eating, and communication. Thus, the mobilization of physio-
logical variables does not occur in an all-or-nothing manner and is not exclusive of
extreme “fight or flight” conditions; rather, they can occur with graded intensities in
4 1  The Enlarged Autonomic Nervous System

ordinary daily situations. During such active periods, cardiovascular, respiratory,


and body temperature regulation needs to be adjusted to situational demands, which
differ from those during resting states, by modulating or resetting homeostatic
points. For instance, if an increase in blood pressure (BP) is needed, a greater gain
in the response can be obtained by the concomitant enhancement of sympathetic
activity and inhibition of parasympathetic activity. This is obtained not only in the
CNS areas, but also via reciprocal inhibitory connections at the level of the heart
wall [5].
Autonomic nervous system activity adapts to body changes and supports somatic
reactions. The autonomic response may be secondary to the somatic, parallel to, or,
most commonly, before (the “autonomic posture”) [3]. The ANS is designed to
produce sustained actions, compared with those of the somatic nervous system. On
the other hand, the ANS performs this multisystem control in a continuous and
constant manner throughout the life cycles, both at rest and during activity in the
three body configurations within a 24 h cycle discussed in Chap. 2.
One of the most surprising features of the ANS is the speed and intensity with
which it modifies the visceral functions. In a few seconds, for example, the heart
rate and BP can increase up to double the normal rates, there may be intense sweat-
ing, the urinary bladder may empty, or digestive motility and secretion may be
activated.
The sympathetic stimulation causes general responses in the organism for two
reasons. First, there is a high degree of irradiation of the connections between pre-
ganglionic and postganglionic neurons. The proportion of preganglionic fibers to
postganglionic neurons in sympathetic ganglia is 1/4–1/30, and since the degree of
axonal branching is high, it is estimated that each preganglionic axon contacts an
average of 120 neurons. For this reason, the activity of a few preganglionic neurons
is highly amplified numerically via their ganglionic connections. Second, sympa-
thetic activity produces a stimulation of the secretion of catecholamines from the
adrenal medulla. In circulation, almost all epinephrine (E) found is derived from the
adrenal medulla, but only 20% of the NE is. The rest of the circulating NE comes
from the peripheral sympathetic terminals, i.e., as the neurotransmitter that has
escaped presynaptic reuptake.
A massive activation of the sympathetic system produces a set of reactions
defined as an alarm response (fight or flight) [6]. The most obvious visceral phe-
nomena of this massive activation are:

• Pupil dilation, to increase the visual field


• Piloerection, to simulate a larger body size
• Sweating, to lose heat that is produced by muscular activity
• Increased cardiac activity and BP, to provide greater blood flow to the muscles
• Bronchodilation, to increase the entrance of air into the lungs
• Hyperglycemia
• Inhibition of the digestive function
• Inhibition of the urinary and genital functions
The Control of Homeostasis Is the Most Important Function of the ANS 5

In contrast, the activity of the parasympathetic system is related to protective and


conservation functions, which favor the proper functioning of the different visceral
organs. The functional components of the parasympathetic system do not act simul-
taneously under normal conditions, but participate in specific reflexes or in inte-
grated reactions to promote a specific visceral function. Thus, stimulation of
different nuclei of parasympathetic neurons promotes responses such as:

• Pupillary constriction, to protect the retina from excessive illumination


• Decreased heart rate, to avoid excessive activity
• Bronchoconstriction, to protect the lungs
• Increased motility and digestive secretions, to promote digestion
• Urinary activity and urination
• Genital activity (erection)

Therefore, the parasympathetic effects are, in consonance with their fundamental


purpose, more localized. In the parasympathetic ganglia, each preganglionic neuron
contacts a few postganglionary neurons, whose divergence is much smaller than
that of the sympathetic system [6].
Both efferent systems exert a control of the function of the visceral organs in vari-
able forms. Some effector organs receive innervation from a single system. For
example, the smooth muscle of most blood vessels is only controlled by sympathetic
vasoconstrictor innervation (except for vessels of genital organs, which also receive
parasympathetic innervation, and cholinergic vasodilator fibers in the skeletal mus-
cles and brain). In these cases, control of the visceral activity depends on the varia-
tions of the frequency of discharge of the impulses by sympathetic innervation. As
the autonomic nerves have a basal tonic activity (of low pulse frequencies, 0.5–5 Hz),
the activity may increase or decrease. In most vessels, an increase in sympathetic
frequency determines vasoconstriction and a decrease, vasodilation [7].
Sympathetic and parasympathetic nerve fibers simultaneously innervate many
effector organs. The activity of the organ depends on the interaction or balance
between the signals of both systems, which exert antagonistic effects. This interac-
tion can be developed by opposite actions on the same effector cells, as in the heart,
where the sympathetic excites the nodal cells thus increasing the heart rate, whereas
the parasympathetic inhibits the same cells, thus reducing their frequency.
Another possibility is the action on different cells, with opposite effects. In the
iris, the sympathetic excites the meridian muscular fibers, which produces mydria-
sis, whereas the parasympathetic excites the circular fibers and produces miosis. In
these cases, the functional balance is relatively complex. In normal circumstances,
the two systems are reciprocally active, because the central and reflex signals excite
one system and inhibit the other. When both sympathetic and parasympathetic sys-
tems innervate the same target cells, more complex interactions arise. For example,
the simultaneous activation of sympathetic fibers may exaggerate the cardiovascular
response to parasympathetic activity and vice versa. This phenomenon is mediated
by mutual influences at the presynaptic and postsynaptic domains [7].
6 1  The Enlarged Autonomic Nervous System

 here Are Functional Similarities in the Hierarchical


T
Organization of the Autonomic and Somatic Motor Pathways

To understand the hierarchical organization of the ANS, several concepts derived


from the somatic motor system are useful [3]. Any movement, even the simplest
one, involves an enormous amount of information processing and the participation
of numerous neuronal groups. On the other hand, each movement of our body is
based on a posture; thus, it is important to consider how the motor system provides
global responses for both components, the posture needed, and the movement itself.
To achieve this, there are four hierarchical levels in which the somatic motor
system is organized (Fig. 1.1): (a) spinal cord; (b) brainstem; (c) motor cerebral
cortex; (d) premotor cortical areas.
The circuits of the three basic motor reflexes, i.e., the myotatic reflex, the tendon
reflex, and the withdrawal reflex, are found in the spinal cord, the centers of regula-
tion of the dorsolateral and ventromedial motor neuron systems are located in the
brainstem, and the motor programs defined by the secondary motor areas (premotor,
motor supplemental area, parietal cortex) are situated in the primary motor cortex.

Premotor cortex
Cerebellum Supplementary Basal ganglia
motor cortex

Primary
motor cortex

Brain stem

Locomotor
Spinal cord
System

Fig. 1.1  Hierarchical organization of the somatic motor system. Modified with permission from
Cardinali [3]
Hierarchical Organization of Motor Pathways 7

THE AUTONOMIC NERVOUS Level 4:


Limbic System.
SYSTEM ENLARGED Emotions, social behavior,
learning, memory

Cerebellum Basal
Level 3: Limbic System ganglia
Hypothalamus.
(Autonomic + neuroendocrine +
behavioral components)
• Defense behavior Hypothalamus
• Thermoregulation
• Feeding behavior
• Sexual behavior
• Maternal behavior Brain Stem Parasympathetic
• Biologic rhythm control

Level 2: Spinal Cord Sympathetic


Brain Stem.
Complex autonomic reflexes
(respiratory, cardiovascular,
urination, defecation)

Level 1:
Spinal Cord.
Metameric autonomic
reflexes (viscero-visceral,
viscero-cutaneous)

Fig. 1.2  Hierarchical organization of the autonomic nervous system (ANS). The figure was pre-
pared in part using image vectors from Servier Medical Art (www.servier.com), licensed under the
Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

Also, two other brain areas, the cerebellum and the basal ganglia, have an essential
regulatory function with regard to the somatic motor system [3].
In the ANS, a similar arrangement can be found (Fig. 1.2). The four major levels
are: (a) spinal cord; (b) brainstem; (c) hypothalamus; (d) limbic system.
The spinal cord contains connections that mediate segmental autonomic reflexes
involved in visceral function. These localized autonomic phenomena acquire inter-
segmental significance in the brainstem. In the brainstem, many complex autonomic
responses, such as cardiovascular and respiratory regulation, are found.
The other two hierarchical levels of the ANS organization are the hypothalamus
and the limbic system. In the hypothalamus, autonomic motor programs acquire
their homeostatic nature. The cardiocirculatory and respiratory response to hemor-
rhage is completed by a neuroendocrine response, i.e., the secretion of arginine
vasopressin (AVP) and adrenocorticotropin (ACTH), and by a behavioral response
(thirst and fluid intake, etc.). These neuroendocrine mechanisms have strong simi-
larities to those of the somatic posture needed for the execution of orders derived
from the top level of the motor system.
The limbic system gives the homeostatic reaction its emotional tone and social
significance [8]. The amygdala provides affective or emotional value to incoming
sensory information and has multiple downstream targets that participate in the
autonomic and neuroendocrine–immune responses. The anterior cingulate cortex is
8 1  The Enlarged Autonomic Nervous System

interconnected with the anterior insula and is subdivided into the ventral and dorsal
regions. The insular cortex is the primary interoceptive cortex that integrates vis-
ceral, pain, and temperature sensations. This hierarchical level also comprises the
cerebellum, which participates in the coordination of the executed autonomic pro-
grams, and the basal ganglia, which are relevant for the selection of the autonomic
program best adapted to a given situation [9].
It is useful to compare the performance of somatic and autonomic motor
responses with that of a building under construction [3]. In this case, we can observe
three basic categories, in hierarchical order: masons, foremen (overseers), and
architects. Schematically, the architects are responsible for planning (activity before
the start of the work), the foremen for the management and coordination, and the
masons for the construction itself. All three functions are indispensable for building
and the failure of one of them will affect the success of the whole work.
The main “architects” of the ANS, which are responsible for the layout of the
autonomic programs, are in cortical and subcortical regions of the limbic system.
They include the limbic association cortex and the subcortical areas such as the
amygdala, hippocampus, septal nuclei, olfactory bulb, and portions of the basal
ganglia (ventral striatum, nucleus accumbens). The ventral portion of the basal gan-
glia are part of a loop that begins and ends in the limbic areas. The limbic associa-
tion cortex projects to ventral striatum (nucleus accumbens) then to the thalamus,
and finally back to the limbic cortex. The overall function of this loop is to select a
sequence of autonomic actions (behaviors) while suppressing others.
The “overseers” or areas of execution in the motor system are linked to the move-
ment itself (dorsolateral system: primary motor cortex, red nucleus) and the posture
(ventromedial system: vestibular nuclei, inferior colliculus superior, reticular forma-
tion) needed for that movement. In the case of the ANS, the hypothalamus plays such
a function. The main autonomic behaviors coordinated by the hypothalamus are:

• Defense behavior
• Nutritive or appetitive behavior
• Thermoregulatory behavior
• Maternal and sexual behavior

Typically, autonomic behaviors involve the coordinated expression of auto-


nomic, neuroendocrine, somatic, and motivational mechanisms. Recent observa-
tions underlined the role of the cerebellum in the appropriate coordination of these
components [4].
The “masons” of the somatic motor system, which represent the final common
pathway of the system and are directly responsible for muscle contraction, are: (a) the
motor units of the spinal cord; (b) the motor units in the nuclei of the cranial nerves.
The “masons” of the ANS include the motor neurons (postganglionic) of the sym-
pathetic, parasympathetic, and enteric systems. These neurons are separate functional
units (vasomotor muscular, cutaneous vasomotor, sudomotor, pilomotor, visceromo-
tor) that exercise a specific and appropriate control over a target organ or cell.
From data derived from the somatic motor system, there are three important
aspects to be considered for the physiological understanding of the hierarchical
organization described above: (a) there is somatotopy, i.e., an orderly anatomical
Hierarchical Organization of Motor Pathways 9

map at each level of organization; (b) at each level of the motor system, information
from the periphery is received and modifies the descending order of command; (c)
the upper levels have the capacity to control or suppress the information that reaches
them (afferent control).
For other authors, the hierarchical organization of the integrated control of auto-
nomic responses involved in homeostasis, adaptation, and emotional and goal-­
oriented behaviors includes the following levels: (a) spinal; (b) bulbopontine; (c)
pontomesencephalic; (d) forebrain [10]. The bulbopontine (lower brainstem) level
is involved in the reflex control of circulation, respiration, gastrointestinal function,
and micturition. The pontomesencephalic (upper brainstem) level integrates auto-
nomic control with pain modulation and responses to stress. The brainstem auto-
nomic areas include the periaqueductal gray, the parabrachial nucleus (PBN), the
nucleus tractus solitarius (NTS), the ventrolateral medulla, and the medullary raphe.
The forebrain regions involved in the control of autonomic functions include the
insular cortex, the anterior cingulate cortex, the amygdala, and the hypothalamus.
However, such a view considers only partially the role of the hypothalamus as a
central component in the ANS hierarchy. Maintaining the hypothalamus as an inde-
pendent level has the advantage of recognizing its role in integrated autonomic,
neuroendocrine–immune, and behavioral responses. Therefore, the areas of the cen-
tral autonomic network: (a) are reciprocally interconnected; (b) receive converging
visceral and somatosensory information; (c) generate stimulus-specific patterns of
autonomic, endocrine, and motor responses; (d) are regulated according to the
respective body configuration: wakefulness, slow wave or non-REM (NREM) sleep,
REM sleep [3].

Central Body Postural


command movements disturbance

Postural
adjustment
Feed-forward Feedback
(for expected (for unexpected
postural changes) postural changes)

PREDICTIVE HOMEOSTASIS REACTIVE HOMEOSTASIS

Fig. 1.3  Body posture is the position of the trunk relative to that of the limbs, and both, as a whole,
in space. The postural reflexes are a set of antigravity reflexes, articulating with each other as a
program. The postural adjustment includes anticipatory feed-forward programs and compensatory
servo-assisted, feed-back mechanisms. In the same way, there is an “autonomic posture” that
includes the mechanisms of anticipatory predictive homeostasis and the corrective mechanisms of
reactive homeostasis. Modified with permission from Cardinali [3]
10 1  The Enlarged Autonomic Nervous System

The Autonomic Posture

The adjustment program of body posture includes compensatory and anticipatory


mechanisms. The spinal motoneurons are under the continuous influence of descend-
ing impulses from the upper regions and from the corresponding muscular and skin
territories. One of the fundamental descending programs regulating the activity of
motoneurons is that of posture, derived from neurons located in the brainstem [11].
The term “posture” defines the position on the trunk and the limbs. The postural
reflexes are a set of antigravity reflexes, articulating each other as a program. This
postural adjustment program includes feed-back and feed-forward compensatory
mechanisms (Fig. 1.3) In the same way, there is an “autonomic posture” that
includes the anticipatory mechanisms of predictive homeostasis and the corrective
mechanisms of reactive homeostasis.
In the case of body posture, we anticipate with a proper body position the pre-
dictable changes given by muscle activity and the force of gravity, and correct this
appropriate position with compensatory changes triggered by sensory information
(Fig. 1.4). The position of the body in space varies accordingly to the movements
performed. Therefore, we do not have a “single posture,” but the correct one that is
adapted to the movements performed. This maintenance of the postural equilibrium
requires, in addition to the movements, advanced programming, and an on-line reg-
ulation of the process to adapt to the changes. For this, the integration of four senso-
rial modalities is indispensable: (a) vision; (b) position of the head; (c) proprioception;

a b

Popliteus
Biceps EMG Quadriceps
Gastrocnemius

Tibialis anterior
0 100 200 ms
0 100 300 500
ms
Ring
Gastrocnemius EMG
Popliteus
Quadriceps
Gastrocnemius
0 100 300 500
ms
Tibialis anterior
Ring
0 100 200 ms

Fig. 1.4  With any motor plan, the proper posture program is executed first. Note that postural responses
are always triggered before voluntary movements (a). If a normal individual is placed on a forward-
leaning platform, the extension of the lower limb stabilizes the body, which causes the extensor (anti-
gravitational) reflex to increase in the lower limbs as it is practiced (b). If the platform is tilted backward,
the extension of the lower members, antigravitational in the previous case, now produces destabiliza-
tion. In this second case, and after a few repetitions of the test, the extensor reflex diminishes, until it is
completely extinguished. There is strong evidence for the role of the cerebellum in the mechanisms of
both body posture and autonomic posture. Modified with permission from Cardinali [3]
The Autonomic Posture 11

(d) exteroception (touch). Based on these data, the nervous system produces an
early postural program suitable for movement and provides a series of automatic
adjustments if unanticipated problems occur (Fig. 1.4) [12].
In the case of the autonomic posture, and as discussed in Chap. 2, the circadian
system generates a map of acrophases (maximal neurovegetative functions con-
trolled by the ANS) that allows the adequate neuroendocrine–immune configuration
for each of the three autonomic configurations of the body to be anticipated in a
24-h cycle (wakefulness, NREM sleep, REM sleep) [3]. In the face of unexpected
demands, the modification of the predetermined neuroendocrine–immune configu-
ration and the readjustment of the autonomic function occur (Fig. 1.5).

Homeostasis

Homeostatic mechanisms

PREDICTIVE REACTIVE

Autonomic
posture

Fig. 1.5  In the case of autonomic posture, the circadian system generates a map of acrophases (max-
imum neurovegetative functions controlled by the ANS), which allows the adequate neuroendocrine–
immune configuration for each autonomic configurations of the body (wakefulness, slow wave sleep,
REM sleep) (predictive homeostasis) to be anticipated. Based on interoception, when unexpected
demands arise, the modification of the predetermined neuroendocrine–immune configuration and the
readjustment of the autonomic function arise (reactive homeostasis). The figure was prepared in part
using image vectors from Servier Medical Art (www.servier.com), licensed under the Creative
Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)
12 1  The Enlarged Autonomic Nervous System

GEOPHYSICAL SOCIAL AND


ENVIRONMENT PSYCHOLOGICAL STIMULI
Light Abstract thinking
Noise Emotions
Odors Stress
Electromagnetic field

NEURO-
AUTONOMIC ENDOCRINE
NERVOUS SYSTEM
SYSTEM Hypothalamus
Hypophysis
Pineal gland Immune signals
Neural and endocrine signals

IMMUNE SYSTEM
Bone marrow
Thymus
Spleen
EXTERNAL
Lymph nodes
PATHOGENIC
GIimmune tissue
STIMULI INTERNAL
Bronchial immune tissue
Bacteria PATHOGENIC
Tonsils
Virus STIMULI
Parasites Inflammation
Autoantigens
Tumor cells

Fig. 1.6  Basis of the autonomic posture. The way in which the nervous system communicates
with the immune system is twofold: (a) through the neuroendocrine system (hypothalamic–pitu-
itary axis and pineal gland), via the secretion of pituitary, adrenocortical, thyroid and gonadal
hormones and melatonin, thus modulating the immune response; (b) through the ANS, in both the
parasympathetic and sympathetic divisions, which supplies the lymph nodes, thymus, spleen, and
bone marrow. Various groups of hypothalamic neurons react to humoral signals (cytokines) pro-
duced by immunocompetent cells

The neuroendocrine–immune mechanisms involved in the “autonomic pos-


ture” are summarized in Fig. 1.6. The link between the activity of the nervous and
immune systems has been the subject of numerous investigations in the last
50 years, giving rise to psychoneuroimmunoendocrinology as a discipline [13].
Multidirectional interactions among the immune, endocrine, and nervous systems
have been demonstrated in humans and nonhuman animal models. Neuroendocrine–
immune interactions can be conceptualized using a series of feedback loops,
which culminate in distinct neuroendocrine–immune phenotypes. Behavior can
exert profound influences on these phenotypes, which in turn reciprocally modu-
late behavior [13].
The way in which the nervous system communicates with the immune system is
twofold: (a) through the neuroendocrine apparatus (hypothalamic–pituitary axis and
pineal gland), via the secretion of pituitary, adrenocortical, thyroid, and gonadal hor-
mones, and melatonin, all of which have a modulatory effect on the immune response;
(b) through the ANS, both the sympathetic and parasympathetic divisions, which
innervates the lymph nodes, thymus, spleen, and bone marrow (Fig. 1.6). Both path-
ways carry the link among the limbic, motivational, and immune response. Galen
(second century AD), in his writings “On tumors against nature,” had sensed this
association, stating that breast cancer appeared in women whose menstruation was
either abnormal or nonexistent because of the accumulation of “waste of black bile”
The Autonomic Posture 13

(melancholia). A depressed patient is prone to developing inadequate immune


responses; in contrast, a normal emotional balance contributes to a normal immune
defense.
On the other hand, and because of the immune reaction, important changes in neu-
ronal activity are verifiable. Several groups of central neurons react to humoral signals
produced by immunocompetent cells (cytokines), such as interleukin (IL) 1 and 6,
tumor necrosis factor-α (TNF-α), or interferon-γ (IFN-γ) [13]. These cytokines give
rise to signs and symptoms that accompany acute or chronic infection (loss of appetite
or anorexia, depressed motor activity, loss of interest in daily activities), in addition to
activating the adrenal pituitary axis and producing thermogenesis (disease behavior).
The ambiome is defined as the set of nongenetic, changing elements that surround
the individual and that contribute to the development and building of the human being,
and therefore the state of health or the appearance of disease. It is part of the biopsycho-
social–ecological reality of the individual from which the microbiome has been extracted
as being very important in recent years. The microbiome defines the set of microorgan-
isms that are normally located in different places in the human body, in particular the
digestive tract [14]. The microbiome is in a commensal symbiotic relationship with the
host. These microbial components aid in the digestion of food, produce vitamins, and
protect against the colonization of other microorganisms that may be pathogenic. There
are few physiological and immunological parameters that are not deeply affected by the
presence and nature of the microbiome, with host resistance to infections being one of
the most prominent factors. The gut microbiome is highly dynamic, exhibiting daily
cyclic fluctuations that have repercussions for the host metabolism and provide evidence
for the cross-­regulation of prokaryotic and eukaryotic circadian rhythms [14]. We could
see the microbiome forming part of our internal environment.
Genetic factors explain only some (<30%) of the changes linked to health and
disease, as revealed by studies in twins. The ambiome and the microbiome corre-
spond to the rest of the changes, which are essentially epigenetic. The biopsychoso-
cial–ecological nature of the individual changes in the three autonomic body
configurations, wakefulness, slow-wave sleep, and REM sleep is discussed in Chap. 2.
Therefore, the autonomic posture is an essential and prior program for the homeo-
static responses of organs and systems [15].
Hence, rather than being a mere top–down or reflex regulation, signals from the
organs influence the functioning of the brain. For example, the reflex regulation of BP
and the heart rate is not only subject to modulation by ascending information from the
body, but also by descending information from several areas in the hypothalamus and
cortex. The CNS has the capacity to control its output via the ANS using an amazing
differentiation. For example, not only do the biological clock and prefrontal cortex
contain neurons that influence the parasympathetic or sympathetic motor neurons,
they also contain different neurons that project to diverse body compartments.
In the end, this leads to integrated responses whereby visceral sensory information
reaches higher centers in the CNS via vagal or spinal sensory pathways, causing a reac-
tion that considers factors such as the time of day, the season, the reproductive status,
or the mood. Based on all this information, the brain sets the balance of the different
parts of the ANS, causing its output to change its emphasis as per the situation. A dis-
turbed balance, either as a result of behavior or of disease of any of the organs, leads to
pathological conditions affecting the functioning of the entire individual [15].
14 1  The Enlarged Autonomic Nervous System

Basic Neuronal Organization of ANS

Although dozens of types of neurons have been described in the nervous system by
their morphological characteristics, when the length of the axon (indicative of the func-
tion they play) is considered, only two types of neurons are distinguished (Fig. 1.7) [3]:

• Golgi type I neurons with an identifiable axon, which are involved in the transfer of
information between brain regions or in providing a basal tone of excitation to wide-
spread brain areas. The difference between the two subsets of Golgi I neurons is the
degree of axonal branching. In projection neurons, ramifications are limited to one or
a few brain areas, whereas in widely distributed neurons (“spider web” neurons), axo-
nal arborization ends in many areas or in some cases most of the cerebral cortex.
• Golgi type II neurons, with no identifiable or poor developed axon, that fulfill the
function of interneurons in local circuits.

These two neuronal types (Golgi I and Golgi II) generate the three basic circuits:

• Local circuits, consisting of interneurons.


• Projection circuits or “point to point” connections, which relate to distant local
circuits between them.

a b
Dendrite

Axon

Dendrite

Axon

Fig. 1.7  Golgi type I (a) and Golgi type II neurons (b). Modified with permission from Cardinali [3]
Basic Neuronal Organization of ANS 15

• “Spider web” circuits, by which local modifications of brainstem nuclei are


transformed into global states of CNS, e.g., wakefulness or sleep. These circuits
are of fundamental importance for understanding the homeostatic function of the
ANS.

Synaptic potentials are how a neuron can modify the membrane potential of the
cells with which it connects. For this, the presynaptic neuron releases a chemical
transmitter or, less frequently, the transmission is performed by an electrical mecha-
nism [11].
In chemical transmission, the neurotransmitter interacts with receptors on the
surface of the postsynaptic membrane resulting in the generation of synaptic poten-
tials, which may be inhibitory – inhibitory postsynaptic potential (IPSP; i.e., of a
hyperpolarizing nature) – or excitatory – excitatory postsynaptic potential (EPSP;
i.e., of a depolarizing nature). The duration of synaptic potentials varies from a few
milliseconds to, in some cases, seconds, or minutes (Fig. 1.8). These potentials are
local and summable.
The integration signal (Fig. 1.8) is observed in the “trigger area” of the neuronal
membrane, where various local potentials, propagated electrotonically, are summed,
giving rise to the action potential. Generally, but not always, the “trigger zone” is in
the axonal cone. This area is characterized by a high concentration of Na+- and K+-
dependent voltage channels and constitutes the lowest threshold portion of the entire
cell membrane. If the sum of the synaptic potentials reaches the threshold, an action
potential is generated; hence, the signal produced is called “integrative” [11].
The driving signal is the action potential (Fig. 1.8). Whereas receptor synaptic
potentials only passively propagate with sharp decreases in amplitude as a function

Sensory Motor
Interneuron
neuron neuron
Graded
Input potentials
signal EPSP
IPSP

Integrative
signal

Action
Conduction potential
signal

Secretory
potential
Output
signal
Effector cell

Fig. 1.8  The different signals of reception, integration, conduction, and secretion in neurons (left).
The different electric potentials found in each segment (right). Modified from Cardinali [3]
16 1  The Enlarged Autonomic Nervous System

of distance, the action potential (or “spike potential”) has the following properties:
(a) it actively propagates along the axons or in certain cases, as the pyramidal neu-
rons of the cerebral cortex, also in dendrites; (b) it does not diminish its intensity as
a function of distance; (c) it is of an “all or nothing” nature; (d) it is similar in all
neurons, regardless of the neuron’s function. The action potential amplitude is
approximately 100 mV and the duration potential is 0.5–2 ms.
Although the Na+-dependent action potential is the fastest method of signal con-
duction in the CNS, in the dendrites of the central neurons there are Ca2+ voltage
channels displaying most of the properties of the Na+ action potential. The main
difference is the amplitude, i.e., a few mV for Ca2+ action potential vs 100 mV of
Na+ action potential.
The output signal (Fig. 1.8) is observed in synaptic axon terminals, where depo-
larization causes the release of neurotransmitter (chemical type synapses) or dis-
turbs the resting potential of the postsynaptic neuron (electric type synapses) owing
to the apposition of the membranes. In the case of chemical synapses, transmitter
release depends on the entry of Ca2+ and involves the generation of a secretory
potential [11].
Based on their conduction velocity (Fig. 1.9), nerve fibers are generally classi-
fied into:

• A fibers, myelinated, 2–20 mm thick with a velocity of 15–120 m/s. They are the
sensory or motor fibers found in the somatic nerves. They comprise four sub-
groups, from highest to lowest speed: α, β, γ, δ. In the ANS the A fibers found
are Aδ.

4
V

Aα fiber
3

Aδ fiber
1

Fig. 1.9  Compound action C fiber


potential in a peripheral
nerve with peaks generated
by different types of nerve
fibers. Modified from 0 1 2 3 4 5 6 7
Cardinali [3] Latency, ms
References 17

• B fibers, myelinated, 1–3 mm thick with a velocity of 3–15 m/s. They constitute


the white communicating branches (preganglionic afferents) of the sympathetic
chain.
• C fibers, unmyelinated, <1 mm thick, with a velocity of <2 m/s. They are the
afferent amyelinic fibers of the visceral nerves and the sympathetic postgangli-
onic nerves.

References
1. Cannon WB. Organization for physiological homeostasis. Physiol Rev. 1929;9:399–431.
2. Moore-Ede MC, Czeisler CA, Richardson GS. Circadian timekeeping in health and disease.
Part 1. Basic properties of circadian pacemakers. N Engl J Med. 309:469–76.
3. Cardinali DP. Neurociencia aplicada. Sus fundamentos. Buenos Aires: Editorial Médica
Panamericana; 2007.
4. Saper CB. The central autonomic nervous system: conscious visceral perception and auto-
nomic pattern generation. Annu Rev Neurosci. 2002;25:433–69.
5. Hoit BD, Walsh RA. Chapter 5. Normal physiology of the cardiovascular system. In: Fuster V,
Walsh RA, Harrington RA, editors. Hurst's the heart. 13th ed. New York: McGraw-Hill; 2011.
6. Waxman SG. Chapter 20. The autonomic nervous system. In: Clinical neuroanatomy. 27th ed.
New York: McGraw-Hill; 2013.
7. Barrett KE, Barman SM, Boitano S, Brooks HL. Autonomic nervous system. In: Ganongs
review of medical physiology. 25th ed. New York: McGraw-Hill Education; 2016.
8. Waxman SG. Chapter 19. The limbic system. In: Clinical neuroanatomy. 27th ed. New York:
McGraw-Hill; 2013.
9. Strata P. The emotional cerebellum. Cerebellum. 2015;14:570–7.
10. Benarroch EE. Central autonomic network. Functional organization and clinical correlations.
Leander: Futura Publishing; 1997.
11. Kibble JD, Halsey CR. Neurophysiology. In: Medical physiology: the big picture. New York:
McGraw-Hill Education; 2015.
12. Waxman SG. Chapter 13. Control of movement. In: Clinical neuroanatomy. 27th ed. New York:
McGraw-Hill; 2013.
13. Ashley NT, Demas GE. Neuroendocrine-immune circuits, phenotypes, and interactions. Horm
Behav. 87:25–34.
14. Erny D, Hrabe de Angelis AL, Prinz M. Communicating systems in the body: how microbiota
and microglia cooperate. Immunology. 2017;150:7–15.
15. Buijs RM. Chapter 1. The autonomic nervous system: a balancing act. In: Buijs RM, Swaab
D, editors. Handbook of clinical neurology. Autonomic nervous system. New York: Elsevier;
2013, p. 1–11.
The Timed Autonomic Nervous System
2

Abstract
As the Earth rotates on its axis, it has two distinct environments: light and darkness.
As the axis of rotation of the Earth is tilted, the relative duration of the periods of light
and darkness changes systematically during the year. Because of the process of evo-
lution, living beings have responded to these two situations by developing specific
mechanisms for prediction and have successfully adapted to the time of day and the
seasons of the year. The biological rhythms are the basis of predictive homeostasis.
This Chapter describes the sleep/wake cycle as the major 24-hour rhythm as well as
the components of the three different ANS physiological programs that occur during
it. The significance of the glymphatic system and its link to sleep are discussed.

Keywords
Clock genes • Dreaming • γ-Aminobutyric acid • Glymphatic system • Melatonin •
Non-REM sleep • Orexin • Phase maps • Polysomnography • REM sleep •
Seasonality • Suprachiasmatic nucleus • Ventrolateral preoptic area • Zeitgeber

Objectives
After studying this chapter, you should be able to:

• Explain why the biological rhythms are the basis of predictive homeostasis.
• Summarize the behavioral and electroencephalographic (EEG) character-
istics of each of the stages of NREM and REM sleep and the mechanisms
responsible.
• Describe the pattern of normal night-time sleep in adults and the variations
in this pattern from birth to old age.
• Describe the interplay between brainstem neurons that contain norepinephrine
(NE), serotonin (5-HT), and ACh in addition to γ-aminobutyric acid (GABA)
and histamine (His) in mediating transitions between sleep and wakefulness.
• Discuss the sleep/wake cycle as the major 24-h rhythm and the role of the
suprachiasmatic nuclei (SCN) in its regulation.

© Springer International Publishing AG 2018 19


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1_2
20 2  The Timed Autonomic Nervous System

• Explain the components of the three different ANS physiological programs


(“body configurations”) that occur in a normal 24-h day/night cycle.
• Explain the meaning and mechanisms of dreaming.
• Describe the significance of the glymphatic system and its link to sleep.

Biological Rhythms Are the Basis of Predictive Homeostasis

As the Earth rotates on its axis, it has two distinct environments: light and darkness.
As the axis of rotation of the Earth is tilted, the relative duration of the periods of
light and darkness changes systematically during the year. Because of the process of
evolution, living beings have responded to these two situations by developing spe-
cific mechanisms for prediction and have successfully adapted to the time of day
and the seasons of the year.
The brain pacemaker creates a “day” and “night” in the body, as an approximate
mirror of the outside world. We wake up every day at about the same time, relatively
independently of the previous time devoted to sleep. We have a greater tendency to
carry out certain tasks (physical or mental) at certain times of the day or night
depending on what chronotype we have (an early chronotype, “larks,” a late chrono-
type, “owls,” or an intermediate one).
We perceive the seasons in our emotionality, physical strength, or ability to lose
weight. The disturbances originating in a traveler of transmeridian flights, the emo-
tional imbalances that often accompany the onset of winter and the troubles experi-
enced by those workers who must comply with rotating shifts are proof of the existence
of biological clocks and calendars, in conjunction with the geophysical cycles.
Although every physiological response exhibits a 24-h rhythm, there are differences
between these rhythms at the time when a peak occurs. The “phase maps” are the
graphic description of these maxima for many physiological periodic changes (Fig. 2.1).
Such a sequence and spacing of the maximum values of daily rhythms reveal the ordered
cause–effect relationships in bodily processes of all kinds, from genomic to behavioral,
and their normality is what we could define as the quintessence of health [1].
These maps may experience temporary disruptions when the body is forced to make
a quick adjustment phase, as happens after a transmeridian flight [2]. In such circum-
stances, the different rhythmic functions do not resynchronize with the same speed and
the normal temporal relationships between phases are lost. Full resynchronization
requires a few days (about 1 day for every hour of phase shift), and during this period the
“jet-lag” syndrome is seen. Phase maps are also distorted in chronic or acute diseases,
even mild ones [3, 4]. As we discuss in Chap. 8, full recovery is achieved by controlling
for both the underlying disease and the accompanying chronobiological disruption.
It is now established that clock mechanisms are genomic. Since life originated
about 4,500 million years ago, in an environment where day and night already
existed, successful species reproduced in their genome such geophysical reality.
Thus, the day and night have left an indelible mark, which is as universal as the
genetic code in all forms of life. In every living cell, a cyclic mechanism of interac-
tion between transcription factors, genes, and proteins exists close to 24-h periods
(in humans, slightly longer than 24 h; Fig. 2.2).
Biological Rhythms Are the Basis of Predictive Homeostasis 21

White blood cells Lymphocytes


TSH
Gastric secretion Melatonin
Eosinophils
Alkaline phosphatase
GH, PRL
Insulin, Leptin
Uric acid
24 ACTH
Platelets 22 2
20 4 FSH, LH
Triglycerides

Cholesterol 18 6 Cortisol
Aldosterone
GHrelin
16 8 Testosterone
Catecholamines
14 10
12 Platelet aggregation,
Body temperature blood viscosity

NK cells
Forced expiratory
Hemoglobin
volume Erythrocytes

Fig. 2.1  Phase maps of various circadian rhythms in man. As the individual is in normal light–dark
conditions, each of the rhythms shown is a period of exactly 24 h. What varies for each one of them is the
time in which the daily maximum (or “acrophase”) of the rhythm, shown in the diagram) is presented.
A synchronized phase map characterizes normality. Reproduced with permission from Cardinali [1]

These intertwined feedback loops involve a small number of clock genes [5].
The positive arm of the daily clock consists of transcription factors Bmal1 and
Clock. The protein products of these genes form heterodimeric complexes that
control the transcription of other clock genes, in particular, Per (Per1, Per2,
Per3) and cryptochrome (Cry1, Cry2), which in turn provide the negative signal
feedback that inhibits Bmal1 and Clock to complete the circadian cycle. Other
clock genes (Rev-erbα, Rorα, NR1D1, timeless) provide additional force to the
translation/transcription loops. The expression of clock genes is cyclic and is
regulated in part by phosphorylation of proteins, thus controlling protein stabil-
ity, nuclear reentry, and transcription complex formation (Fig. 2.2). Collectively,
the molecular circadian clock operates in every cell of the body to regulate at
least half the genome [6].
To generate physiological and behavioral responses consistently, the phases of
these myriads of cellular clocks must be orchestrated by a pacemaker. One of para-
mount importance resides in the SCN of the anterior hypothalamus [7]. This core
master clock is a key regulator of 24-h cycles in many body functions, including
sleep and wakefulness, thermoregulation, glucose homeostasis, and fat metabolism
[8]. SCN integrity is necessary for the generation and maintenance of the 24-h
rhythms, and for their synchronization by light–dark cycles. Although the complex
behaviors such as sleep, waking or feeding involve many brain areas working in a
network, in the case of circadian rhythms, the participating brain region is single
and has minimal volume.
22 2  The Timed Autonomic Nervous System

Cry
Per1 Per3

Per2
Clock Per2
E-Box Cry
Per2 Cry
Clock
BMAL1
E-Box Cry
BMAL1 Cry E-Box
Clock Cry Per2
Rev-Erba BMAL1 E-Box
E-Box Per2 Per1
Rev-Erba
E-Box
RORA E-Box Per1
RORA Per3
E-Box
Per3

E-Box
12 CLOCK CONTROLLED GENES
11 1 CSNKle

10 2 P Per2
P P
CSNKle
9 3
METABOLIC RHYTHMS Cytoplasmic degradation
8 4
Neuronal Behavior
7 5 activity Hormone Gene
6 secretion expression

Fig. 2.2  Simplified model of the circadian intracellular mechanisms. The process begins when the
CLOCK and BMAL1 transcription factors are dimerized and trigger the transcription of the Per
(Per1, Per2, and Per3) and Cry (Cry1 and Cry2) genes. Per and Cry are translated into their respec-
tive proteins, which increase throughout the day. When the PER and CRY proteins reach a certain
level, they form heterodimers that enter the nucleus and regulate the CLOCK-mediated BMAL1
transcription of their own genes. This process takes about 24 h. There are also accessory mechanisms
to the clock (e.g., Rev.-ERBA and RORA). Almost 50% of the transcriptome shows a 24-h variation

The circadian apparatus includes [7]: (a) a hypothalamic pacemaker, the SCN; (b) a
series of physiological outputs under the control of SCN; (c) molecular clocks present in
cells of all tissues and organs. The SCN has hypothalamic (endocrine) and extrahypotha-
lamic projections of a behavioral type. The circadian effects of SCN are exerted: (a) on
neuroendocrine neurons of the hypothalamic paraventricular nucleus (PVN); (b) on auto-
nomic PVN neurons; (c) on hypothalamic structures associated with sleep generation
(e.g., the ventrolateral preoptic area, VLPO); (d) other hypothalamic areas (sub-PVN
zone, sPVz; dorsomedial hypothalamus, DMH; median preoptic area, MnPO), interme-
diate between the SCN and autonomic and neuroendocrine neurons; on extrahypotha-
lamic structures (lateral geniculate body, paraventricular thalamic nucleus) for the
synchronization of hypothalamic conducts and locomotor activity (Fig. 2.3) [9].
The SCN contains local projection neurons that communicate with one another
and with other hypothalamic structures [10]. The axons of many SCN neurons ter-
minate within the nucleus itself, thus forming local circuit connections and/or col-
laterals from longer range projections. The SCN core projects densely to the SCN
shell, which projects only sparsely back to the core. Neuronal cell bodies in the
SCN are small (~10 μm), have simple dendritic arbors, and are closely apposed.
Neurons in the SCN core and shell regions are distinguished by their neuro-
chemical content. The neuropeptide vasoactive intestinal peptide (VIP) is found in
Biological Rhythms Are the Basis of Predictive Homeostasis 23

Hypothalamic nuclei
receiving projections Subparaventricular Suprachiasmatic
from SCN nucleus nuclei (SCN)

Dorsomedial Paraventricular
nucleus nucleus

Medial preoptic
area

Pontine Hypothalamic Circadian


arousal homeostatic Superior cervical
temperature
system mechanisms ganglia
control

Sleep / wake cycle Circadian


Locus neuroendocrine
Diencephalic Sleep / wake
coeruleus control (ACTH, GH,
arousal switch Pineal melatonin
TSH, etc.)
system

Seasonal rhythms in
Circadian function of systems, tissues and organs
bodily functions

Fig. 2.3  Transmission of circadian information from the oscillator to the hypothalamic systems
that control circadian rhythms, including the sleep–wake rhythm. The key steps identified include
multisynaptic transmission from the suprachiasmatic nuclei (SCN) to the hypothalamic control
systems through adjacent nuclei of the anterior hypothalamus, multisynaptic transmission to the
pineal gland controlling the secretion of melatonin, direct SCN pathways to the sleep-promoting
and awakening regions, and the integration of reactive and predictive homeostasis in relation to the
sleep–wake rhythm in the medial preoptic area. Modified with permission from Cardinali [8]

about 10% of all SCN cells, whereas AVP is present in 20% of all cells [11]. The
VIP-positive neurons are mainly located in the ventral and central parts of the SCN
(core). In humans, the volume of the VIP core subdivision is 0.03 mm3 and contains
about 1,700 VIP-immunoreactive neurons, with a mean density of about 63,000
neurons/mm3 [12]. Besides neurons containing VIP, substance P, gastrin-releasing
peptide (GRP), calretinin- and calbindin-containing neurons are also found in the
SCN core (Fig. 2.4). Most of the AVP-positive neurons are in the dorsomedial part
of the SCN (the shell). In humans, the volume of the AVP subdivision is 0.2 mm3
and contains about 6,900 AVP-immunoreactive neurons, with a mean density of
29,000 neurons/mm3. In this region, neurons containing cholecystokinin (CCK) and
prokineticin 2 are found in addition to AVP neurons (Fig. 2.4).
In most SCN neurons, neuropeptides are colocalized with GABA, and almost
all synapses among SCN neurons are GABAergic. Electrophysiologically, it has
been shown that glutamate (Glu) is also a transmitter in the efferent pathways of
the SCN.
The increased electrical activity of SCN during the day occurs in both nocturnal
mammals such as the hamster or the rat and in daytime species such as the primates.
In primates, however, the secretion of corticosteroids and the onset of activity and
phase of sympathetic predominance and temperature rise occur at the beginning of
24 2  The Timed Autonomic Nervous System

INPUTS CELL TYPES

AVP
3V CCK
PK2
SCN SHELL

VIP
Glu. PACAP SP SCN CORE
GABA, NPY, GRP
5-HT CALB
OC CALR

Fig. 2.4  Suprachiasmatic nuclei organization illustrating the compartmentalization of inputs and
cell types and in a coronal plane. On the left, inputs originating from melanopsin-containing retinal
ganglion cells (Glu; PACAP, pituitary adenylate cyclase-activating polypeptide), intergeniculate
leaflet of the thalamus (NPY, neuropeptide Y; GABA) and dorsal raphe (5-HT). On the right the
different cell populations described, containing AVP, CCK cholecystokinin, PK2 prokineticin 2;
VIP; SP substance P, GRP gastrin-releasing peptide, CALB calbindin, CALR calretinin. 3V third
ventricle, OC optic chiasm. The figure was prepared in part using image vectors from Servier
Medical Art (www.servier.com), licensed under the Creative Commons Attribution 3.0 Unported
License (http://creativecommons.org/license/by/3.0/)

the light phase, and not during the dark phase, as in the rat. That is, the signal pro-
duced by the SCN on the different effectors mentioned is interpreted in different
ways in diurnal and nocturnal species.
Research into animals and humans has shown that only a few key environmental
periodic clues, relevantly the light–dark cycle, are effective at synchronizing the
internal clocks (with periods slightly longer than 24 h) to exactly 24 h [2]. A syn-
chronizer agent can “reset” or modify the phase of the biological clock (Zeitgeber,
time-giver in German). The variability of the response is always predictable,
depending on when the synchronizer stimulus is applied, circadian rhythms are
phase-advanced, phase-delayed, or remain unchanged (Fig. 2.5).
Biological Rhythms Are the Basis of Predictive Homeostasis 25

Light Light
delays the advances
clock the clock
Body temperature

Light delays
the clock

00:00 05:00 12:00 18:00 00:00

Melatonin Melatonin
advances delays the
the clock clock
Body temperature

Melatonin
advances
the clock

00:00 05:00 12:00 18:00 00:00

Fig. 2.5  Central body temperature response to the application of a light pulse or administration of
melatonin (3 mg). A pulse of light during the evening or the early part of the night, slows the clock
and sleep begins later on subsequent days. A light pulse during the second half of the night and
early morning advances the clock and sleep begins earlier on subsequent days. Melatonin has the
opposite effect. Neither light nor melatonin changed the clock if applied during the day (dotted
line). Reproduced with permission from Cardinali [1]
26 2  The Timed Autonomic Nervous System

Without the action of external time cues, the period of these oscillators tends to
be longer than 24 h (Fig. 2.6). The rate is set at exactly 24 h by the action of light,
which is the main (although not the unique) Zeitgeber in humans. Brief exposures
to morning light are sufficient to adjust the clock to the precise 24-h solar time. This
action requires an intact SCN (Fig. 2.6).
A group of ganglion cells located in the periphery of the retina and containing
melanopsin as a photopigment projects to the SCN and other hypothalamic areas
and is linked with the neuroendocrine response to light observed in all vertebrates
including man [13]. They do this through specific neural projections (the retinal–
hypothalamic tract), resulting in genomic activation of neurons in the SCN (Fig. 2.7).

Synchronized

10

Phase delayed
Days

20

Free running

SCN lession
30

Desynchronized

40
0 12 24
Time of day

Fig. 2.6  Locomotor activity rhythm in the hamster. The activity on successive days (blue bars)
coincides with the dark phase in the synchronized situation. If a few hours’ delay is made at the
beginning of the night, the animal adapts with a delay (equivalent to “jet-lag”). In permanent dark-
ness, the locomotor activity rhythm follows the endogenous period (greater than 24 h). After SCN
lesion, an abolition of rhythm is observed. Modified with permission from Cardinali [8]
Biological Rhythms Are the Basis of Predictive Homeostasis 27

Cerebral cortex Photoperiod Dawn


Dusk
Striatum, Limbic system

Thalamus

Hypothalamus, Forebrain SCN


Retina

Hypothalamic effects O
NH
CH3
Predictive & reactive H3C O
homeostasis N
H
Melatonin
Diencephalic Sleep Sleep
arousal sustem switch homeostat
Central and peripheral
Brainstem Circadian clocks
Pontine arousal
Ultradian oscilator REM-nREM
sustem

Fig. 2.7  The mechanisms triggering and maintaining sleep are in the hypothalamus and basal
forebrain. The ultradian rhythm slow-wave sleep/REM sleep depends on mechanisms of the brain-
stem. Melatonin has effects on both central and peripheral oscillators. Reproduced with permission
from Cardinali [1]

The information generated in these nuclei is transmitted to specific areas of the


basal hypothalamus, which control the two major channels of body communication:
the endocrine system and the ANS.
In humans, light during the first part of the night delays the clock and that during
the second part of the night and early morning accelerates the clock (Fig. 2.5). At
other times of the day, exposure to light exerts no appreciable effect to advance or
delay the phase of the circadian rhythms. This mechanism explains why exposure to
artificial light fastens the endogenous clock in the morning, whereas during the first
part of the night it tends to perpetuate and aggravate sleep deprivation by causing a
phase delay [2].
A major synchronizer of the SCN clockwork is pineal melatonin (Figs. 2.7 and
2.8) [1]. Melatonin synchronizes the human circadian system as per a phase response
curve that is about 12-h out of phase with the phase response curve produced by
light (Chap. 8). Projections of the SCN driving the daily melatonin rhythm inhibit
the firing of neurons in the sPVz of the hypothalamus. From this zone, a multisyn-
aptic pathway begins that includes the medial forebrain bundle, reticular formation,
and the intermediolateral cell column of the cervical spinal cord, the superior cervi-
cal ganglia (SCG), and the postganglionic sympathetic fibers that end near the
pineal cells to stimulate melatonin synthesis (Figs. 2.7 and 2.8).
28 2  The Timed Autonomic Nervous System

MELATONIN

1200 2400 1200


PIN

Peripheral SCN
Circadian clocks

RHT

SCG
Dawn
Dusk ILC

Fig. 2.8  Control of melatonin synthesis by environmental light. Retinal ganglion cell receptors
project via the retino-hypothalamic tract (RHT) to the SCN. From here, and through a multisynap-
tic pathway including the cervical sympathetics, pineal melatonin release is modulated. Melatonin
both feedback at the SCN and affects peripheral clocks. Reproduced with permission from
Cardinali [1]

Melatonin phase-shifts circadian rhythms in the SCN by acting on MT1 and MT2
melatonin receptors in SCN neurons [14]. The phase-and amplitude-altering effect
of melatonin is caused by its direct influence on the electrical and the metabolic
activity of the SCN. The circadian rhythm of melatonin secretion has been shown to
be responsible for sleep rhythm in both normal and blind subjects (i.e., in the
absence of the synchronizing effect of light). If an individual who usually falls
asleep after midnight wishes to advance his or her sleep schedule to rise early for
work, the indication is the administration of melatonin at 18:00–19:00 h to achieve
the desired phase advance of the circadian clock (Fig. 2.5) [1].
The SCN communicates day–night cycle information to the rest of the body through
neural and humoral signals, including the ANS and the neuroendocrine system [7]. By
this information, the peripheral cellular circadian clocks become synchronized to exactly
24 h. As is discussed in Chap. 4, the clocks at the periphery are also able to respond to
other environmental cues, such as food, altering their phases to these cues accordingly.
Both melatonin and cortisol secretion are controlled by the circadian clock and
in turn are essential for synchronizing peripheral circadian rhythms. The secretion
of melatonin is very consistent from day to day for a given period of life and shows
Biological Rhythms Are the Basis of Predictive Homeostasis 29

Spring Summer

Autumn Winter Autumn Winter

Pregnancy

Breeding season Breeding season

LH

Progesterone

16D 16D

Breeding Anovulatory Breeding


season period season

Fig. 2.9  Reproductive seasonal rhythm in the sheep (a short-day breeder). Modified from
Cardinali [8]

far fewer contingent changes because of the stress that cortisol shows. Thus, mela-
tonin reflects more accurately the circadian signal given by the SCN [1].
From an evolutionary point of view, it is not difficult to imagine that a successful
adaptation to the environment in which animals compete for nutrients that are
always scarce needs to optimize processes of high-energy consumption, such as
reproduction [15]. Thus, almost all species undertake seasonal mating in the wild
(Fig. 2.9). The most appropriate signal to the circadian system to transmit informa-
tion about the season is the duration of photoperiod. Other environmental signals
(temperature, humidity, etc.) do not have the same degree of reproducibility year
after year as the length of the photoperiod.
In humans, there is seasonality in reproduction. The statistics on births indicate
the seasonality in the reproductive process, with maximum activity during the sum-
mer. If multiple births are computed, which are indicative of ovarian overstimula-
tion and are therefore independent of social factors such as sex behavior, the
seasonal differences are even more significant. These studies have been conducted
in populations in the northern hemisphere in periarctic zones (Scandinavia, Labrador
Peninsula). In these countries, the activity of the pituitary–ovarian axis and the inci-
dence of conception in human populations decline during the dark months of the
year.
30 2  The Timed Autonomic Nervous System

In comparative studies between summer and winter conducted in Finland,


increased melatonin secretion was observed during the winter, coinciding with the
decline of ovarian hormone release [16]. The overnight pulse of melatonin in
the plasma extends for a few hours in the morning during the winter because of the
short duration of the day (3–4 h) and low light intensity caused by the very oblique
incidence at a latitude of 65° north, light intensity being insufficient to suppress
melatonin secretion. This may be the signal that triggers gonadal involution.
Menarche has a seasonal incidence, with peaks in the spring and summer. Seasonality
in the reproductive process also occurs in men [17].
Long day breeders such as rodents, with a gestational period of less than a month,
mate at a time that allows the pups to be born at a time of the year that maximizes
survival (the summer). In a long breeder rodent such as the rat kept under optimal
light conditions and laboratory feeding, the seasonal changes tended to disappear.
However, it can be restored by the administration of melatonin in the drinking water
in a form that resembles a photoperiod exposure to winter [18].
In short day breeders such as sheep or deer, mating takes place in the autumn,
when nights lengthen and prolonged levels of melatonin secretion occur (Fig. 2.9).
This is because the gestational period in this species is longer (5 months); thus, the
birth of the pups occurs in the period when chances of survival are likely to be great-
est (spring). In these animals, melatonin stimulates reproductive function by decreas-
ing the sensitivity of the hypothalamus to the negative feedback of gonadal steroids.
Another seasonal rhythm is that of mood [19]. Although seasonal trends of mood
and emotions have been recognized in the medical literature for over 100 years, it
was not until the 1980s that the distinguishing features of an affective disease
involving the recurrent winter depression were established. This form of affective
illness (“seasonal affective disorder”) is now the subject of numerous investigations
[19]. Every fall or winter, patients with this condition get tired easily, eat high-­
calorie carbohydrate diets, show weight increases, and have exaggerated anxiety or
sadness. With the arrival of spring, patients emerge out of depression, and in certain
circumstances, may show moderate manic symptoms.
Light therapy is useful in seasonal affective illness: the morning exposure to light
(intensity of at least 2,500 lux, equivalent to the light intensity of a sunset) for 2 h
daily is recommended [19].
One could see seasonal affective disorder as the human equivalent of hiberna-
tion. Decreased libido in the winter reduces the chances of birth the following win-
ter (an inappropriate time of the year for newborn survival). In addition, food intake
increase in the winter facilitates reproductive success because overweight mothers
tend to have larger fetuses. In the winter, at high latitudes, nocturnal melatonin
secretion in humans is about 40 min longer than in the summer [16]. In a hibernat-
ing animal species, the prolongation of 30 min in the secretion of melatonin is suf-
ficient to signal winter.
Evidence has now accumulated that a seasonal change in thyroid hormone avail-
ability within the brain is a crucial element for seasonality [15]. This is mediated by
local control of thyroid hormone metabolizing enzymes within ependymal cells lin-
ing the third ventricle of the hypothalamus. Within these cells, deiodinase type 2
The Sleep/Wake Cycle as the Major 24-h Rhythm 31

enzyme is activated in response to the length of summer days, converting metaboli-


cally inactive thyroxine to tri-iodothyronine. The pars tuberalis of the pituitary
gland plays an essential role. Specialized thyrotroph cells are regulated by the
changing signal of day length, with long days activating TSH. In mammals, the pars
tuberalis is regulated by the nocturnal melatonin signal [15].

The Sleep/Wake Cycle as the Major 24-h Rhythm

A fundamental circadian rhythm is that of sleep/wakefulness. The Greeks called


sleep “the brother of death,” because they thought that in the sleeping man, the
soul temporarily abandoned the body and wandered through the world. Indeed,
sleep is an active process, the metabolism of several brain regions being greater
than in wakefulness. Moreover, sleep is complex, with two different electroen-
cephalographically (EEG) defined stages, NREM and REM sleep. It is also
endogenous and relatively independent of exogenous factors. During REM sleep,
O2 consumption in various brain areas of the limbic system exceeds that found in
wakefulness [20].
In rats, total sleep deprivation (NREM plus REM sleep) causes death in about
20 days. Hair loss and discoloration, tail and leg skin lesions and increased food
intake, together with 20% body weight loss and a doubling of energy expenditure,
are seen in sleep-deprived rats. If the deprivation is only of REM sleep, rats die in
about 40 days, with the same abnormalities. On average, humans cannot live with-
out sleep for more than a few days (about 2 or 3 days). After sleep deprivation,
humans recover approximately 33% of the total sleep time, 100% of the slow-wave
sleep, and 30–50% of the REM sleep lost [20].
A reduction of daily sleep hours is common in contemporary society, which pri-
oritizes a timeless organization, with continuous activity 24 h, 7 days a week (“24/7
society,” Chap. 8). On average, we sleep about 2 h less every day than just 40 years
ago, and this affects our cognitive performance, emotional stability, and health [1].
Electroencephalography recording and ocular musculature activity during sleep
has allowed the study of the “basic architecture” of the neural process. Four stages
of sleep are distinguished according to the type of EEG brain activity (Fig. 2.10).
Stages N1–N3 correspond to the progressive slowing of the brain waves, from the α
rhythm (8–13 Hz) to the δ rhythm (<3 Hz). A fourth stage, REM sleep, corresponds
to a desynchronized EEG, similar to that of wakefulness [20].
Rapid eye movement (REM) sleep is accompanied by a profound loss of mus-
cle tone (except for the diaphragm, the muscles of the middle ear, and the crico-
esophageal sphincter, all of which have very few muscle spindles),
ponto-geniculo-­occipital (PGO) spikes on the EEG, eye movements synchronous
with PGO spikes, and ANS signs, such as increased variability in BP and heart
rate and the tendency toward poikilothermia (loss of control of body temperature).
The more complex mechanisms of cardiocirculatory, respiratory, and thermal
feedback temporarily cease to function during REM sleep, with only the auto-
nomic spinal reflexes remaining.
32 2  The Timed Autonomic Nervous System

clocktime
23 24 1 2 3 4 5 6 7

W (wakefulness)

N1
Slow wave N2
sleep
N3
1 cycle (approx. 90 min)
R
REM sleep

Fig. 2.10  Sleep stages based on electroencephalography (EEG). There is a slowing of the brain
waves (N1–N3 stage) depending on the progression of sleep, reaching deep slow-wave sleep in
about 30–45 min. At this time, a progressive acceleration of EEG occurs, running from stage N3
to N1 in about 30–45 min to reach the REM sleep stage (R), which lasts about 10–15 min. There
are between 4 and 6 cycles per night. Reproduced with permission from Cardinali [1]

The normal polysomnographic register in humans indicates a slowing of the


cerebral waves (stages N1–N3). An acceleration of the EEG rhythm then occurs,
passing from stage N3–N1 in about 30–45 min (Fig. 2.10). Abruptly, a stage of
REM sleep occurs that lasts about 10–15 min, beginning a new cycle of slowdown
and subsequent acceleration of brain waves (Fig. 2.10). There are four to six REM
sleep episodes per night, but a δ rhythm (N3 stage) is only detected during the first
half of the night. In the second part of the night, the inter-REM periods show
increased EEG frequency, but not passing the stage 2 or 3 in the final hours of sleep.
In contrast, the REM periods are longer in the second part of the night. On average,
normal adult polysomnographic recordings demonstrate 25% REM sleep and 75%
NREM sleep (slow-wave sleep; 50% in stages N1 and N2, and 25% in N3). Video
recordings detect a postural change of importance every 20 min, approximately, all
of them corresponding to NREM sleep.
Three mechanisms have been identified as responsible for sleep (Fig. 2.11) [21]:

• A process called “S” (for sleep), determined by the previous individual history of
sleep and wakefulness. The “S process” is manifested in the increased propensity
to sleep after sleep deprivation. It is the accumulation of sleep debt, like the
mechanism of an hourglass.
• A process called “C” (for circadian), controlled by the endogenous biological
clock. It is independent of the previous history of sleep and wakefulness. The “C
process” comprises the trend toward falling sleep with the decrease in body tem-
perature (during the first part of the night) and the termination of sleep during the
increase in body temperature (during the second part of the night). This is
because of the “C process,” that is, after a night spent awake one is sleepier at
04:00–05:00 h than at 07:00–08:00 h, regardless of the 2–3 h increase in sleep
debt. Therefore, sleep is like a bank debt: it is impossible to pay the debt when
the “bank window” (C process) is “closed.” For a night worker who after a
The Sleep/Wake Cycle as the Major 24-h Rhythm 33

Process S

Sleep propensity
Sleep

2300 h 2300 h 0700 h

Process C
Sleep propensity

Sleep

2300 h 2300 h 0700 h

Ultradian
Slow wave sleep
Slow wave intensity

REM Sleep

2300 h 2300 h

Fig. 2.11  Three interacting processes regulate the timing, duration, and depth, or intensity, of
sleep: a homeostatic process that maintains the duration and intensity of sleep, a circadian rhythm
that determines the timing of sleep, and an ultradian rhythm given by the NREM sleep–REM sleep
sequence. Reproduced with permission from Cardinali [1]

sleepless night wants to sleep in the morning, he must wait until a more appropri-
ate time (e.g., the siesta after lunch) to have a restful sleep.
• An ultradian component (frequency of about 90 min), perceptible both in sleep
(slow-wave sleep and REM sleep alternation) and in wakefulness (periodicity of
about 90 min with maximal and minimal attention, the so-called basic rest–activ-
ity cycle, BRAC) [22].

It is important to note that these three mechanisms apply not only to sleep, but
also to the ultradian and circadian variations of most of the physiological phenom-
ena in which they have been examined. Studies conducted in sighted individuals
under normal sleep–wake cycles cannot establish whether the rhythm is endogenous
and self-sustained (the definition of a circadian rhythm) or is driven by external fac-
tors such as sleep–wake or rest–activity cycles, the light–dark cycle, or feeding
cycles. The constant routine protocol is a gold-standard method employed in circa-
dian biology to differentiate whether or not a rhythm is intrinsically generated and
sustained and, compared with measures taken under ambulatory baseline
34 2  The Timed Autonomic Nervous System

conditions, can measure the extent to which it is driven by external factors such as
sleep, light, meal timing, or posture [2]. During a constant routine procedure, par-
ticipants remain awake in bed for typically 30–50 h (1–2 circadian cycles) in a
semirecumbent posture under dim light conditions and are fed hourly isocaloric
meals. This procedure removes the direct impact of sleep, light, activity, and posture
on rhythm expression and distributes calorie intake uniformly across the circadian
cycle, hence removing or minimizing many invoked effects due to external factors
that may mask the underlying endogenous circadian rhythm. An important point to
consider in relation to the ultradian/circadian relation is that the ultradian period of
90 min is a harmonic of 24 h and that its progressive consolidation (rhythms of
1.5 h, 3 h, 6 h, 12 h, 24 h) is seen as the ontogenic development of the sleep/wake
rhythm in human newborns.
The C process involves circadian changes in the promotion of wakefulness given
by the SCN (Fig. 2.12) [7]. In primates that have lesions in the SCN, not only the
desynchronization of circadian rhythms occurs, but a significant increase in the total
time asleep (having removed a key area for maintaining wakefulness) is seen
(Fig. 2.6). During the day, the electrical activity of the SCN increases and peaks
toward the evening (around 18:00 h). SCN neurons help to counteract the increased
pressure of sleep debt that accumulates during wakefulness [7].
Pineal melatonin begins to be released at the end of the afternoon, at about
18:00 h. Melatonin acts on specific receptors located at the SCN, reducing their
electrical activity and therefore their ability to neutralize the pressure of the

Fig. 2.12  The evening rise in melatonin feeds back to inhibit the wakefulness-promoting effect of
the SCN. This is the trigger for consolidated sleep. Reproduced with permission from Cardinali [1]
The Sleep/Wake Cycle as the Major 24-h Rhythm 35

Homeostatic
drive of sleep
Melatonin
secretion starts

SCN feedback via MT1


Circadian drive receptors
of sleep

Inhibition of arousal
system

Sleep begins

Fig. 2.13  The homeostatic sleep pressure increases during the day and is counteracted by a strong
circadian promotion of wakefulness from the SCN. The secretion of melatonin inhibits SCN elec-
trical activity, thus triggering sleep. Reproduced with permission from Cardinali [1]

S process [1]. There is evidence that this abrupt change in sleep propensity is crucial
for sleep induction. Thus, melatonin is considered to be the signal that “opens the
gates of sleep” (Fig. 2.13).
The best candidate in the search for the hypnogenic substance that mediates the
homeostatic (S process) is an astrocyte-derived nucleoside, adenosine, acting on the
sleep-active GABAergic VLPO and MnPO, and on the basal forebrain (BF;
Fig. 2.14). Adenosine is a cellular product that accumulates because of metabolic
activation in tissues and thus indicates the degree of activity. Caffeine, theobromine,
and other xanthines are inhibitors of adenosine receptors, and coffee and tea may
promote wakefulness by interfering with this mechanism [23].
The extracellular level of adenosine in the brain increases during prolonged wak-
ing. In several brain regions, the stimulation of presynaptic A1 adenosine receptors
depresses Glu release and reduces the amplitude of excitatory postsynaptic currents.
This is the mechanism by which increased adenosine levels after sleep deprivation
influence the light responsiveness of the circadian clock. Indeed, after a 6-h sleep
deprivation, the light response in the SCN was reduced compared with the control.
Systemic injection of caffeine restored this attenuation of the SCN light response
almost completely [24], supporting the interactive role of adenosine and Glu. The
reduced response to light of SCN neuronal activity after sleep deprivation provides
evidence that the pacemaker may be modified by sleep homeostatic pressure.
In addition, the cortical neuronal population expressing neuronal nitric oxide
(NO) synthase has recently emerged as another candidate for involvement in the
homeostatic physiological sleep response (S process) [23]. The accumulation of
adenosine during the increase in IL-1 in infectious processes, given by the increase
in prostaglandin (PG) D2 in the organum vasculosum laminae terminalis (OVLT)
explains in part the symptomatology of the disease behavior (Chap. 4). IL-1 has also
been linked to daily sleep debt (Fig. 2.14).
36 2  The Timed Autonomic Nervous System

Sleep/wake cycle
Adenosine
IL-1 NO, PGE2
NFκ-B
GHRH
Thalamus

ACh

nREM Sleep GABA


VLPO
HIST

5-HT
NE

Fig. 2.14  Effect of humoral signals on slow-wave sleep. NFκB transcription factor kappa B

Neurophysiology of Sleep

The pioneering research of Bremer at the mid-1930s (Fig. 2.15) demonstrated that


lesions at a low medullary level in cats did not modify the sleep–wake cycle
(“encephale isolé”), whereas a cut between the pons and intercollicular midbrain
produced chronic sleepiness (“cerveau isolé”) [25]. Years later, other investigations
demonstrated that a midpontine section suppressed REM sleep, an indication that
the REM sleep generator was located below this level (Fig. 2.15).
Bremer’s observations ruled out the initial concept that sleep was a merely pas-
sive phenomenon produced by the inactivation of sensory stimuli arriving at the
diencephalic/telencephalic structures. Following Bremer’s ideas, Moruzzi and
Magoum demonstrated that the electrical stimulation of pontine reticular formation
activated the cortex and produced awakenings, and concluded that the forebrain was
kept alert by the tonic activity of the pontine reticular formation [26]. Thus, during
those years, the overall framework of knowledge held that the reticular formation
maintains an ascending activation from wakefulness to the thalamus and cortex,
producing recordings of EEG activation or unambiguous wakefulness. It was
hypothesized that a passive inactivation of the pontine reticular formation caused
the reduction of sensory input and generated sleep. However, many observations did
not fit with this theory. For example, thalamic stimulation produced sleep or wake-
fulness depending on the frequency of stimulation. Likewise, transection of the
pons, rostral to the fifth cranial nerve, induced wakefulness, thus questioning the
Neurophysiology of Sleep 37

Continuous NREM sleep

Sleep/wake cycle without REM sleep

Normal sleep/wake cycle


Thalamus
Cerveau isole

Mediopontine pre-trigeminal
Cerebellum

Encephale isole

Fig. 2.15  In 1935, and working on sleep EEG in cats, Frédéric Bremer identified the changes
produced by surgical cuts. He noticed that in the cerveau isolé there was permanent sleep, whereas
in the encéphale isolé the wake/sleep rhythm was maintained. Medial pontine cuts eliminated
REM sleep, leaving the slow sleep/wake rhythm intact. This located the origin of REM sleep in the
structures of the brainstem

theory of a passive inactivation of pontine reticular formation. Indeed, the findings


were compatible with the view that inputs from the lower pons or medulla inhibit a
wakefulness center in the rostral pons or rostral sites (thalamus) to induce sleep.
Thus, sleep was not merely a result of the deactivation of arousal centers, but an
active state of the brain.
After World War I, a worldwide epidemic of influenza with neurotropic sequelae
led the Rumanian neuropathologist von Economo to identify in patients three types
of lesions associated with different premortem statuses of sleep and waking [27].
These were:

• Type 1: lesions from the posterior hypothalamus variably extending to mesence-


phalic reticular formation associated with premortem somnolence or coma.
• Type 2: lesions of the anterior hypothalamus (VLPO and MnPO) and nearby
areas of basal forebrain associated with insomnia
• Type 3: lesions of the posterior–lateral hypothalamus, commonly in type 1 survi-
vors of a somnolence/coma syndrome, associated with narcolepsy.

From these observations, von Economo concluded that the posterior hypothala-
mus contains promoters of wakefulness and the anterior hypothalamus promoters
for sleep. These findings were confirmed by subsequent research demonstrating that
lesions of the anterior hypothalamus or preoptic/BF (substantia innominata and
horizontal limb of the diagonal band of Broca) reduced sleep and conversely their
electrical stimulation produced sleep onset [23].
38 2  The Timed Autonomic Nervous System

Histaminergic and orexin-containing neurons of the posterior–lateral hypo-


thalamus constitute the posterior hypothalamic wake promotion center.
Tuberomammillary nucleus (TMN) histaminergic neurons are the only source of
His in the brain and these cells extensively project innervation to the forebrain and
brainstem (Chap. 4). The TMN firing rate has a decreasing pattern from a con-
tinuum from wakefulness to NREM sleep and REM sleep, and plays a fundamen-
tal role in the generation of wakefulness [23].
The orexinergic neurons are found only in the lateral hypothalamic area (LHA)
and project widely to the brain and spinal cord. The orexin A and B neuropeptides
are excitatory, acting via OX1 and OX2 receptors and promote waking by activating
forebrain and brainstem wake-active cell groups (Fig. 2.16). The most important
evidence that orexins play a fundamental role in the regulation of wakefulness and
sleep was the demonstration that in narcolepsy with cataplexy a loss of orexin sig-
naling occurs. Orexin neurons act as a conductor of orchestration for vigilance
states, behaviors, and autonomic functions. Body temperature regulation by orexin

Suprachiasmatic
LDT/PPT NSQ
nuclei
Acetylcholine

VTA Amygdala
Dopamine DMH
LC
Norepinephrine
Dorsal raphe LHA
VLPO
Serotonin Orexin
GABA
PH
MB
Histamine
SN r ROX1
GABA Leptin Ghrelin Glucose ROX2
Metabolic
Excitation
signals
Inhibition

Fig. 2.16  Interactions of neurons containing orexin with other regions of the brain involved in
sleep and wakefulness regulation. Orexinergic neurons in the lateral hypothalamic (LHA) and
posterior hypothalamus (PH) are strategically located to serve as a link between the limbic system,
the systems involved in energy and monoaminergic homeostasis, and cholinergic neurons in the
brainstem. The SCN sends signals to orexin neurons through the dorsomedial hypothalamus
(DMH). VLPO ventrolateral preoptic area, NDR dorsal raphe, LC locus coeruleus, LDT laterodor-
sal tegmental nucleus, PPT pedunculopontine tegmental nucleus, SNR substantia nigra pars reticu-
lata, TMN tuberomammillary nucleus
Neurophysiology of Sleep 39

neurons seems to be mediated by one of its cotransmitters, whereas cardiovascular


and respiratory regulation are mediated by orexin itself [28].
The ascending reticular activating system involved in arousal and EEG activa-
tion, and consequently in wakefulness, is schematized in the left panel of Fig. 2.17.
In addition to the posterior hypothalamic wake-promoting center described above,
it is composed of the cholinergic laterodorsal tegmental (LDT) nucleus and pedun-
culopontine tegmentum (PPT) nucleus, the noradrenergic locus coeruleus (LC),
5-HT-containing dorsal raphe nucleus (DRN) neurons, and dopamine (DA)-
containing neurons in the ventral tegmental area (VTA) and substantia nigra that

LHA +
(orexin/ + LHA
MCH) (orexin/
+ MCH)

+
+ + + -
-
+-
BF -
(ACh, GABA) BF -
(ACh, GABA) -
VLPO, PPT/LDT (ACh) PPT/LDT (ACh)
MnPO VLPO, DRN (5HT)
PB/PC DRN (5HT)
(GABA) MnPO LC (NE)
(Glu) LC (NE) PB/PC
TMN SN/VTA (GABA)
(His/GABA LCα (Glu) TMN (Glu) LCα (Glu)
(DA) PZ SN/VTA
to VLPO) (GABA) MN (His/GABA PZ
(DA) MN
(Glycine) to VLPO) (GABA)
(Glycine)

Fig. 2.17  Left: schematic representation of the neuronal pools promoting arousal to the forebrain.
During wakefulness, the histaminergic neurons in the ventral tuberomammillary nucleus (TMN) at
the bottom of the posterior hypothalamus provide a strong inhibitory influence on the VLPO/
MnPO (median preoptic area). The components of the ascending reticular activating system fur-
ther include the raphe nuclei (5HT neurons), the locus coeruleus (LC, noradrenergic neurons), the
pedunculopontine and laterodorsal tegmenti (PPT/LDT) ACh-containing neurons, DA-containing
neurons in the substantia nigra (SN) and the ventral tegmental area (VTA), Glu-containing neurons
in the parabrachial/precoeruleus area (PB/PC) and the basal forebrain (BF), mainly cholinergic,
but also containing a population of GABAergic neurons whose stimulation produces sustained
wakefulness and EEG gamma activity [6]. Orexin-melanocyte concentrating hormone (MCH)
neurons on the LHA provide stimulatory input to the wakefulness-promoting areas. Right: sche-
matic drawing for a possible dual control, rostro-caudal and caudo-rostral, of sleep generation and
maintenance. The rostral sleep-promoting pathway includes the VLPO/MnPO area, which is
active during NREM sleep, and inhibits the activity of the arousal centers of the brainstem, hypo-
thalamus, and cortex. The caudal sleep-promoting pathway includes the parafacial zone (PZ) area
in the rostral medulla, which inhibits the PB/PC area, thus decreasing its wakefulness-promoting
activity on BF. The LC-α and magnocellular nuclei (MN), participating in REM-induced atonia,
are also depicted. Red labels denote activation, whereas blue labels represent inhibition. The figure
was prepared in part using image vectors from Servier Medical Art (www.servier.com), licensed
under the Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/
license/by/3.0/)
40 2  The Timed Autonomic Nervous System

stimulate the cortex directly and indirectly via the thalamus, hypothalamus, and
BF. Recent evidence indicates that the activation of a population of BF GABAergic
neurons produce sustained wakefulness and EEG gamma activity, and that the pon-
tomesencephalic PBN and precoeruleus area glutamatergic pathway to BF is very
important for EEG activation and wakefulness, as their lesions produced a coma-­
like state (Fig. 2.17) [23].
The VLPO and MnPO play a key role in promoting sleep. The neurons in these
nuclei contain GABA and the neuropeptide galanin and they innervate all the
arousal-promoting regions, including the LDT/PPT, LC, VTA, substantia nigra,
DRN, TMN, and the orexin neurons. Thus, the VLPO and MnPO promote sleep by
coordinating the inhibition of arousal regions during NREM and REM sleep
(Fig.  2.17, right panel). Recently, the identification of a slow-wave GABAergic
sleep-promoting center in the rostral medullary brainstem, in the parafacial zone
(PZ), added a new pathway. PZ neurons inhibit the parabrachial (PB)/precoeruleus
(PC), thus decreasing its wakefulness-promoting activity on BF and producing
NREM sleep [23]. The medullary Pz area inhibiting the PB/PC area could be the
counterpart of the VLPO for the dual control, rostro-caudal and caudo-rostral, of
NREM sleep generation and sleep maintenance (Fig. 2.17).
Since Aserinsky and Kleitman discovered REM sleep in humans in the mid-­
1950s, an impressive amount of research has demonstrated that the pons plays a
key role in the generation of REM sleep. In the mid-1970s the proposal that the
NREM/REM cycles arise from a reciprocal interaction between REM-off (mono-
aminergic) and REM-on cells (cholinergic) in the medial pons was made. A flip–
flop model for REM-off and REM-on neurons was proposed (Fig. 2.18). Two
neuronal pools of mutually inhibitory neurons in the upper pons form a switch for
controlling transitions between NREM and REM sleep. GABAergic neurons in the
ventrolateral periaqueductal gray and the adjacent lateropontine tegmental area fire
during NREM states to inhibit entry into REM sleep. During REM sleep, they are
inhibited by a population of GABAergic neurons in the sublaterodorsal region that
fire during REM sleep. This mutually inhibitory relationship produces a REM–
NREM flip–flop switch, promoting rapid and complete transitions between the two
states [20].
The core REM switch is also modulated by other neurotransmitter systems.
Noradrenergic neurons in the LC and serotoninergic neurons in the DRN inhibit
REM sleep by actions on both sides of the flip–flop switch (exciting REM–off and
inhibiting REM–on neurons), and during REM sleep they are silent. Cholinergic
neurons from the PPT/LDT promote REM sleep by having opposite actions on the
same two neuronal populations. The orexin neurons inhibit entry into REM sleep by
exciting neurons in the REM-off population (and by presynaptic effects that excite
monoaminergic terminals), whereas the VLPO neurons promote the entry into REM
sleep by inhibiting this same target. During REM sleep, a separate population of
glutamatergic neurons activates a series of inhibitory interneurons in the medulla
and spinal cord, which inhibit motor neurons, thus producing the atonia of REM
sleep. Withdrawal of tonic excitatory input from the REM-off regions may also
contribute to the loss of muscle tone. At the same time, ascending projections from
Neurophysiology of Sleep 41

vlPAG-LP

MCH
LDT-PPT

BF
PB/PC LC-DR

SLD
Ventromedial medulla

Ventral medulla
Spinal interneuron

REM ATONIA Motor neuron

Fig. 2.18  Schematic drawing illustrating the hypothetical circuitry involved in REM sleep regula-
tion. A flip–flop mutual inhibition between the sublaterodorsal tegmental area (SLD) REM-on
neurons and ventrolateral periaqueductal gray matter/lateropontine tegmental areas (vlPAG/LPT)
REM-off neurons are proposed to regulate transitions into and out of REM sleep. Pedunculopontine
tegmental/laterodorsal tegmental area (PPT-LDT) REM-on neurons inhibit lateropontine tegmen-
tal areas/ventrolateral periaqueductal gray matter (LPT-vlPAG) REM-off neurons, but they are not
mutually inhibited by the latter and thus they are not part of the REM flip–flop switch. The same
unidirectional relationship occurs with the serotoninergic dorsal raphe nucleus and noradrenergic
LC (DRN/LC) that activate REM-off neurons, but are not inhibited by the SLD REM-on neurons.
During REM sleep, glutamatergic neurons from the PB/PC area project rostrally to the basal fore-
brain (BF) and regulate the EEG components of REM sleep and caudally to the ventromedial
medulla and spinal cord to activate GABA/glycine interneurons inhibiting motor neurons to cause
atonia. The inhibition of the SLD given by vlPAG/LPT is overcome during REM sleep by neurons
containing MCH and other neurotransmitters. A population of GABAergic cells located caudally
at the ventral medulla also participates in the REM sleep switch, indicating a “dual command,”
rostral–hypothalamic and dorsal–medullary, for REM sleep regulation. Solid lines denote path-
ways that are active during REM sleep; dashed lines pathways that are inactive during REM sleep.
The figure was prepared in part using image vectors from Servier Medical Art (www.servier.com),
licensed under the Creative Commons Attribution 3.0 Unported License (http://creativecommons.
org/license/by/3.0/)
42 2  The Timed Autonomic Nervous System

glutamatergic neurons in the PBN activate forebrain pathways that drive EEG
desynchronization and hippocampal theta rhythms, thus producing the characteris-
tic EEG signs of REM sleep. Further research using optogenic tools showed that
REM episodes (duration and frequency) can be increased by the photostimulation
of melanin-concentrating hormone (MCH) projections in the TMN and median sep-
tum. Thus, MCH neurons constitute another input to consider in the REM–NREM
flip–flop model. Likewise, the finding of a ventral medulla GABAergic control of
REM sleep suggests an extended hypothalamic/midbrain/brainstem, perhaps redun-
dant, controlling REM sleep (Fig. 2.18). A “dual command,” rostral–hypothalamic
and dorsal–medullary, for REM sleep was thus proposed [29].
The different types of sleep change throughout life. In humans, REM sleep pre-
cedes slow sleep, which prevails in very early stages of life, so that it may play an
important role in the developing CNS (Fig. 2.19). Because REM sleep involves the
activation of many neural circuits, it is assumed to have a powerful internal drive
necessary for brain development and maturation in newborns.

10
Total sleep time
9
Hours/day

30 Percent time in REM sleep


Percentage

25

20

25 Percent time in N3 sleep

20
Percentage

Fig. 2.19 
Polysomnography 15
recording throughout a
lifetime. The depth of
10
NREM sleep decreases
with age. Reproduced with
permission from Cardinali 0 20 40 60 80 100
[1] Age (years)
Neurophysiology of Sleep 43

Slow sleep decreases exponentially with age and often disappears after
60 years of age [30]. This decrease in the depth of sleep causes frequent awaken-
ings and a return to an ancestral pattern of interrupted sleep. The duration of
slow-wave sleep decreases rapidly (almost 30 min per decade); thus, deep slow-
wave sleep (stage N3) after 50 years of age is less than 10% of the total sleep
period. This decrease in the depth of sleep is accompanied by an increase in the
N1 and N2 stages, whereas the duration of REM sleep and total sleep time
remains stable (Fig. 2.19). Many of the elderly often complain of sleep distur-
bances. In these cases, the physician should always investigate for the quality of
wakefulness. If there is no daytime sleepiness, sleep is sufficient and productive,
this is merely a complaint about sleep disruption resulting from the belief that
sleep is normal if not interrupted [20].
In humans, the maximum threshold for awakening occurs in the N3 stage,
whereas the minimum threshold for awakening corresponds to REM sleep. On aver-
age, 75% of normal individuals report dreaming when awakened during
REM. Although this percentage is higher than those recorded in the NREM stages,
what changes is the content of dreaming. As the time spent in REM sleep tends to
be proportionally greater as the night progresses, there is a tendency to wake up
toward the end of the REM sleep period, with more vivid and memorable dreams,
at least for a short period.
There are other differences between slow-wave sleep and REM sleep. The slow-­
wave sleep period depends upon waking and is prevalent when the individual has
not slept or slept little the previous night. In fact, the most accurate indicator of
homeostatic “debt” is the time spent in slow-wave sleep. This link with previous
waking does not exist for REM sleep. Both types of sleep also differ in their rela-
tionship to hypnotic drugs: alcohol and barbiturates decrease REM sleep, whereas
benzodiazepines affect it less, although sufficiently to result in a nonphysiological
sleep.
What are the bases of the physiological process by which stages of sleep and
wakefulness are defined? The cortico-thalamic circuit is instrumental in defining
wakefulness, slow-wave sleep, and REM sleep. Each of these stages is character-
ized by the level of thalamic “gate” to permit or not the passage of ascending sen-
sory information. Wakefulness and REM sleep are characterized by an “open gate,”
which allows the arrival of sensory information (exteroceptive in waking, interocep-
tive in REM sleep) into the cortex. In NREM sleep the gate is closed and there is
minimal information input into the cerebral cortex [30].
In Fig. 2.20, the physiological basis of this function is summarized. In thalamic
relay nuclei, most synapses on the main cells are derived from the cerebral cortex
(cortico-thalamic connections), brainstem, and inhibitory interneurons that control
the various modes of activity of the main cells and thus how the information is sent
from the thalamus to the cortex. The many inputs that operate on the thalamic relay
cells use various neurotransmitters [20]. Retinal and cortical afferents are always
excitatory and use the Glu as a neurotransmitter. Approximately 25% of the neurons
of the relay nuclei (e.g., lateral geniculate body) are cells with a small soma whose
dendritic tree extends along the sheet where they are located and do not project
44 2  The Timed Autonomic Nervous System

Sensory and motor cortex Cortico-thalamic Corticothalamic neuron


neuron activtiy after stimulation
Reticular neuron activity after cortico- of thalamo-cortical
thalamic stimulation neuron

Thalamo-cortical neuron activtiy


after stimulation of cortico-thalamic
Reticular hypothalamic neuron neurons
activity after stimulation of brain
stem cholinergic neurons
Reticular
thalamo
neuron

Thalamo-cortical neuron activtiy Thalamo-cortical neuron


after stimulation of brain stem
cholinergic neurons
Thalamus

Brain stem
cholinergic neuron

Fig. 2.20  Thalamic–cortical mechanisms. Represented neurons include pedunculopontine teg-


mental–laterodorsal tegmental cholinergic neurons, the GABAergic neurons of the reticular
nucleus of the thalamus, the glutamatergic neurons of the thalamic nuclei, and the glutamatergic
cortical neurons that project the thalamus. Different recordings, based on the transmission in the
form of relay or of gate of the circuit are shown. Cholinergic projections depolarize relay neurons
and hyperpolarize thalamic reticular neurons. The final result depends on the stimulation of
thalamo-­cortical and cortico-thalamic neurons on each one of them and the response characteris-
tics in salvos (spindles) of the thalamic reticular neurons (see text). Modified with permission from
Cardinali [8]

outside the core. They are interneurons with an inhibitory function and they use
GABA as a neurotransmitter [30].
Cortico-thalamic glutamatergic afferents synapse with reticular nucleus cells,
with inhibitory interneurons and with relay cells, so that the influence on core per-
formance thalamic relay has both excitatory (direct) and inhibitory (through inter-
neurons) components.
Afferents coming from the brainstem are heterogeneous. Mesencephalic reticu-
lar formation acts as a functional unit with the thalamus in what is defined as “gates.”
Important regulation of these gates is given by cholinergic influences, from the PPT/
LDT area. DRN serotonergic neurons and LC noradrenergic neurons also affect the
“gates.” To a lesser extent, VTA dopaminergic projections are also involved. These
inputs modulate globally the excitability of thalamic relay neurons, thereby influ-
encing how much information is directed to the cortex.
Three Different ANS Programs (“Body Configurations”) Occur in a 24-h Day/Night Cycle 45

There are thus two functional modes for thalamic relay nuclei cells (Fig. 2.20):
(a) “Open gate,” in which the transmission of information is carried out with the
triggering of action potentials of variable frequency only. Under these conditions,
transmission through the thalamic relay nuclei reliably reflects the arrival of infor-
mation from sensory pathways and the transfer rate approaches 1 (one PEPS pro-
duces one action potential).
(b) “Closed gate” when the thalamic relay cell is sufficiently hyperpolarized
(membrane potential = −75 mV or more negative), it operates in a salvo mode,
characterized by an initial depolarization with a variable number (2–7) of spikes
linked to the Na+ voltage-dependent channel. This is the basis of the “closed-gate
mode” that appears in slow sleep, and coincides with a very characteristic EEG pat-
tern, consisting of springs of waves at the frequency of 7–13 Hz, called sleep spin-
dles. This mechanism is cyclic and hyperpolarization leaves the cell ready to start
the process again.
Interestingly, in the “closed-gate” form, giant δ spikes are capable, as a grand
mal epileptic focus, to co-opt the activity of most CNS neurons at their own pace.
Indeed, it is as if a physiological “seizure focus” becomes the absolute regulator of
brain activity. This state of being prone to epilepsy, has a clinical relevance: the
EEG of epileptic patients during sleep records show a clear tendency toward the
predominance of epileptic seizures in slow-wave sleep stage [30].

 hree Different ANS Programs (“Body Configurations”) Occur


T
in a 24-h Day/Night Cycle

Sleep is not just a neurological phenomenon and a common mistake is to consider


it an exclusive phenomenon of the CNS. Together with wakefulness, slow-wave
sleep and REM sleep comprise three different ANS programs (Fig. 2.21 and
Table 2.1) [1, 8].
Several physiological functions vary both in the passage from wakefulness to
sleep and within each sleep stage; the sympathetic and parasympathetic systems,
key regulators of the automatic functions of the body, are responsible for these
changes. The sympathetic nervous system has evolved as predominant in wakeful-
ness and in response to major threat to our species during evolution, i.e., physical
trauma. It is thus linked to the consumption of energy (catabolism) to fight or flight
at the threat and promote mechanisms to mitigate the consequences of trauma: vaso-
constriction, increased blood coagulability, increased innate and humoral immunity
(that keep wounds germ-free), etc. (Fig. 2.21 and Table 2.1). The sympathetic ner-
vous system dominates wakefulness. This is a stereotypical hyperactivity, directed
to place the individual in a situation of defense in the face of circumstantial danger,
real or potential. Sympathetic overstimulation leads to variations in visceral func-
tions designed to protect the integrity of the organism and to ensure survival. In fact,
a sympathectomized animal hardly survives if left free in its natural environment.
46 2  The Timed Autonomic Nervous System

Fig. 2.21  The three different “bodies,” wakefulness, slow-wave sleep (NREM sleep), and REM
sleep, must necessarily follow each another harmoniously to ensure health. A 76-year-old man
sleeping 8 h daily lives 50 years in the physiological state of wakefulness, 20 years in slow-wave
sleep, and 6 years in REM sleep

Table 2.1  The three physiological states (“bodies”) of our life


Wakefulness NREM sleep REM sleep
“Active brain in an “Inactive brain in an “Hallucinating brain in a
active body” active body” paralyzed body”
Neurochemical Tonic firing of neurons Inhibition by VLPO Prevalent cholinergic
“microclimate” in the locus coeruleus area of the arousal activity (PPT nucleus)
(noradrenergic) and systems. Decreased concomitant with
raphe nuclei aminergic activity in extreme reduction of
(serotonergic) driven the face of a progressive aminergic activity. REM
by orexinergic increase in cholinergic sleep and wakefulness
hypothalamic neurons. activity (tonic firing of are states of cortical
Phasic discharge of the PPT neurons). Both activation with different
pedunculo-pontine responsible for neuromodulating pattern
nucleus of the pontine decreased (cholinergic vs
tegmentum, PPT consciousness noradrenergic) and
(cholinergic) different contents of
consciousness
Three Different ANS Programs (“Body Configurations”) Occur in a 24-h Day/Night Cycle 47

Table 2.1 (continued)
Wakefulness NREM sleep REM sleep
Afferent Actively functioning. Thalamo-cortical circuit Thalamic activity
Thalamocortical circuit in “closed-gate fashion” changes to operation
in “open-gate fashion” (which prevents sensory “open-gate fashion” as
so that sensory information from during wakefulness
information can reach reaching the cerebral
the cerebral cortex. cortex. A 25% decrease
Activated dorsolateral in cerebral blood flow
prefrontal cortex and oxygen
(working memory) consumption. Synthesis
of neurotrophins.
Glymphatic flow
increased.
Efferent Actively functioning Episodic muscle Skeletal muscle
activity, hypotonia paralysis as a protective
mechanism to prevent
the locomotor correlates
of a highly activated
brain
Content Attention, logical Disconnection, episodic Dream activity
awareness thinking, memory memory characterized by vivid
hallucinations, illogical
thinking and intense
emotion.
Perception Externally generated Absent Generated internally,
preferential activation of
the pons and limbic
system with deactivation
of the dorsolateral
prefrontal cortex
Physiological Predominance of Parasympathetic Disconnection of
pattern in organs sympathetic activity. hyperfunction in organs autonomic regulatory
and systems Augmented plasma NE and systems. GH, system (this prevents
and cortisol prolactin and insulin expression of dreaming
secretion emotions).
Antihomeostatic
physiology
Reproduced with permission from ref. [1]
PPT pedunculopontine tegmentum, NE norepinephrine, VLPO ventrolateral preoptic area, GH
growth hormone

Our species is programmed to eat only sporadically (every 2–3 days); thus, a


mechanism that optimizes the maximal intake at the right time was selected for
wakefulness. Thus, wakefulness is linked to an increased food intake given by the
secretion of orexinergic hormones such as ghrelin. Note that the same chemical
signal that increases appetite, orexin, is also a central neurotransmitter in maintain-
ing alertness (Fig. 2.16).
For modern man, the trauma has become a minor factor and instead new diseases
resulting from the prolongation of life and the type of diet and living conditions
48 2  The Timed Autonomic Nervous System

manifest. During evolution, endothelial injury and organ hypoxia were associated
almost exclusively with trauma. Today, endothelial injury is precipitated by stress-
ors such as hypertension, diabetes, or dyslipidemia. It is postulated that the patho-
physiology of cardiovascular disease implies a prominent amplification of a triple
response to trauma, (a) adrenergic response; (b) inflammation; (c) coagulation,
which are aggravated by the sleep deprivation conditions described above. Selected
components act to limit bleeding, defend infection of wounds, and initiate cell
reconstruction. These mechanisms are highly conserved, as indicated by the phylo-
genetic age of the renin–angiotensin system [31].
The parasympathetic system in slow-wave sleep serves as the anabolic counter-
part of the predominance of the catabolic sympathetic system during wakefulness.
It promotes energy accumulation, adaptive immunity, and augmented secretion of
anabolic hormones, such as growth hormone (GH), and of anorexic hormones, such
as leptin and insulin. The eighteenth-century French clinicians believed that the
parasympathetic system was the “master of sleep.” Today, we know that this must
be rephrased to point out that “the parasympathetic is the master of slow wave
sleep”, that is, of about 75% of the night (Fig. 2.21 and Table 2.1) [8].
Although the common view is that we humans are homeotherms (that is, we have
a regulated body temperature, Chap. 5), in a substantial part of our life we lack that
control. Wakefulness is characterized by a constant interaction of the hypothalamic
(automatic) and behavioral mechanisms (facultative: I have cold and seek shelter)
that control body temperature. In the passage to slow-wave sleep, the inactivation of
behavioral control occurs, but the temperature is still regulated by the automatic
processes discussed in Chap. 5. During REM sleep, the situation changes radically:
at this stage both forms of temperature control are halted and there is no heat pro-
duction to compensate for the cold. That is, during REM sleep, we acquire a similar
state to amphibians and reptiles, whose body temperature depends on ambient tem-
perature (poikilothermic animals).
Indeed, at REM sleep, most supraspinal autonomic reflex mechanisms are sup-
pressed: the complex mechanisms of cardiovascular, respiratory, and thermal con-
trol temporarily stop working, with only the basic autonomic reflexes of the spinal
cord persisting. Like a transoceanic flight in which most control mechanisms of
aircraft become disconnected for 10–15 min and the risk of an accident is high,
during REM sleep, there is a greater risk for strokes, heart attacks, and other acute
episodes. As in the latter part of the night, there is a prevalence of REM sleep,
such accidents tend to be higher late at night/early morning. This state of regional
disconnection is the equivalent of leaving the body without its basic homeostatic
mechanisms.
The impoverishment of slow-wave sleep and the consequent decrease in para-
sympathetic tone have strong effects on the neuroendocrine–immune network [32].
The observed immune changes include reduction of acquired immunity, particularly
of cellular immunity, whereas innate and humoral immunity tends to increase.
Many conditions that depend on an adequately controlled cellular immune response
(viral diseases, oncology, autoimmunity) are aggravated by this imbalance. In turn,
cancers and viral diseases are accompanied by a significant reduction in slow-wave
Three Different ANS Programs (“Body Configurations”) Occur in a 24-h Day/Night Cycle 49

sleep (and thus a greater parasympathetic withdrawal), either because they alter
directly NREM sleep via the inhibition of the secretion of melatonin or because
some of its symptoms trigger arousal (for example, coughing in lung disorders).
However, it should not be forgotten that there is no absolute predominance of one
system over the other, but a delicate interplay between the sympathetic and
­parasympathetic that is responsible for each of these body system configurations
(Chap. 1).
For all that has been said up to now, it is obvious that we cannot skip over the
slow-wave sleep repair period after spending several hours in the physiological set-
ting of wakefulness (of sympathetic predominance of the catabolic type, with high
energy consumption and potential damage to organs and tissues). Everything is pre-
pared during NREM sleep for the anabolic recovery, with the release of hormones
such as GH and typical responses of cellular immunity. This intricate and subtle
mechanism is altered during sleep deprivation [32].
Finally, it should be noted that not only are the mechanisms of sleep neural, but
that important humoral components also exist. The idea of a humoral origin of sleep
dates from ancient times. In the pre-scientific stage, it was thought that “vapors”
derived from the digestive system produced sleep. By the late nineteenth century,
“fatigue substances” or hypnotoxins were postulated from experiments in which the
cerebrospinal fluid (CSF) of sleep-deprived dogs produced sleep when injected into
control dogs. The hypnotoxin theory lost strength after Von Economo’s studies on
sleeping sickness, which clearly showed that lesions of the anterior hypothalamus
produce insomnia whereas lateral hypothalamic lesions lead to hypersomnia.
In the initial search for humoral factors inducing sleep, several substances were
identified. In the cerebral venous effluent of sleeping rabbits a nonapeptide was iso-
lated (δ sleep-inducing peptide). Three other substances isolated from the brain areas
of sleeping animals were characterized as uridine pyrimidine nucleoside, oxidized
glutathione tripeptide, and a “factor S,” later identified as muramyl peptide. It is of
interest that this peptide induces the synthesis of IL-1 by astrocytes.
The following criteria have been proposed to consider a humoral substance as a
regulator of sleep [33]: (a) it must induce or maintain physiological sleep or induce
sleep equivalent to that obtained after sleep deprivation; (b) the concentration, turn-
over or its receptors should vary with changes in sleep propensity; (c) normal sleep
should be inhibited by the effect of the inhibiting substance; (d) it should promote
sleep by acting on one or more parts of the neural circuitry shown to operate in
sleep; (e) the conditions that promote or inhibit sleep must be accompanied by
changes in the amount or metabolism of the substance; (f) the induced sleep should
be rapidly reversible without significant physiological consequences.
More recently, several substances have satisfied most of the requirements stated
above. An important one is the GH-releasing hormone (GHRH), a peptide of the
secretin/glucagon family. Slow-wave sleep is reduced in transgenic mice deficient
in receptors for GHRH or in GHRH production. GHRH action is exerted on the
VLPO area on a subset of GABAergic neurons. These neurons also respond to IL-1,
which raises the possibility that the mechanism by which IL-1 promotes sleep is by
increasing the response of these neurons to GHRH [34].
50 2  The Timed Autonomic Nervous System

Two cytokines are of great importance in inducing sleep, IL-1 and TNF-α. The
following evidence indicates their importance: (a) central or peripheral administra-
tion of IL-1 or TNF-α increases slow-wave sleep and suppresses REM sleep; (b) the
selective inhibition of IL-1 or TNF-α reduces NREM sleep; (c) knockout mice for
receptor type I IL-1 or p55 TNF show less slow-wave sleep; (d) sleep deprivation
increases mRNAs for IL-1 and TNF-α in the brain; (e) there is a parallel rate of
increase in mRNA and protein levels of IL-1 and TNF and of slow-wave sleep in
rodents.
The intracerebroventricular administration of TNF-α or IL-1 toward the latter
part of the period of wakefulness increases NREM sleep. By contrast, if adminis-
tered during the first part of the waking period TNF-α or IL-1 inhibits NREM sleep,
presumably by the increased activity of the pituitary–adrenal axis. TNF seems to
exert somnogenic effects by promoting the attraction of microglia and their pro-
cesses to the vicinity of the dendrites and synapses [35]. Both cytokines induce
fever, but this action is not linked to their somnogenic activity. Production of IL-1
and TNF-α by neurons, glial cells, and by endothelial cells in the CNS has been
shown, and it varies with the state of sleep propensity.
It is noteworthy that the injection of TNF-α or IL-1 into the somatosensory cor-
tex induces an increase in ipsilateral slow-wave sleep that is unobserved contralater-
ally. Conversely, the inhibition of cytokines produces local changes in EEG activity
that are not observed contralaterally. These results indicate a local action of cyto-
kines and are consistent with the idea that there is a “local sleep” in neural networks
that depends on the local metabolic activity and release of cytokines and other para-
crine or autocrine substances [36]. Some of the cytokines with effects on sleep are
listed in Table 2.2.

Table 2.2  Cytokines with activity on sleep


Prosomnogenic Antisomnogenic
IL-1 IL-4
Tumor necrosis factor (α and β) IL-10
IL-2 IL-13
IL-6 TNF soluble receptor
IL-8 IL-1 soluble receptor
IL-15 Insulin-like growth factor-1
IL-18
Epidermal growth factor
Fibroblast growth factor
Neural growth factor
Brain-derived neurotrophic factor
Neurotrophins 3 and 4
Glial-derived neurotrophic factor
Interferon (α and β)
Granulocyte macrophage colony-stimulating
factor
Tumor growth factor
Prokineticin 2
The Meaning of Dreaming 51

The Meaning of Dreaming

Rapid eye movement sleep is widely distributed on the zoological scale.


Therefore, it should be considered a process of great evolutionary significance.
Indeed, the body is jeopardized in the phase of sleep by the major disconnection
of the motor system and the ANS, a truly antihomeostatic state. One obvious
conclusion is that this paralysis helps to preserve the body of motor and vegeta-
tive dream acting, which would be of great danger near predators. But why
endanger homeostasis to preserve dreaming? The answer is probably the great
importance of sleep in the process of learning and memory, both procedural and
declarative (Chap. 6) [37].
Dreaming is a state characterized by illogical thinking, hallucinations, and emo-
tional changes. Strictly speaking, the dream is a “delirium” as it presents hallucina-
tions, disorientation, memory loss, and confabulation. For Sigmund Freud (The
Interpretation of Dreams, 1900) there is a manifest content of the dream story with
a latent or symbolic content. For Freud, dreams are the hallucinatory fulfillment of
desire and have the biological function of protecting sleep (insomnia occurs when
this mechanism fails).
Using a portable EEG recorder that distinguishes wakefulness, slow-wave sleep,
and REM sleep during normal activity of the individual, several aspects of dreaming
were examined [38, 39]. This led to the formulation of the activation–synthesis
model of dreaming, according to which the physical substrate of dreaming is the
cholinergic activation of an aminergically demodulated cortex during REM sleep.
This situation leads to a synthesis of visual hallucinations, loss of reflective knowl-
edge, emotional intensification, and memory loss [38, 39].
According to the model of reciprocal interaction, the REM phases and asso-
ciated dreams are activated/deactivated through specific nerve connections at
the brainstem. “REM-on” nerve cells (cholinergic) and “REM-off” (NE, 5-HT)
nerve cells interact with each other. According to the activation–synthesis
model, during sleep, brainstem neurons, and immediately after, areas of the
cortex and the limbic system, are stimulated (“activation”). Dreaming content
originates from random nerve impulses, produced by PGO (cholinergic) waves
in REM sleep. With these random signals, the sleeping brain attempts to do the
same thing it does in waking: integrate nerve impulses and give them meaning
(“synthesis”). That is, according to Hobson’s view, dreams are of an exclusively
random nature [38, 39].
Freud believed that the amount of sleep was determined by the day’s experience,
which activates emergency-related memories. However, from experimentally
obtained data it could be concluded that the dreams do not represent narrative or
episodic memories, but rather the aggregation of discrete memory and incomplete
narrative fragments to create a new synthetic script of dreaming [39].
However, there is way to link both scenarios, psychodynamic and neurocogni-
tive. Dreaming is a side effect of neural activity generated by interoception (ANS).
Interoceptive mechanisms are defined very early in life and are affected by the early
emotional environment (Chap. 5). As it can be postulated that the biological
52 2  The Timed Autonomic Nervous System

representation of Freudian preconscious is given by interoception, dreaming may


not be an aleatory event, but may depend heavily on the emotional history of the
individual.
The absence of episodic memory in dreams reflects the relative inaccessibility of
information from the hippocampus at the time (Chap. 6). Elevated levels of ACh
suppress the flow of information from the hippocampal cortex both in wakefulness
and in REM sleep. In individuals awakened during slow-wave sleep, there are ele-
ments of episodic memory (what happened during the day) that are characteristic of
dreaming at this stage.
The forebrain structures activated in REM sleep (and paralimbic/limbic areas)
give emotional and social content to dreaming. There is a very important activation
of the basal ganglia in REM sleep, which is not surprising in view of the relevant
participation of the limbic and cognitive circuitry of the basal ganglia in higher
brain functions (Chap. 6). The cerebellar vermis is also stimulated, indicating the
important role of the cerebellum in cognition and emotion [40].
Dreams have a visually predominant component. There are two main aspects in
the visual analysis: (a) spatial vision in the medial temporal and parietal regions; (b)
recognition of the object in the inferior temporal cortex [8]. It has long been known
that these medial temporal cortex and occipital areas involved in visual processing
are responsible for generating higher order visual imagery in dreams. These areas
are selectively activated during REM sleep. The disabling of executive areas of the
dorsolateral prefrontal cortex during slow-wave sleep and their failure to wake dur-
ing REM explain the prominent executive deficiencies occurring in sleep (disorien-
tation, lack of logic, less active memory, and amnesia).

The Glymphatic System and Sleep

The classic view of cerebrovascular physiology has been that blood flow and cere-
bral metabolism are tightly coupled under the influence of substances such as H+,
adenosine, NO, and K+ that ensure a rapid and matched supply of blood. In part
driven by the use of cerebral blood flow measurements by functional brain imaging,
it has become clear that astrocytes also play a role in modulating functionally asso-
ciated changes in cerebral blood flow. The autonomic innervation of the cranial
circulation has both a sympathetic component that arises predominantly from the
SCG and a cranial parasympathetic component that traverses the pterygopalatine
(sphenopalatine) and otic ganglion [41]. Neuropeptide transmitters such as neuro-
peptide Y (NPY), VIP, and pituitary adenylate cyclase-activating peptide (PACAP)
have each been identified in components of the system.
A previously unrecognized system that drains waste from the brain has recently
been characterized as active in sleep [42]. This system acts as a glial cell-dependent
pipe and has been called the “glymphatic system.” This pathway consists of CSF
influx into the brain parenchyma via para-arterial spaces, exchange of solutes
The Glymphatic System and Sleep 53

Para-arterial influx Cerebral arterial blood

capillary
Brain cells Ventricles with
Blood-brain barrier
choroid plexus
(Blood-CSF
barrier)
Convection

CSF
formation
Brain ECF
Ependyma
Para-venous clearance

Venules Arachnoid
villous

CSF-absorption:
500 ml daily

Cerebral venous blood

Fig. 2.22  Although the cerebrospinal fluid (CSF) influx is driven by arterial pulsation, the
exchange of solutes with the interstitial fluid and fluid movement through the brain parenchyma
are driven by convective bulk flow

(soluble proteins, waste products, metabolic wastes, and excess extracellular


fluid) with the interstitial fluid and clearance along para-venous spaces (Fig. 2.22)
[43, 44].
Although the CSF influx is driven by arterial pulsation, the exchange of sol-
utes with the interstitial fluid and fluid movement through the parenchyma are
driven by convective bulk flow rather than diffusion [43]. The exchange of sol-
utes between the CSF and the interstitial fluid occurs mostly during NREM sleep,
when the cortical interstitial space increases by more than 60% and provides a
low-resistance path for the movement of CSF and interstitial fluid in the brain
parenchyma (Fig. 2.23).
Brain water homeostasis is mediated by integral membrane pore proteins called
aquaporins, which transport and regulate water movement in the brain. Aquaporin-4
is predominantly present in astrocytic endfeet near capillaries and in cells lining the
ventricles, which are key sites for water movement among the cellular, vascular, and
ventricular compartments [44]. The continuous aquaporin-4 expressing astrocytic
endfeet lining the cerebral blood vessels create a low-resistance para-vascular chan-
nel for the movement of CSF. Postinjury reduction of aquaporin-4 expression has
been associated with exacerbated glymphatic system dysfunction and aquaporin-4
54 2  The Timed Autonomic Nervous System

Skin
Bone
Dura mater Awake
Arachnoid • Reduced interstitial space
Pia mater • Restricted CSF flow
• Metabolites accumulate

Cerebral cortex

Asleep
• 60% increase in interstitial space
• Increase in CSF flow
• Effective clearance of metabolites

Tissue perfused by CSF

Fig. 2.23  The exchange of solutes between the CSF and the interstitial fluid occurs during NREM
sleep, when the cortical interstitial space increases by more than 60% and provides a low-­resistance
path for the movement of CSF and interstitial fluid in the brain parenchyma. The figure was pre-
pared in part using image vectors from Servier Medical Art (www.servier.com), licensed under the
Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

knockout mice exhibited slowed CSF influx and ~70% reduction in interstitial fluid
solute clearance, indicating that the aquaporine-4 water channel mediates/facilitates
the glymphatic pathway [44].

References
1. Cardinali DP. Ma Vie en Noir. Fifty years with melatonin and the stone of madness. Switzerland:
Springer; 2016.
2. Golombek DA, Casiraghi LP, Agostino PV, Paladino N, Duhart JM, Plano SA, Chiesa JJ. The
times they’re a-changing: effects of circadian desynchronization on physiology and disease.
J Physiol Paris. 2013;107:310–22.
3. Oldham MA, Lee HB, Desan PH. Circadian rhythm disruption in the critically ill: an opportu-
nity for improving outcomes. Crit Care Med. 2016;44:207–17.
4. Agorastos A, Kellner M, Baker DG, Otte C. When time stands still: an integrative review
on the role of chronodisruption in posttraumatic stress disorder. Curr Opin Psychiatry.
2014;27:385–92.
5. Hastings MH, Brancaccio M, Maywood ES. Circadian pacemaking in cells and circuits of the
suprachiasmatic nucleus. J Neuroendocrinol. 2014;26:2–10.
References 55

6. Zhang R, Lahens NF, Ballance HI, Hughes ME, Hogenesch JB. A circadian gene expres-
sion atlas in mammals: implications for biology and medicine. Proc Natl Acad Sci U S A.
2014;111:16219–24.
7. Saper CB. The central circadian timing system. Curr Opin Neurobiol. 2013;23:747–51.
8. Cardinali DP. Neurociencia Aplicada. Sus Fundamentos. Editorial Médica Panamericana:
Buenos Aires; 2007.
9. Cardinali DP, Pandi-Perumal SR, editors. Neuroendocrine correlates of sleep/wakefulness.
New York: Springer; 2006.
10. Evans JA. Collective timekeeping among cells of the master circadian clock. J Endocrinol.
2016;230:R27–49.
11. Reghunandanan V, Reghunandanan R. Neurotransmitters of the suprachiasmatic nuclei.

J Circadian Rhythms. 2006;4:2.
12. Hofman MA, Zhou JN, Swaab DF. Suprachiasmatic nucleus of the human brain: an immuno-
cytochemical and morphometric analysis. Anat Rec. 1996;244:552–62.
13. Guido ME, Garbarino-Pico E, Contin MA, Valdez DJ, Nieto PS, Verra DM, Acosta-Rodriguez
VA, de ZN, Rosenstein RE. Inner retinal circadian clocks and non-visual photoreceptors: novel
players in the circadian system. Prog Neurobiol. 2010;92:484–504.
14. Hardeland R, Cardinali DP, Srinivasan V, Spence DW, Brown GM, Pandi-Perumal

SR. Melatonin – a pleiotropic, orchestrating regulator molecule. Prog Neurobiol.
2011;93:350–84.
15. Wood S, Loudon A. Clocks for all seasons: unwinding the roles and mechanisms of circadian
and interval timers in the hypothalamus and pituitary. J Endocrinol. 2016;228:X1.
16. Paakkonen T, Leppaluoto J, Makinen TM, Rintamaki H, Ruokonen A, Hassi J, Palinkas
LA. Seasonal levels of melatonin, thyroid hormones, mood, and cognition near the Arctic
Circle. Aviat Space Environ Med. 2008;79:695–9.
17. Levitas E, Lunenfeld E, Weisz N, Friger M, Har-Vardi I. Seasonal variations of human sperm
cells among 6455 semen samples: a plausible explanation of a seasonal birth pattern. Am
J Obstet Gynecol. 2013;208:406.
18. Jimenez-Ortega V, Cano P, Pagano ES, Fernández-Mateos P, Esquifino AI, Cardinali

DP. Melatonin supplementation decreases prolactin synthesis and release in rat adenohy-
pophysis. Correlation with anterior pituitary redox state and circadian clock mechanisms.
Chronobiol Int. 2012;29:1021–35.
19. Melrose S. Seasonal affective disorder: an overview of a assessment and treatment approaches.
Depress Res Treat. 2015;2015:178564.
20. Saper CB. The neurobiology of sleep. Continuum (Minneap Minn). 2013;19:19–31.
21. Borbely AA, Daan S, Wirz-Justice A, Deboer T. The two-process model of sleep regulation: a
reappraisal. J Sleep Res. 2016;25:131–43.
22. Kleitman N. Basic rest-activity cycle – 22 years later. Sleep. 1982;5:311–7.
23. Garay A, Cardinali DP. New concepts in the neurophysiology of sleep and wakefulness.
Physiol Minirev. 2016;9:26–36.
24. Van Diepen HC, Lucassen EA, Yasenkov R, Groenen I, Ijzerman AP, Meijer JH, Deboer
T. Caffeine increases light responsiveness of the mouse circadian pacemaker. Eur J Neurosci.
2014;40:3504–11.
25. Bremer F. Preoptic hypnogenic area and reticular activating system. Arch Ital Biol.

1973;111:85–111.
26. Moruzzi G, Magoun HW. Brain stem reticular formation and activation of the

EEG. Electroencephalogr Clin Neurophysiol. 1949;1:455–73.
27. Reid AH, McCall S, Henry JM, Taubenberger JK. Experimenting on the past: the enigma of
von Economo’s encephalitis lethargica. J Neuropathol Exp Neurol. 2001;60:663–70.
28. Kuwaki T. Thermoregulation under pressure: a role for orexin neurons. Temperature (Austin).
2015;2:379–91.
29. Peever J, Fuller PM. Neuroscience: a distributed neural network controls REM sleep. Curr
Biol. 2016;26:R34–5.
56 2  The Timed Autonomic Nervous System

30. De Andres I, Garzon M, Reinoso-Suarez F. Functional anatomy of non-REM sleep. Front


Neurol. 2011;2:70.
31. Nishimura H. Renin-angiotensin system in vertebrates: phylogenetic view of structure and
function. Anat Sci Int. 2017;92:215–47.
32. Pandi-Perumal SR, Cardinali DP, Chrousos G, editors. Neuroimmunology of sleep. New York:
Springer Science+Business Media, LLC; 2007.
33. Krueger JM. The role of cytokines in sleep regulation. Curr Pharm Des. 2008;14:3408–16.
34. Geiger SS, Fagundes CT, Siegel RM. Chrono-immunology: progress and challenges in under-
standing links between the circadian and immune systems. Immunology. 2015;146:349–58.
35. Karrer M, Lopez MA, Meier D, Mikhail C, Ogunshola OO, Muller AF, Strauss L, Tafti M,
Fontana A. Cytokine-induced sleep: neurons respond to TNF with production of chemokines
and increased expression of Homer1a in vitro. Brain Behav Immun. 2015;47:186–92.
36. Mader EC Jr, Mader AC. Sleep as spatiotemporal integration of biological processes that
evolved to periodically reinforce neurodynamic and metabolic homeostasis: the 2m3d para-
digm of sleep. J Neurol Sci. 2016;367:63–80.
37. Stickgold R. Sleep-dependent memory consolidation. Nature. 2005;437:1272–8.
38. Hobson JA. REM sleep and dreaming: towards a theory of protoconsciousness. Nat Rev
Neurosci. 2009;10:803–13.
39. Pace-Schott EF, Hobson JA. The neurobiology of sleep: genetics, cellular physiology and sub-
cortical networks. Nat Neurosci Rev. 2002;3:591–605.
40. Strata P. The emotional cerebellum. Cerebellum. 2015;14:570–7.
41. Goadsby PJ. Chapter 16 – Autonomic nervous system control of the cerebral circulation. In:
Buijs RM, Swabb D, editors. Handbook of clinical neurology. Autonomic nervous system.
Elsevier; 2013, p. 193–201.
42. Jessen NA, Munk AS, Lundgaard I, Nedergaard M. The glymphatic system: a beginner’s
guide. Neurochem Res. 2015;40:2583–99.
43. Venkat P, Chopp M, Chen J. New insights into coupling and uncoupling of cerebral blood flow
and metabolism in the brain. Croat Med J. 2016;57:223–8.
44. Tang G, Yang GY. Aquaporin-4: a potential therapeutic target for cerebral edema. Int J Mol
Sci. 2016;17.
First Level: Peripheral Sympathetic
and Parasympathetic Nervous System 3

Abstract
The periphery of the autonomic nervous system (ANS) comprises two parts: the
sympathetic and the parasympathetic systems. The identity and functional feature
of the ANS innervating the gastrointestinal is often considered the third section of
the ANS, i.e., the enteric ANS. This Chapter discusses the neurochemical bases of
the peripheral motor and sensory components of the ANS, including the neurotrans-
mitters and receptors involved and their participation in spinal autonomic reflexes.
The composition and functions of the enteric nervous system are also discussed.

Keywords
Adrenergic neurotransmission • Cholinergic neurotransmission • Co-transmission
• Enteric nervous system • Ionotropic transmission • Metabotropic transmission
• Parasympathetic nervous system • Prevertebral and paravertebral sympathetic
ganglia • Sensory autonomic neurons • Spinal autonomic reflexes • Sympathetic
nervous system

Objectives
After studying this chapter, you should be able to:
• Underline why cholinergic and adrenergic neurotransmission and peptide-
rgic co-transmission are the basis of the peripheral motor constituents of
the autonomic nervous system (ANS).
• Describe the location of the cell bodies and axonal trajectories of pregan-
glionic and postganglionic sympathetic and parasympathetic neurons.
• Name the neurotransmitters that are released by preganglionic autonomic
neurons, postganglionic sympathetic neurons, postganglionic parasympa-
thetic neurons, and adrenal medullary cells.
• Name the types of receptors on autonomic ganglia and on various target
organs and the processes involved in neurotransmission within the ANS.

© Springer International Publishing AG 2018 57


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1_3
58 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

• Define sensory autonomic neurons, their anatomy, and how the basic sen-
sorial dimensions are coded.
• Describe the homologies and differences between spinal motor reflexes
and spinal autonomic reflexes.
• Describe the anatomical and physiological basis of genitourinary reflexes.
• Describe the anatomical and physiological basis of defecation.
• Describe the anatomical and physiological basis of pupillary reflexes.
• Describe the composition and functions of the enteric nervous system.
• Understand the meaning of local autonomic projections in neuroendocrine
communication.
• Define how the ANS contributes to the maintenance of healthy bone
tissue.

 he Organization of the ANS at a Peripheral Level Comprises


T
Two Neurons in a Series

The periphery of the autonomic nervous system (ANS) comprises two parts: the
sympathetic and the parasympathetic systems (Figs. 3.1 and 3.2). The identity and
functional features of the ANS innervating the gastrointestinal is often considered
the third section of the ANS, i.e., the enteric ANS. Conceptually, however, the
regulatory mechanisms at the digestive level, although complex, do not escape the
general characteristics of the sympathetic and parasympathetic systems discussed
below [1].
Moreover, the ANS has also been conceptualized as having five components: the
sympathetic noradrenergic system, the sympathetic cholinergic system, the para-
sympathetic cholinergic system, the sympathetic adrenergic system, and the enteric
nervous system [2]. The reason for the differential noradrenergic vs adrenergic
responses is that sympathetic noradrenergic system activation dominates the
responses to orthostasis, moderate exercise, and exposure to cold, whereas sympa-
thetic adrenergic system activation dominates responses to glucoprivation and emo-
tional distress [2].
Effectors of the sympathetic system are generally the smooth muscle, the heart,
exocrine and endocrine glands, the adipose tissue, liver, lymphohematopoietic
organs, and kidney. Most sympathetic ganglia are remotely located to the innervated
organ, and therefore postganglionic axons (i.e., the axons of ganglion neurons) are
long (Fig. 3.3). An exception is the short adrenergic neurons located on the wall of
the genitourinary organs, which have a similar pattern to the parasympathetic array
[3, 4].
Prevertebral ganglia (celiac, superior mesenteric, inferior mesenteric) give rise to
postganglionic fibers, which through plexuses or nerves innervate organs of the
abdominal and pelvic regions. The sympathetic preganglionic neurons have their
bodies in the intermediolateral column of the spinal cord, between segments T1 and
The Organization of the ANS at a Peripheral Level Comprises Two Neurons in a Series 59

Fig. 3.1  The sympathetic


branch of the autonomic
nervous system. Modified
with permission from
Cardinali [1] Ciliary muscle
Lacrimal gland
Salivary glands

Larynx and 1
trachea 2
3
Lungs 4
5
Celiac ganglion
Heart 6
7
Stomach 8
Small intestine 9
10
Adrenal medulla
11
12
Super mesenteric 1
2
ganglion 3
Large
intestine
Kidney
Bladder
Sexual Inferior
organs mesenteric
ganglion

L2, so that the sympathetic system is also called thoracolumbar (Fig. 3.1). The pre-
ganglionic fibers exit through the corresponding anterior roots and follow the white
communicating branches to the paravertebral sympathetic ganglion chain, from
where four pathways can follow (Fig. 3.4) [3, 4]:

1. To synapse with ganglion neurons of the corresponding level or higher or lower


levels. The postganglionic fibers of these ganglion neurons leave the ganglion
through the gray communicating branch and join the somatic peripheral nerves
to innervate blood vessels, exocrine or endocrine glands, immune tissue, and
piloerector muscles in their territories. The postganglionic fibers present a vari-
able segmental distribution, much more poorly delimited than that of the sensory
and motor fibers.
60 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Fig. 3.2 The Ciliary ganglion


parasympathetic branch
Sphenopalatine Ciliary muscle
of the autonomic nervous ganglion III pair
system. Modified with
Submandibular Lacrimal
permission from VII pair
ganglion glands
Cardinali [1]
Salivary
glands
Parotid
gland
Otic ganglion
Larynx 1
and trachea 2
3
Lungs
4
Heart
5
6
Gastrointestinal tract
7
Abdominal 8
organs X pair 9
10
11
Large 12
1
intestine 2
3 S2-S4

Kidney
Bladder
Sexual
organs

2. The superior thoracic roots form synapses in the cervical ganglia, whose post-
ganglionic fibers are distributed to innervate cranial structures (superior cervical
ganglion, SCG) or to innervate thoracic organs via visceral, cardiac, and pulmo-
nary nerves, by innervating these thoracic organs (middle cervical ganglion, stel-
late ganglion).
3. The preganglionic fibers that emerge from the spinal roots below the diaphragm
pass through the paravertebral sympathetic chain and they give rise to the
splanchnic nerves, to end up forming synapses in the prevertebral ganglia. From
these, the postganglionic fibers form small visceral nerves that, in the form of
plexuses, are distributed to the abdominal and pelvic viscera.
4. Finally, some preganglionic fibers also follow this path to form synapses directly
with the chromaffin cells of the adrenal medulla, which represent the paraverte-
bral sympathetic ganglia corresponding to the postganglionic cells.

These neurons form a bilateral chain parallel to the vertebral column, between
the base of the skull and the coccyx. Normally, the anatomical distribution of each
ganglionic chain consists of about 24 ganglia. In the cervical region, there are three
The Organization of the ANS at a Peripheral Level Comprises Two Neurons in a Series 61

Fig. 3.3 Adrenergic Sympathetic Parasympathetic


and cholinergic
neurotransmission in the
Central Nervous System
autonomic nervous system.
Modified with permission Preganglionic neuron
from Cardinali [1]

Autonomic ganglion

ACh (nicotinic)

ACh (nicotinic)

Postganglionic
neuron

Autonomic ganglion

NE ACh (muscarinic)
EFFECTOR CELL

ganglia, superior, middle, and inferior, although the latter is often fused with the
first thoracic ganglion forming the stellate ganglion. In the other regions, there are
usually a pair of ganglia for each spinal metamere, 10 or 11 thoracic ganglia, 4
lumbar, 3 or 4 sacral, and a single coccygeal.
The sympathetic prevertebral ganglia are located around the branches of the
abdominal aorta, forming the abdominal or solar plexus. The main prevertebral gan-
glia are celiac, superior mesenteric, inferior mesenteric, and aortorenal.
The preganglionic neurons of the parasympathetic system have their perikarya in
nuclei of the brainstem or in the lateral column of the sacral spinal cord, which is
why it receives the designation of craniosacral division (Fig. 3.2).
In the cranial portion the preganglionic neurons are in the visceral efferent nuclei,
from where they distribute their fibers peripherally via the cranial nerves [5]:

• Edinger–Westphal nucleus: common ocular motor nerve (III).


• Superior and lacrimal salivary nucleus: facial nerve (VII).
• Lower salivary nucleus: glossopharyngeal nerve (IX).
• Dorsal nucleus of the vagus and ambiguous nucleus: vagus nerve (X).

These fibers are distributed to terminal ganglia, close to the effector organs, from
which short postganglionic fibers are formed, giving rise to the innervation of the
viscera of the head and neck, the thoracic cavity, and much of the abdominal
cavity.
62 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Autonomic reflexes
(efferent pathways)

Spinal nerve

White ramus
communicans
(preganglionic)
Gray ramus communicans
(postganglionic) Paravertebral
ganglion
Autonomic motor axons
(vasomotor, pilomotor,
sudomotor)
Splanchnic nerve

Prevertebral
ganglion

Postganglionic innervation Adrenal medulla


(blood vessels, viscera,
endocrine glands, immune
system)

Fig. 3.4  Neurons of the autonomic reflex, integrated at the level of the intermediolateral column
of the spinal cord. The visceral afferents enter through the spinal nerve. The sympathetic and pre-
vertebral ganglia and the gray and white communicating branches are also shown. Modified with
permission from Cardinali [1]

On the other hand, the preganglionic neurons of the sacral portion of the spinal
cord, located in the intermediolateral columns between the S2 and S4 segments,
send their fibers through the anterior roots to form the pelvic nerves and end in the
terminal ganglia of the pelvic plexus, innervating the descending colon and the uro-
genital organs [6].
The parasympathetic terminal ganglia are formed by small clusters of neurons,
located on or in the walls of the viscera, where the preganglionic fibers form syn-
apses. In the cephalic region, there are four relatively large ganglia, associated with
branches of the trigeminal nerve: the ciliary, sphenopalatine, optic, and subman-
dibular ganglia.
In the cervical, thoracic, and pelvic regions, there are small ganglia that form
visceral plexuses, whereas in the digestive tract they are in the myenteric plexus of
Auerbach and the submucosal plexus of Meissner.
Cholinergic and Adrenergic Neurotransmission and Peptidergic Co-transmission 63

The cell bodies of the parasympathetic preganglionic neurons are located in the
brainstem and sacral cord. Although some of their axons are myelinated, most are
unmyelinated. Compared with the sympathetic system, the preganglionic axons are
longer and their postganglionic fibers considerably shorter. This is because the para-
sympathetic ganglia are located in the immediate vicinity of the innervated organs [1].
Effectors of the parasympathetic system include smooth muscle and glands of
the digestive tract, the excretory organs, the genital system, heart, and lung, lym-
phohematopoietic and endocrine organs and intraocular muscles. Except genital
arteries and possibly the brain, there is no parasympathetic innervation of vascular
smooth muscle. Nor is there a parasympathetic innervation of the skin [1]. By using
different experimental models of autonomic hyper- or hypofunction, the conclu-
sions listed in Table 3.1 can be reached [7].

 holinergic and Adrenergic Neurotransmission


C
and Peptidergic Co-transmission Are the Basis
of the Peripheral Motor Constituents of the ANS

Transmitters identified, partially or totally, in neural pathways comprise five large


families [1]:

• Biogenic amines: norepinehrine (NE), epinephrine (E), acetylcholine (ACh),


5-HT, dopamine (DA), His
• Amino acids: Glu, aspartate, GABA, Glycine (Gly), taurine and purine deriva-
tives (adenosine, ATP, included in this group although they are not amino acids).
• Neuropeptides (over 70 different structures identified to date)
• Gases, such as NO or CO (carbon monoxide)
• Lipids, such as anandamide.

The criteria for a substance to be considered a neurotransmitter are as follows


(Fig. 3.5):

• It must be synthesized by the presynaptic neuron and stored in synaptic vesicles


(although exceptions are found, like the gases and lipids)
• The physiological neural stimulation must release it
• It must act on the postsynapsis in a similar manner to normal stimulus analyzed
(action identity criterion)
• There should be effective mechanisms for terminating its action (reuptake in
neural terminal, metabolism, passage to extrasynaptic space), to ensure the nec-
essary speed and transience of transmitter action

Among them, the action identity criterion is the most important, as it involves the
physiological effect itself [1].
Listed criteria have been completely fulfilled in the peripheral ANS, for the neu-
rotransmitter nature of ACh or NE (various preganglionic and postganglionic
64 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Table 3.1  Autonomic nervous system effects on effector organs


Parasympathetic
Organ, tissue stimulation Sympathetic stimulation
Pupils
Radial muscle of the iris – Contraction, α1 adrenoceptor mediated
Sphincter muscle Contraction, M3, –
of the iris M2-receptor mediated
Ciliary muscle Contraction, M3, Weak relaxation, β2 receptor mediated
M2-receptor mediated
Tracheobronchial Contraction, M2, Relaxation, β2 adrenoceptor mediated
muscle M3-receptor mediated
Heart Bradycardia, negative Tachycardia, positive inotropism, β1
inotropism, M2 adrenoceptor mediated
receptor mediated
Arteries
Skin, mucosas – Constriction, α1 adrenoceptor mediated
Abdomen – Constriction, α1 adrenoceptor mediated
Dilatation (blood epinephrine), β2
adrenoceptor mediated
Skeletal muscle – Dilation, M2 receptor mediated
Dilatation (blood epinephrine), β2
adrenoceptor mediated
Coronary arteries – Constriction, α1 adrenoceptor mediated
Dilation, β2 adrenoceptor mediated
CNS arteries Dilation (?) Constriction, α1 adrenoceptor mediated
Penis Dilation, M2 receptor –
mediated
Kidney – Constriction, α1 adrenoceptor mediated
Veins – Constriction, α1 adrenoceptor mediated
Endothelium Stimulation of NO
synthase, M3 receptor
mediated
Bladder
Detrusor muscle Contraction, M3 Relaxation, β2 adrenoceptor mediated
receptor mediated
Internal sphincter Relaxation, M2 Constriction, α1 adrenoceptor mediated
receptor mediated
Genital tract
Seminal vesicles – Constriction, α1 adrenoceptor mediated
Vas deferens – Constriction, α1 adrenoceptor mediated
Intercourse Erection, M3 receptor Ejaculation, α-1 adrenoceptor mediated
mediated
Uterus – Constriction, α1 adrenoceptor mediated
Relaxation, β2 adrenoceptor mediated
Gastrointestinal tract
Circular and Contraction, M3 Relaxation (direct) β adrenoceptor mediated
longitudinal muscle receptor mediated Parasympathetic inhibition, α1 adrenoceptor
mediated
Sphincters Relaxation Constriction, α adrenoceptor mediated
Exocrine glands
Salivary Secretion. M3
receptor mediated
Cholinergic and Adrenergic Neurotransmission and Peptidergic Co-transmission 65

Table 3.1 (continued)
Parasympathetic
Organ, tissue stimulation Sympathetic stimulation
Lacrimal Secretion. M3 –
receptor mediated
Digestive Secretion. M3 Inhibition, α adrenoceptor mediated
receptor mediated
Nasopharyngeal Secretion. M3 –
receptor mediated
Sweat – Secretion. M3 receptor mediated
Bronchial Secretion. M3 Weak inhibition, α-1 adrenoceptor mediated.
receptor mediated Weak stimulation, β adrenoceptor mediated
Adipose tissue – Lipolysis, β1 adrenoceptor mediated
Immune system
Lymph nodes, spleen Stimulation Inhibition, α-1 adrenoceptor mediated
Splenic capsule – Constriction, α-1 adrenoceptor mediated
Endocrine system
Juxtaglomerular – Renin secretion, β1 adrenoceptor mediated
apparatus
Thyroid follicles Thyroid growth, Inhibition of T4 secretion, α1 mediated.
secretion of T4. M Weak stimulus of T4 secretion, β
receptor mediated adrenoceptor mediated
Thyroid C cells Inhibition of CT Inhibition of CT secretion, α1 mediated.
secretion Weak stimulus of CT secretion, β
adrenoceptor mediated
Parathyroid glands Inhibition of PTH Inhibition of PTH secretion
secretion
Pineal gland – Melatonin release, β2 adrenoceptor mediated.
Weak stimulus of melatonin release, α1
adrenoceptor mediated
Adrenal medulla Preganglionic stimulation, M1 receptor
mediated
Anterior pituitary Postganglionic SCG sympathetic neurons
inhibit FSH, LH, GH, PRL, and TSH
secretion and augments ACTH secretion
Posterior hypophysis – Postganglionic SCG sympathetic neurons
inhibit ADH secretion, β adrenoceptor
mediated
Exocrine pancreatic Secretion. M1 Inhibition, α adrenoceptor mediated
glands receptor mediated
Pancreatic islets, α cells Secretion. M1 Inhibition of glucagon secretion, α2
receptor mediated adrenoceptor mediated. Weak stimulation of
glucagon secretion, β2 adrenoceptor mediated
Pancreatic islets, β cells Secretion. M1 Inhibition of insulin secretion, α-2
receptor mediated adrenoceptor mediated. Weak stimulation of
insulin secretion, β2 adrenoceptor mediated
Liver Glycogen synthesis Glycogenolysis, α1 adrenoceptor mediated
(?) Gluconeogenesis, β2 adrenoceptor mediated
Lipolysis, β1 adrenoceptor mediated
NO nitric oxide, CT calcitonin, PTH parathyroid hormone, SCG superior cervical ganglion, FSH
follicle-stimulating hormone, LH luteinizing hormone, GH growth hormone, PRL prolactin, TSH
thyroid-stimulating hormone, ACTH adrenocorticotropic hormone, ADH antidiuretic hormone
66 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

a
Resting Synapse

Cleft
Presynapse Postsynapse

b
Active Synapse
Action potential Action potential

Electrical secretory potential


Effector activity
events

Ca2+ Transmitter
Active zone diffusion Nerve Action potential and
transmitter release
Muscle
Contraction,
relaxation
Gland
Secretion
Chemical Transmitter Transmitter
events release interaction with post- Triggered function
synaptic receptors

Fig. 3.5  Chemical synapse. (a) In the resting synapse the pre- and postsynaptic membranes are
normally polarized. (b) In the active synapse, depolarization of the neural terminal (secretory
potential) results in the release of the transmitter, which diffuses through the synaptic gap and
produces local synaptic currents and potentials in the postsynaptic membrane, which initiate effec-
tor activity, neuronal transmission, neurotransmitter release, hormonal secretion, muscle contrac-
tion). Modified with permission from Cardinali [1]

territories). On the other hand, in central ANS neurons, the physiological character-
ization of neurotransmitters has been difficult, owing to the large number of syn-
apses present, to the neuronal plurality of most brain regions, and to the very
common coexistence of two or more neurotransmitters in the same synapse.
Biogenic amines participate in approximately 5% of the brain synapses, localizing
in certain subcortical projection pathways to encephalic or descending spinal cord
regions. The noradrenergic and serotonergic cortical, cerebellar, and subcortical
innervation originates, almost exclusively, in brainstem nuclei that project in the form
of a “spider web” to large cerebral areas, an anatomical disposition that speaks of its
general modulatory function (Chap. 1). Something similar occurs for the central dopa-
minergic, cholinergic, and histaminergic systems. It is not surprising, then, that these
monoaminergic systems have been linked to generalized alterations in brain function,
such as psychiatric illness, emotionality, wakefulness, and sleep (Chap. 4).
Cholinergic and Adrenergic Neurotransmission and Peptidergic Co-transmission 67

In the ANS, ACh is the transmitter of all the preganglionic synapses, of all para-
sympathetic postganglionic neurons, and of some sympathetic postganglionic neu-
rons (muscular vasodilator system, sweat glands). NE is the neurotransmitter of all
the remaining postganglionic sympathetic neurons [3, 4].
For most chemical synapses, there is an exocytotic release from the presynapsis
of the neurotransmitter substance contained in the synaptic vesicles. Exceptions to
this rule are gases identified as neurotransmitters (NO, CO), which pass through
membranes by simple diffusion, and neurotransmitter lipids such as anandamide,
which are not stored in the vesicles.
In the case of NO, which is one of the most important local regulators of BP,
three different NO synthase (NOS) isoforms have been described, two constitu-
tively present (eNOS and nNOS), and one inducible (iNOS). In the CNS, both
eNOS and nNOS are present. Centrally, NO functions mainly as a neuromodulator,
having both sympathoinhibitory and sympathoexcitatory actions [8]. At the periph-
ery, locally released NO from endothelial cells acts on adjacent vascular smooth
muscle cells to produce vasodilatation. In addition, NO synthase activity has been
demonstrated in preganglionic autonomic fibers innervating vascular smooth mus-
cle. Furthermore, NO release from autonomic nitrergic nerves interferes with the
release of NE. Both central and peripheral effects of NO under normal conditions
are masked by the baroreflex [9] (Chap. 4).
In general, the action of these signals is exerted at the level of specific receptors
in the postsynapsis. The gases are again the exception as they traverse the postsyn-
aptic membrane and exert their action intracellularly. In the chemical synapse the
synaptic message is unidirectional (it ranges from pre- to postsynapsis) and when it
comes from synapses by exocytotic release of neurotransmitters present in vesicles,
it shows a synaptic delay. This synaptic delay is largely due to the transmitter release
process, and to a minimal extent by the passage of the transmitter through the syn-
aptic gap. Its duration is about 0.5–1.0 ms [10].
In the ANS, the synaptic cleft is broad (synaptic varicosities or nondirected syn-
apses) and thus differs from directed synapses, such as the neuromuscular plaque
(Fig. 3.6). This allows the “transmission by volume” into the ANS, a name that is
given to the wide diffusion of the autonomic transmitter to several postsynaptic cells.
Autonomic postganglionic fibers present a series of synaptic dilations or vari-
cosities when they approach their targets, which have a length of approximately
1 μm and a diameter of between 0.5 and 2 μm. These varicosities contain a high
density of mitochondria and synaptic vesicles. The synaptic space has a variable
width; for example, in the innervation of the smooth muscle, it varies from 20 nm in
the vas deferens to 1–2 μm in the large arteries. Transmitter release occurs in a pas-
sage, and propagates nerve impulses along the autonomous axon. The availability of
varicosities and the wide synaptic space allow the released neurotransmitter to dif-
fuse variable distances in the target organ and to activate multiple receptors, which
expand the effect of autonomic activity (Fig. 3.6) [3, 4, 11].
This classic view of autonomic transmission is, however, a simplification, as it is
now known that many of these neurons also use other types of molecules as neu-
rotransmitters or as neuromodulators. Initially, these were encompassed under the
68 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Preganglionic
neurons

Micro-
tubule

NE
DA-β-hydroxylase
ATP
chromogranin

Varicosities

Nerve terminals

Immune cells
Connective tissue

Smooth muscle
Endocrine and
exocrine glands

Cardiac
muscle
Secretory epithelium

Fig. 3.6  Postganglionic noradrenergic neuron with its distribution in varicosities (“beads”).
Modified with permission from Cardinali [1]
Cholinergic and Adrenergic Neurotransmission and Peptidergic Co-transmission 69

denomination of non-adrenergic–noncholinergic transmission, which mainly


includes purinergic transmitters and neuropeptides. Purinergic nucleotides, such as
ATP, are important transmitters in nerve fibers that innervate the intestinal smooth
muscle, the urinary bladder, and the vas deferens, where they act directly as neu-
rotransmitters or modulate the effects of NE or ACh. Several neuropeptides have
been identified, widely distributed, in autonomic ganglia, the enteric system, and
peripheral fibers. The most well-known include: enkephalin/endorphin, vasoactive
intestinal peptide (VIP), substance P, calcitonin gene-related peptide (CGRP), neu-
ropeptide-­Y (NPY), somatostatin, bombesin, galanin, neurotensin, angiotensin, and
cholecystokinin (CCK)/gastrin.
The coexistence of neuropeptides with classical transmitters in different auto-
nomic neurons is the rule, for example, sympathetic and parasympathetic postgan-
glionic neurons containing ACh and VIP, or adrenergic neurons containing NE and
NPY. The classical neurotransmitter and the neuropeptide may be separately
released under different excitation conditions (Fig. 3.7). It is postulated that neuro-
peptides might act as transmitters by themselves or as neuromodulators, altering the
action of classical transmitters. Thus, stimulation of cholinergic fibers from the sali-
vary or sweat glands causes release of ACh, which has a secretory effect, and VIP,
which produces vasodilation [3].
The neuropeptides co-released with small molecule neurotransmitters in auto-
nomic nerves do not usually act as co-transmitters, but rather as prejunctional neu-
romodulators or trophic factors. Autonomic co-transmission offers subtle, local
variation in physiological control mechanisms, rather than the dominance of inflex-
ible central control mechanisms envisaged earlier [12]. The variety of information
imparted by a single neuron then greatly increases the sophistication and

Release

Biogenic amine Biogenic amine


+ neuropeptide

Neural activity

Fig. 3.7 Differential
release of transmitters
contained in small
dense-cored vesicles
(biogenic amine) and large
dense-cored vesicles
(neuropeptide). Modified
with permission from
Cardinali [1]
70 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Table 3.2 Neurochemical Nerve fibers Neurotransmitters


code in different types of
Vasoconstriction NE, NPY
autonomic nerve fibers
Vasodilation ACh, VIP
Sudomotor ACh, VIP, CGRP
Piloerector NE, dynorphin
Sialomotor CGRP, NPY
Intestinal ganglia NE, somatostatin
NE norepinephrine, NPY neuropeptide-­Y, ACh ace-
tylcholine, VIP vasoactive intestinal peptide, CGRP
calcitonin gene-related peptide

complexity of local control mechanisms. Co-transmitter composition shows con-


siderable plasticity in development and aging, in pathophysiological conditions,
and following trauma or surgery. For example, ATP appears to become a more
prominent co-transmitter in inflammatory and stress conditions [12]. In situations
such as congestive cardiac failure that are characterized by high levels of cardiac
sympathetic drive, sympathetic co-transmitters such as NPY can be released in
addition to NE. Even in the presence of β-adrenoceptor blockers, NPY is able to
bind to its own receptors located on the cholinergic ganglia and ventricular myo-
cytes, thus inhibiting ACh release during vagus nerve stimulation and limiting the
subsequent bradycardia [13].
Most experimental evidence has shown that the autonomic sympathetic and
parasympathetic control is organized in different functional units or groups of neu-
rons that innervate specific target organs. Thus, in the sympathetic ganglia, subtypes
of vasomotor, sudomotor, pilomotor, and visceromotor neurons are distinguished.
Ganglion neurons of each subtype present a unique chemical code (Table 3.2).
Today, we know that the rule is that the ANS releases a combination of neurotrans-
mitters whose proportion depends on the intensity and frequency of stimulation
(Fig. 3.7).
It is useful to pay attention to a differential aspect among the different families of
neurotransmitters, namely, the mechanism of synthesis [1]. Biogenic amines, amino
acids, gases, and lipids are synthesized by an enzymatic process at the synaptic
terminals. The specific enzymes migrate to the terminal via axoplasmic transport,
forming parts of vesicles, and catalyze at the terminal the synthesis of transmitter
from specific precursors. Because of its catalytic nature, an enzyme molecule par-
ticipates in the synthesis of thousands of molecules of the transmitter. This prevents
the rapid reduction of the contents of the transmitter.
In contrast, neuropeptides are synthesized in the neuronal body as part of a
higher molecular weight prepropeptide, which is incorporated into the synaptic
vesicles and is processed (by acetylation, glycosylation, or hydrolysis reactions)
while these vesicles migrate by axonal transport toward the neural terminal. It
should be noted that both the axon and the presynaptic terminals are almost devoid
of ribosomes, and hence of the ability to synthesize peptides or proteins (there is
only a restricted synthesis of some components of the cytoskeleton). In this, they
differ from the dendrites [3].
Cholinergic and Adrenergic Neurotransmission and Peptidergic Co-transmission 71

It is for this reason that after a prolonged neuronal stimulation there is a greater
possibility of neuropeptide transmitter depletion than that of any co-transmitter
(Fig. 3.7). The peptide transmitter deposits are depleted at terminals remote from
the neuronal body, the terminals at close range being maintained for longer. This is
the reason why, on certain occasions, terminals of the same neuron can release dif-
ferent combinations of transmitting substances [12]. Likewise, the coexistence of
neurotransmitters does not mean that they are released with the same kinetics before
presynaptic stimulation. As shown in Fig. 3.7, small vesicles with a dense center
with a biogenic amine (e.g., NE) are released before the large vesicles (e.g., NPY).
Generally, the opening or closing of channels at the postsynaptic membrane by
neurotransmitter in the autonomic synapses (Fig. 3.8) is produced by: (a) the direct
association of neurotransmitter with the postsynaptic receptor coupled to a channel
(ionotropic transmission); (b) by synthesizing intracellular second messengers, trig-
gered by the association of the autonomic transmitter with its receptor, this second
messenger being responsible for the modification of membrane conductance
(metabotropic transmission) [7].
An example of ionotropic transmission is the transmission given by the nicotinic
cholinergic receptor of the autonomic preganglionic synapse (Fig. 3.9). The recep-
tor forms a constituent part of an ion channel that permeates Na+ and to a lesser
extent K+, and that opens when ACh binds to the receptor.
An example of metabotropic transmission is the β action of NE, which occurs by
increasing cAMP and the subsequent phosphorylation of K+ channels to inactivate
them. The result of this action is the postsynaptic depolarization that follows the
increases in the concentration of intracellular K+ (Fig. 3.10).

Receptor Pore Pore


Neurotransmitter Receptor
Channel Channel
Effector Neurotransmitter
function ECF

Gate

ICF P
cAMP
Gate
GTP
Effector function
G protein Adenylate
cyclase
NH2 cAMP-dependent PKA
COOH
ECF

ICF
α γ α δ β
COOH

Fig. 3.8  Ionotropic and metabotropic effect of neurotransmitters. Modified with permission from
Cardinali [1]
72 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Fig. 3.9 Cholinergic Synaptic


nicotinic Terminal
neurotransmission.
Modified with permission
from Cardinali [1] Synaptic vesicle
ACh
ACh

Acetate + CoA ACh


+ Choline Calcium-
dependent
ACh release

Choline
Acetylcholinesterase Nicotinic
cholinergic
receptors

Adenylate cyclase
Norepinephrine Plasma
1 membrane
Receptor

b a a
a g
3
2 ATP 6
G protein cAMP 4
Ionophore
5
Prot. kinase Prot. kinase
(inactive) (active)
G protein subunit
dissociation
Protein Ionophore opening (6)
phosphorylation

Fig. 3.10  Sequence of events produced by the interaction of NE with β-adrenoceptors. See text.
Modified with permission from Cardinali [1]

Autonomic ionotropic transmission is rapid, whereas metabotropic transmis-


sion is slower. This fact is due to the time of synthesis and intracellular transloca-
tion of the second messenger synthesized from the cell membrane to its site of
action.
The molecular biological studies on various neurotransmitter receptors indicate
the existence of two structural superfamilies, which correspond to the two types of
transmission (ionotropic and metabotropic) analyzed above [7]:
Cholinergic and Adrenergic Neurotransmission and Peptidergic Co-transmission 73

• Receptors associated with ionophores, whose quaternary structure is part of an


ion channel. This is the case of the nicotinic cholinergic receptor, which com-
prises the following subtypes: muscle, ganglion, neuronal, neuronal CNS, neuro-
nal α7 (more recently they have been molecularly defined as α1*-, α2*-, α3*-,
α4*-, α6*-, α7*-, α8*-, and α9*-acceptors, nACh). This is also the case for the
GABA receptor types A and C, the glycinergic receptor, the AMPA, kainate and
N-methyl-d-aspartate (NMDA) receptors for Glu, the serotonergic receptor
5-HT3 subtype, and the ionotropic purinergic receptors. The structure consists of
highly homologous protein sequence subunits that derive from a common gene.
Each of the subunits contains 15–20 amino acid hydrophobic regions that consti-
tute alpha-helices, which completely cross the plasma membrane.
• Receptors associated with G-proteins: the receptor structure constituting a single
subunit with seven hydrophobic protein portions spanning the cell membrane.
Each of, the α-helices has 15–20 amino acids, which are not arranged in the form
of an ionic channel as in the previous case. The ion channel is located distant to the
receptor and is not part of it; it can instead be used by several receptors (Fig. 3.10).
• This is a very large family of receptors and comprises, among others, α1 adren-
ergic receptors (whose subtypes are identified as α1A, α1B, and α1D), α2 (sub-
types α2A, α2B, α2C), β (subtypes β1, β2, β3, and β4); muscarinic cholinergic
receptors (subtypes: M1, M2, M3, M4, and M5); dopaminergic receptors (sub-
types D1, D2, D3, D4, and D5); serotonergic (5-HT1A, 5-HT1A, 5-HT2A,
5-HT2B, 5-HT2C, 5-HT4A, 5-HT4A, 5-HT5A, 5-HT6, 5-HT7) subtypes; hista-
minergic (subtypes H1, H2, H3, H4); metabotropic Glu receptors (mGlu1 to
mGlu8 subtypes), GABA type B receptor, melatonin receptor (MT1 MT2 sub-
types), conopsins, and receptors of different types of neuropeptides.

There is intense “cross-talking” between the second messenger pathways, as in


many cases individual enzymes, channels or cytoskeletal proteins can be modified
at more than one site of the molecule by different systems of second intracellular
messengers. It is estimated that the metabotropic pathway is about 10,000 times
slower than the ionotropic pathway [11].
The activation of G proteins does not always lead to the synthesis of a second
messenger mediating the effect. Effects of the G proteins are directly exerted on
channels, without the participation of the phosphorylation. For example, the hyper-
polarization produced by ACh in the heart is caused, in a first-stage effect, by a G
protein that opens K+ channels without the mediation of a second messenger. There
are other similar actions described on Ca2+ channels. Finally, there are situations in
which the channel is quickly affected by the G protein, and much later modified by
the second messenger triggered by the same G protein (this is the case of the mus-
carinic M2 receptor type in the heart) [7].
In addition to the early and rapid effects of autonomic neurotransmitters, there
are other late effects on neuronal function. In general, the action of the transmitter
triggers modifications in gene expression, with lasting changes in cellular function
until long after the synaptic action is over. One of the first known late effects was the
trans-synaptic induction of tyrosine hydroxylase, the limiting enzyme in NE
74 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

synthesis, in postganglionic sympathetic neurons because of nicotinic presynaptic


ganglionic activation.
Regarding ACh synthesis, it comprises the acetylation of choline catalyzed by
the enzyme choline acetyltransferase. ACh metabolism is accomplished in the syn-
aptic gap and involves the reversal of the synthesis reaction, i.e., a hydrolysis cata-
lyzed by the enzyme acetylcholinesterase. In muscarinic cholinergic synapses this
hydrolysis is rapid enough to ensure complete inhibition of the synaptic effect of
ACh, a fact that does not occur in nicotinic synapses, such as the preganglionic one.
There are two types of receptors for ACh: muscarinic and nicotinic.
Muscarinic cholinergic receptors mediate the effects of ACh in postganglionic
synapses on smooth muscle, endocrine or exocrine glands, immune cells, and the
heart. They also mediate some of the effects of ACh in the CNS and autonomic
ganglia. Five types of muscarinic receptors (associated with G protein) are known
to exist [11]:

• M1: associated with a decrease in K+ conductance (and therefore excitatory),


present in the cerebral cortex and autonomic ganglia.
• M2: associated with increased K+ conductance (and therefore it is inhibitory),
present in the heart, and presynaptically on several parasympathetic territories.
• M3 and M5: similar to M1, they are present in smooth muscle and glandular
cells.
• M4: similar to M2.

As in the case of other G protein-associated receptors, the action of cholinergic


agonists can be exerted through two mechanisms, one quick, direct, via the protein
G effect on the channel, and a slower one through the second intracellular messen-
ger. M1, M3, and M5 receptors mediate cholinergic responses accompanying post-
synaptic stimulation such as smooth muscle contraction of the bronchi, bowel, or
bladder. M2 and M4 receptors mediate inhibitory responses, bradycardia, and nega-
tive inotropism verifiable in the heart after vagal stimulation or inhibition of presyn-
aptic transmitter release.
Nicotinic cholinergic receptors are in the autonomic ganglia, CNS, and muscle
plate. In all cases, they are associated with ionotropic responses involving the open-
ing of Na + channels (which, although to a lesser extent, also permeate K+), and
consequently depolarization (Fig. 3.9). These channels are composed by pentamers
resulting from combinations of 17 subunits. Nicotinic receptors are broadly classi-
fied into two subtypes based on their primary sites of expression: muscle-type nico-
tinic receptors and neuronal-type nicotinic receptors. In both muscle-type and
neuronal-type receptors, the subunits are somewhat similar, especially in the hydro-
phobic regions. In the autonomic ganglia, hexamethonium is the specific nicotinic
blocker, whereas in the muscle plate, curare (or its active ingredient, tubocurarine)
is the blocker [7, 11].
In relation to NE, their synthesis is started by tyrosine hydroxylation to l-dopa
(l-dihydroxyphenylalanine) catalyzed by the enzyme tyrosine hydroxylase
(Fig.  3.11). This is followed by the decarboxylation of levodopa to DA and by
Cholinergic and Adrenergic Neurotransmission and Peptidergic Co-transmission 75

Sympathetic nerve terminal

Capillary
MAO
Tyrosine
Tyrosine-OH-ase
DA NA
Axon DOPA
DOPA decarboxylase Dopamine β
hydroxylase
COMT

Effector cell

Fig. 3.11  Norepinephrine (NE) synthesis. Tyrosine is converted to l-DOPA by tyrosine hydroxy-
lase. DOPA is then decarboxylated to DA by l-aromatic amino acid decarboxylase. Finally, DA is
β-hydroxylated to NE by DA β-hydroxylase. NE is metabolized by either monoamine oxidase
(MAO) or catechol-O-methyltransferase (COMT)

β hydroxylation of DA to NE. The limiting step in this sequence is the first one, i.e.,
the hydroxylation of tyrosine (Fig. 3.11).
Catabolism of NE is carried out by oxidative deamination by the action of mono-
amine oxidase enzyme (MAO), or by O methylation enzyme by catechol-O-methyl
transferase (COMT; Fig. 3.12). However, under normal conditions, the most wide-
spread form of termination of the action of NE is the presynaptic reuptake in an
intact, unmetabolized form. The time at which this reuptake mechanism becomes
saturated, metabolism by MAO or COMT ensues.
There are two main types of adrenergic receptors (or adrenoceptors): (a)
α-adrenoceptors; (b) β-adrenoceptors.
As in the case of ACh, these receptor types were identified using adrenergic
agonists and antagonists, and more recently, various subtypes have been cloned and
identified. They belong to the superfamily of receptors associated with G proteins.
The α-adrenergic effect is one that shows the following sequence of activity for
agonists: NE = E > > isoproterenol (a synthetic adrenergic agonist).
The β-adrenergic effect is one that shows the following sequence of activity for
agonists: isoproterenol > E = NE.
In turn, each type of adrenoceptor is subdivided into α1 (whose subtypes identi-
fied are α1A, α1B, and α1D), α2 (subtypes α2A, α2B, and α2C), and β1, β2, β3, and
β4 [7, 11].
The subcellular mechanism of action of the α1 adrenoceptor is the increase in
conductance to Ca2+ and the activation of the turnover of inositol phospholipids. The
subcellular mechanism of α2 adrenoceptor is the inhibition of adenylate cyclase.
76 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

OH OH

OH CH - CH2- NH - CH3 OH CH2 - CH2 - NH2

OH OH
Epinephrine Norepinephrine

COMT MAO COMT


OH OH

CH3O CH2- CH2- NH - NH - C3H OH CH - COOH CH3O CH2- CH2- NH2

OH OH OH
Metanephrine Dihydroxymandelic acid Normetanephrine

COMT

OH

CH3O CH - COOH

OH
Vanillylmandelic acid

Fig. 3.12  Metabolism of NE by monoamine oxidase (MAO) and catechol-O-methyltransferase


(COMT). Modified with permission from Cardinali [1]

The subcellular mechanism of action of β1, β2, β3, and β4 receptors is the activation
of adenylate cyclase; α2 adrenergic receptors were initially described as having a
presynaptic location and with negative modulation function of NE release. However,
in the CNS, there are also α2 adrenoceptors at a postsynaptic location. The β1, β2,
β3, and β4 adrenergic receptors have both pre- and postsynaptic localization.
The postsynaptic consequence of the mentioned autonomic ionotropic or
metabotropic mechanisms is:

• Depolarization, which can be reached by a spatial or temporal sum to trigger an


action potential. The depolarizing synaptic potential is called the excitatory post-
synaptic potential, or EPSP (Fig. 1.8).
• Hyperpolarization with decreased postsynaptic excitability. The hyperpolarizing
synaptic potential is called the inhibitory postsynaptic potential, or IPSP (Fig. 1.8).

The arrival of the action potential to the synaptic terminal produces its depolar-
ization (“secretory potential”). Ca2+ voltage channels opened by depolarization are
instrumental for a sharp rise in cytoplasmic Ca2+ concentration that causes mem-
brane fusion of synaptic vesicles with the cell membrane, the opening of synaptic
vesicles, and the exocytotic emptying of its contents into the synaptic cleft (Fig. 3.5).
As the emptying of each vesicle is total, the amount of autonomic transmitter
released in this manner is always a multiple of the unit concentration present in each
vesicle. This is called the quantum release of transmitter.
Cholinergic and Adrenergic Neurotransmission and Peptidergic Co-transmission 77

In view of the importance of Ca2+ in neurotransmission, it is not surprising that


the regulation of autonomic transmitter release is mainly performed at the level of
voltage-dependent Ca2+ channels at the synaptic terminal [1]. This regulation is of
two types:

• Intrinsic to the neuron, through changes in the membrane potential at rest follow-
ing the previous neural activity.
• Extrinsic to the neuron by signals originating outside the cell. These signals may
be the neurotransmitter itself or its precursors, another transmitter, postsynaptic
metabolites, or hormones.

Examples of extrinsic regulation are [1]:

• Self-regulation, given by the same transmitter, which by interacting with the pre-
synaptic autoreceptors that recognize it, modulates its own release. As examples,
in sympathetic noradrenergic synapses, the presynaptic α2 adrenergic receptors
are of inhibitory nature for NE release, whereas β presynaptic adrenergic recep-
tors are excitatory of NE release (Fig. 3.13).
• Trans-synaptic regulation. This involves signals released by the postsynapses as
a result of the action of the transmitter, which across the synaptic cleft, modify
transmitter release. An example is given by various arachidonic acid metabolites,
such as PGs, produced in the postsynapses by the action of NE and that are
inhibitory for transmitter release (Fig. 3.13).
• Heterosynaptic regulation mediated via receptors for different neurotransmitters
in the synaptic terminals. The regulation is exerted by nearby synapses that use a
different type of neurotransmitter. An example of this phenomenon is presynap-
tic inhibition in primary sensory neuron nociception (transmitter: substance P,
co-transmitter: neoendorphin) caused by enkephalinergic interneurons in the
dorsal horn of the spinal cord (Fig. 3.13)
• Hormonal regulation, based on the various central and peripheral neuroendo-
crine phenomena. For example, the increased plasma levels of estradiol released
by growing ovarian follicles cause the activation of neural systems that regulate
the release of gonadotropin-releasing hormone (GnRH) and consequently the
release of luteinizing hormone (LH; Fig. 3.13).

The phenomenon of neurotransmission is fleeting. This transience is achieved by


effective mechanisms of the termination of neurotransmitter action, involving one
of the following three mechanisms.
The first mechanism operable to biogenic amines and amino acids, is the active
reuptake process by the terminal and synaptic vesicles. The passage through the pre-
synaptic membrane is a process associated with a Na/K ATPase symporter, and is
independent of presynaptic receptors. This uptake mechanism, which allows the
reuse of most of the transmitter released, is predominant for NE, DA, and 5-HT; in
the case of ACh, the choline precursor underwent reuptake. Genes are known, and
the sequence has been elucidated, for various presynaptic transporters. Most of them
78 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Fig. 3.13  Modulation of a


transmitter release. (a) Stimulatory Neurotransmitter
autoreceptor
Self-regulation. (b)
Postsynaptic
Trans-synaptic regulation. receptors
(c) Heterosynaptic or
interneuronal regulation.
(d) Humoral regulation.
Modified with permission Inhibitory
Afferent neuron Effector cell
from Cardinali [1] autoreceptor

Trans-synaptic
modulator
b

Afferent neuron Effector cell


Neurotransmitter

Modulatory
c terminal Neuromodulator

Afferent neuron Effector


Neurotransmitter neuron

d Blood vessel Humoral


modulator
Neurotransmitter

Effector
cell

Afferent neuron

comprise proteins with 12 hydrophobic transmembrane portions of 15–20 amino


acids (α helices). Rapid regulation is achieved through changes in transporter plasma
membrane expression and by intrinsic transport activity, mediated via transporter
phosphorylation and the regulated associations of accessory proteins, including pro-
tein phosphatases and scaffolding proteins. Thus, carrier clearance capacity is likely
to involve multiple regulatory proteins that localize, stabilize, and vary the activation
state [14]. Interest in these carriers is that they are the sites of action of many drugs
Sensory Autonomic Neurons 79

such as antidepressants and cocaine. In addition, genetic studies indicated gene poly-
morphisms for monoamine transporters in various psychiatric disorders.
The second inactivation mechanism for transmitters is metabolism (Fig. 3.12).
This is the primary mechanism for ACh, which is inactivated by the acetylcholines-
terase present in the postsynaptic membrane. Because of this inactivation, choline is
produced and is taken up by the cholinergic terminal (Fig. 3.9).
The third mechanism of inactivation is the passage of transmitter from the syn-
aptic cleft into the extracellular fluid or general circulation. Although this process
occurs for all types of transmitter, it is the predominant one for neuropeptides, for
which no presynaptic reuptake exists and the metabolism by extracellular pepti-
dases is slow. The gases are also diluted by diffusion.
In Fig. 3.14, many current concepts of chemical transmission in the ANS are
summarized. They have evolved from the simple consideration of the synapse as
being mediated by a single transmitter, through a single postsynaptic receptor, to the
present one, which includes a multiplicity of signals (M1–M3) acting through vari-
ous pre- and postsynaptic receptors. The main interactions shown in the figure are:

• Inhibition of the release of a neuropeptide (M2) by a conventional transmitter


(e.g., ACh) through presynaptic action (Rp’1, Fig. 3.14).
• Interaction between transmitters (M1 and M2) at a postsynaptic receptor (R’b,
Fig. 3.14).
• Facilitation or inhibition of transmitter release or presynaptic electrical activity
by peptide (M3) through a presynaptic receptor (Rp”, Fig. 3.14).

Sensory Autonomic Neurons

According to a classical definition of senses [15], the following categories occur:


teloreceptive (vision and hearing), proprioceptive (limb position), exteroceptive
(touch, including temperature and pain), chemoreceptive (smell and taste), and
interoceptive (visceral) modalities. For modern neuroanatomy, the existence of an A
sensory system, including teloreception, exteroception/proprioception, and a B sen-
sory system, including interoception/nociception, further classify the senses [16].
The development of small-diameter interoceptive afferents originating from small
(B) cells is coordinated with the development of lamina I and II cells in the dorsal
horn, clearly differing from the large-diameter exteroceptive afferents originating
from large (A) cells that project to the deep dorsal horn and do not connect with
lamina I neurons.
The main sensory input to the ANS comes from small-diameter sensory fibers by
way of lamina I neurons in the superficial dorsal horn (B cells). This pathway sup-
ports organotopic homeostatic control of the body’s condition, but also human feel-
ings from the body, such as temperature, pain, itching, affective touch, muscle ache,
vascular flush, etc. Once homeostatic information from the tissues is decoded, it is
conveyed up to the anterior insula after making synaptic relays at different levels
(spinal cord – lamina I and II, brainstem – homeostatic regions, thalamus) [16, 17].
80 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

a b c

R’p

R’ R’α R’ R’β R’α R’ R’β

Rp”

+
Rp’1
Rp’ M3

M2
M1 R’’’
+

R’α R”
R’β
R’

Fig. 3.14  Chemical neurotransmission. A transmitter acts on (a) a single or (b) several postsyn-
aptic receptors, and (c) presynaptic receptors may also exist. In (d), the current concept of co-­
transmission (see text). Modified with permission from Cardinali [1]

Afferents of receptors in internal organs (interoceptors) are known collectively


as “visceral afferents” (Fig. 3.15) [18]. This group includes afferents from thoracic,
abdominal, and pelvic organs. A list of the principal unconscious sensory modali-
ties measured includes arterial BP, by stretch receptors in the carotid sinus and aor-
tic arch; central venous pressure, by stretch receptors in the walls of the great veins
and atria; inflation of the lung by stretch receptors in the lung parenchyma; tempera-
ture of the blood in the head by neurons in the hypothalamus; arterial PO2 by glo-
mus cells in the carotid and aortic bodies; pH of CSF by receptors on the ventral
Sensory Autonomic Neurons 81

Geniculate ganglion
Medulla VII Middle ear
oblongata
IX
X Uvula
Petrosal ganglion

Tonsils

Nodose Carotid artery


ganglion
Cervical
Lungs
spinal cord

Thoracic Heart
spinal cord
Spleen

Adrenal gland

Stomach

Intestine

Kidneys

Ovary, testis

Lumbar Uterus
spinal cord
Colon

Rectum

Sacral Bladder
spinal cord
Urethra

Fig. 3.15  Visceral afferents. Modified with permission from Cardinali [1]

surface of the medulla oblongata; osmotic pressure of plasma by cells in the orga-
num vasculosum of the lamina terminalis and other circumventricular organs; arte-
riovenous blood glucose difference by cells in the hypothalamus and at the periphery
(glucostats). Additionally, hormones and cytokines are sensed in many central and
peripheral organs. Collectively, these unconscious sensory modalities constitute the
afferent pathways of the autonomic reflexes (Fig. 3.16).
Some of these afferents enter the spinal cord and autonomic somatic pathways,
and their cell bodies are in the spinal ganglia. Others travel through the X pair
(approximately 80% of the vagal fibers are sensory; Fig. 3.17). It is noteworthy
that, through the interoceptive afferents, the ANS builds a kind of “structural
82 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Fig. 3.16  Visceral afferent Autonomic reflexes


(afferent pathways)
pathway as exemplified by
the afferents of a blood Spinal nerve
vessel. Modified with
permission from Cardinali
[1] Peripheral nerve
Rami
Paravertebral communicantes
ganglion of the
sympathetic chain

From
Prevertrbral abdominal
ganglion fat
Splanchnic
nerve

From Blood vessel


Viscera

Fig. 3.17  Afferent and Dorsal motor nucleus of the vagus


efferent fibers of the vagal
nerve. 80% of X pair nerve Nucleus of the solitary tract

fibers are sensory afferents.


Modified with permission Vagus nerve
from Cardinali [1]

Nodose ganglion
Baroreceptors
Ambiguous nucleus Chemoreceptors
Lung receptors
Gastrointestinal receptors
Endocrine & immune tissue

Heart

Striated muscle of Gastrointestinal tract


the larynx, pharynx
and esophagus

configuration of the interior space,” analogous to the contribution of proprioception


to the external configuration of the body image. The important projection of this
sensorial information to the limbic system provides the basis for the hypothesis that
they might constitute a kind of physiological correlate of unconsciousness. These
fibers are also important in visceral nociception.
Visceral afferents present in the sympathetic (spinal) and vagus pathways trans-
mit sensory inputs to the CNS about the physiological and pathological changes in
the local environment of visceral organ systems. Visceral afferent fibers responsive
to mechanical stimuli are either high or low threshold, with the high-threshold end-
ings frequently serving as nociceptors. Many high-threshold mechanosensitive end-
ings are triggered by chemical events and hence are bimodal in their sensitivity.
Chemical stimuli activating these endings depend on the organ in which they are
situated and the condition imposed [18].
Sensory Autonomic Neurons 83

Activation of these visceral afferents by mechanical and chemical stimuli elic-


its important ANS reflex responses. For example, visceral ischemia represents a
pathophysiological condition associated with cardiovascular disease, which leads
to the production and release of many metabolites including protons, bradykinin,
5-HT, His, endothelin, thromboxane and other cyclo-oxygenase products, and reac-
tive oxygen species (ROS), etc. These chemical mediators, both individually and
in combination, are involved in the activation of visceral afferents during ischemia
and reperfusion.
There is very precise information in the CNS with regard to what happens both
outside of and inside our body. Understanding this from a physiological point of
view is important because it helps to explain, for example, how changes in emotion-
ality can be detected as an early consequence of an organic disease such as cancer.
The CNS has information on the functional status of systems that have long been
considered independent of neural control, such as the immune system [1].
Most of the internal organs receive dual sensory innervation via sympathetic and
parasympathetic nerves. Visceral pain information is conveyed mainly through
sympathetic nerves, whereas reflex regulatory functions, for example, in the gastro-
intestinal tract, involved afferent fibers in parasympathetic nerves.
Autonomic sensory fibers travel through the visceral or peripheral nerves, pass
through the autonomic ganglia without interruption and have their neuronal bodies
in the spinal ganglia of the dorsal roots, through which they reach the spinal dorsal
horn. Afferent autonomic pathways are organized into two patterns of connections:

• Reflex circuits in the spinal cord and especially in the brainstem, which allow
adaptive responses of the visceral organs, for example, afferent baroreceptor or
chemoreceptor neurons connect with the medulla oblongata, from where efferent
fibers that control the corresponding effector responses are found (Chap. 4).
• Complex circuits, with upward projections via the anterolateral spinal cord and
brainstem nuclei to the hypothalamus and the limbic system, where information
is integrated. The responses affect multiple systems, autonomic, and emotional
and immunoendocrine, for example, the set of connections that control eating
behavior (Chap. 5).

Based on data obtained mainly in the somatosensory system, several general


principles of sensory organization can be established. The responses of sensory sys-
tems have four basic dimensions: spatiality, temporality, modality, and intensity. All
are deducible by conscious sensation and in the case of visceral perception, inferred
from neuroanatomical and functional neuroimaging studies.
Spatiality and temporality of sensation or perception relate to the real world or to
the internal medium itself. When something touches my skin, I can locate its posi-
tion on my body (spatiality) and identify the beginning and end of the stimulus
(temporality). Various studies indicate that visceral perception has a similar degree
of accuracy and representation in specific brain areas [16].
Modality defines the type of sensation, that is, we do not have the experience of
our environment as a whole, but through discrete elements produced by the
84 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

interaction of appropriate stimuli with their specific sensory receptors. In the case of
interoception, free nerve endings are the most common type of neural receptors.
The intensity, quantitative expression of feeling correlates with the amplitude of
the receptor potential and with the discharge frequency of action potentials in the
sensory nerve. In the case of conscious perception, the intensity is determinable
both objectively and subjectively. In the case of interoception (unconscious), it can
only be measured objectively. In general, it can be said that the intensity of a sen-
sory stimulus is encoded through two common mechanisms: (a) frequency code
(frequency of action potentials); (b) population code (number of sensory fibers
stimulated).
A second form of extraction of properties from a stimulus depends on the phe-
nomenon of adaptation or accommodation of sensory receptors. Invariably, a con-
stant stimulus stops to stimulate the sensory receptor. Rapidly adapting receptors
extract the dynamic characteristics of the stimulus (the time at which it is applied
and when it varies). Instead, slowly adapting receptors extract the static properties
of the stimulus (i.e., the time at which the stimulus is present).
The neural conduction velocity is a third form of abstraction stimulus. The differ-
ent sensory modalities are conducted by nerve fibers of varying diameter (Fig. 1.9).
It should be noted that the pattern of action potentials coming from an afferent is
not sufficient to fully sense the quality of a stimulus. For the brain to know that a
stimulus is a change in osmolarity coming from a given portion of the intestine, a
labeling of which afferent type has been activated must exist. Differing from the
internet, in which messages travel down a shared common line, sensory neurons
give each type of sensor its own private labeled line (Fig. 3.18). This implies the
need for a large number of lines in the spinal cord.
In general, one of the common characteristics of the ordered representation of
the sensory surface (somatotopic in the case of the somatosensory system, retino-
topic for the visual system, tonotopic for hearing, viscerotopic for interoception) is
the center–periphery antagonism of the peripheral fields. This center–periphery
antagonism is a result of the existence of lateral inhibition in synaptic circuits and
of the downward control of neural information, and is exerted at the upper levels of
the sensory pathway. The organization of receptive fields in a center and a periphery
with opposing characteristics increases the discriminative capacity of the stimulus
and improves its contrast [1].
Descending control by the upper structures reduces the entry of irrelevant infor-
mation. This downward, or input, control of sensory information describes the mod-
ulation at different levels of a neural pathway exerted by a structure located more
centrally in the hierarchical level. Because of this top–down control, only a part of
the information generated in the lower levels of the neural pathway reaches the
higher levels.
This is the origin of central analgesia. Together with the lateral inhibition of
circuits, downward control contributes to the center–periphery antagonism. Strictly
speaking, it is an element of the hierarchical information processing.
The anterolateral spinal cord system consists of several ascending pathways,
which together play an important role in the perception of pain and temperature in
Sensory Autonomic Neurons 85

ENCODER DECODER

Thyroid gland

LABELED LINES

Pancreatic islet

Blood vessel

Fig. 3.18  Sensory information is reaching the brain via labeled lines, i.e., by giving each type of
sensor its own private line. This differs from the internet, where the messages travel down a shared
common line. To separate an individual message from the others, each packet of information is
given a tag or label. The figure was prepared in part using image vectors from Servier Medical Art
(www.servier.com), licensed under the Creative Commons Attribution 3.0 Unported License
(http://creativecommons.org/license/by/3.0/)

interoception and constitute a secondary pathway for tactile sensitivity. The antero-
lateral system ends not only in the thalamus, but also in several regions of the brain-
stem (Fig. 3.19). Based on their termination site, three components of the anterolateral
system can be identified: (a) spinothalamic (or neospinothalamic); (b) spinoreticular
(or paleospinothalamic); (c) spinotectal. The last two are linked to interoception.
The analysis of the mechanisms of pain is useful in understanding the phenom-
ena involved in interoception [19, 20]. As interoception, pain is transmitted by spe-
cific neural pathways that start at the level of the superficial and deep free nerve
endings. These receptors respond to specific submodalities or are polymodal. They
are in part chemoreceptors that respond to chemicals produced by the cells. The
neurotransmitter in neurons of the myelinated fibers is the neuropeptide substance
P. It has been demonstrated that there is release of substance P and other neuropep-
tides, such as neoendorphin, somatostatin or CGRP, from these fibers.
Two populations of second-order neurons for pain transmission and interocep-
tion are found in the dorsal horn of the spinal cord: (a) relay neurons, which send
their axons to the thalamus or brainstem; (b) interneurons that connect to other,
similar neurons in the dorsal horn, or to relay neurons [19, 20].
86 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Limbic Somato-
system sensory
Thalamus
system

Brainstem Reticular formation

Spinoreticular
tract Lateral
Dorsal ganglion Dorsal root spinothalamic tract

Spinal nerve Spinal


Ventral cord
root
Sympathetic
chain Splacnic
White ramus nerve
communications

Splacnic ganglion

Fig. 3.19  The somatosensory and limbic pathways for abdominal pain. These connections under-
line the link between emotionality and digestive diseases. Modified from Cardinali [1]

It should be noted that a minority population of the anterolateral cord ascending


fibers ends in the thalamus. Most end diffusely in the reticular formation of the
brainstem, particularly in the area around the aqueduct of Sylvius (periaqueductal
gray), which is closely linked to the limbic system, hence the capacity of these
fibers to produce painful alertness and diffuse autonomic responses, such as stimuli
of respiratory and cardiovascular changes.
In the sympathetic and parasympathetic ganglia, important neural information
processing occurs, provided by similar synaptic circuits to those located at more
complex neural levels. Also, various hormones (steroids, neurotrophins, pituitary
and thyroid hormones) and cytokines affect the activity of the autonomic ganglia.
Let us take, for example, the SCG. In it, as in all other components of the chain
and prevertebral ganglia, the electrical response to preganglionic stimulation
includes the following aspects [4]:

• Quick depolarization, or rapid excitatory postsynaptic potential, responsible for


the genesis of the action potential in the ganglion neurons. The transmitter
involved in this rapid EPSP is ACh, acting through a nicotinic receptor, blocked
by hexamethonium.
• Slow hyperpolarization, or slow IPSP, about 2 s in duration, produced by the
release of DA from a group of catecholaminergic interneurons (small, intensely
fluorescent, SIF, cells).
• Slow depolarization, or slow EPSP, about 30 s duration, produced by ACh acting
through M1 receptors.
Spinal Autonomic Reflexes 87

• A slow late hyperpolarization, or late slow IPSP, produced by the release of neu-
ropeptides, in some cases, GnRH. Neurotransmitters such as GABA are also
present in sympathetic ganglia.

In the author’s laboratory, the SCG has been examined for the presence of hor-
mone receptors and for the effects of various hormones on the neural mechanisms
mentioned above (e.g., M1 cholinergic receptor, ACh and GABA release) [21, 22].
The participation of peripheral sympathetic innervation in neuroendocrine–immune
regulation has been also examined in detail in the SCG territory (Table 3.1).

 pinal Autonomic Reflexes Have Homologies and Differences


S
with Spinal Motor Reflexes

By analogy with the somatic motor system, in which the spinal α-motoneurons are
the “final common pathway,” the sympathetic and parasympathetic ganglion neurons
are considered the final common pathway for the ANS. However, the homology
between the two types of motor neurons, somatic and autonomic, is not complete.
Unlike somatic motoneurons, autonomic ganglion cells receive only a restricted
intermetameric input. Although in the case of a sympathetic paravertebral chain,
ganglion neurons can receive up- or downward influences from other metameres
(via the sympathetic trunk), it is in the intermediolateral column of the spinal cord,
and not in the sympathetic ganglia, that the integration of segmental spinal afferents
with downward influences from a higher level is performed. The cell bodies of pre-
ganglionic autonomic neurons of the intermediolateral spinal column are smaller
and more numerous than those of the α-motoneurons [3].
Synaptic connections between autonomic spinal afferents and efferents are
referred to generically as the autonomic reflex arc. In the cutaneovisceral reflex the
primary afferents come from the skin and the efferent sympathetic system inner-
vates the viscera. When afferents come from viscera (viscerocutaneous and vis-
cerosomatic reflexes), the efferent output can be both sympathetic (e.g., redness of
the skin after visceral irritation) and somatic (e.g., reflex contraction of the abdomi-
nal muscles due to viscera inflammation). There are at least three synapses between
afferent and efferent autonomic neurons. The existence of intervening synapses
facilitates modulatory influences on autonomic reflexes [3, 4].
The spinal organization of the sympathetic system tends to be segmental or
metameric. The preganglionic neurons of a particular spinal segment are in contact
with the visceral afferent entering at that level. In certain organs, this feature is very
pronounced: afferent heart, or excretory organs, make synaptic contacts at the seg-
mental level with preganglionic sympathetic and parasympathetic neurons that
innervate the same organs (intestinal–intestinal reflexes, cardio-cardiac reflexes,
bladder evacuation, etc.).
Such segmental organization of visceral reflexes can be tested clinically. In patho-
logical conditions of clinical significance, as in the case of inflammation of the gall-
bladder or appendicitis, the voluntary muscles of the affected metameres are
88 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

contracted, and the corresponding dermatomes are bloodshot. This situation is


explained by an inhibitory action of the visceral afferents originating from the affected
body on the vasoconstrictor efferent pathway of the same spinal segment (vasodila-
tion, reddening of the skin), and an excitatory action of these afferents on segmental
α-motoneurons (“defense” reflex of the abdominal muscles). Reciprocally, stimula-
tion of thermoreceptors in the skin produces a reflex inhibition of motility viscera
innervated by the same spinal segment through the corresponding sympathetic neu-
rons. This explains the therapeutic effect of applying heat in visceral pain [10].
After a complete spinal cord transection in humans, autonomic reflexes in seg-
ments below the section disappear for about 2–4 months. During the first phase of
this paralysis, the skin is dry and pink, because the sympathetic activity of the fibers
that innervate the sweat glands and vessels is very low. This hypoactivity gradually
reverses, to become hyperreflexia. In this phase, the skin stimulation produces
intense perspiration in areas innervated by deafferented spinal cord.
During the hyperexcitability after spinal cord section, exteroceptive stimuli, such
as nipping the inner thigh or manually dilating the external anal sphincter, or intero-
ceptive stimuli, such as contraction of the bladder muscles or dilation of the intesti-
nal muscles, trigger a mass reflex. During this period, other phenomena, such as
secretion of adrenal catecholamines, hypertension, piloerection, or profuse sweat-
ing, are observed [10].
After recovering from spinal shock, the isolated injured spinal cord is able to
maintain a series of simple autonomic reflexes. For example, heating the skin pro-
duces heat loss by vasodilation and sweating. The transition from supine to upright,
or blood loss, produces vasomotor compensatory segmental reflexes. Stimulation of
the skin at the corresponding metameric level triggers defecation or urination.
Another comparative aspect with motor reflexes is useful in understanding some
aspects of vegetative semiology. By losing neurogenic control, visceral effector
organs usually acquire a greater level of intrinsic activity. For example, after the
section of the sympathetic nerves, vascular smooth muscle is first paralyzed and
vasodilation ensues. However, within a few hours, the myogenic tone increases and,
depending on the internal pressure of the vascular filling, vasoconstriction occurs.
When denervated, most effector organs are hypersensitive to neurotransmitters that
reach them from the circulation. This explains why the miosis injury caused by
postganglionic sympathetic fibers of the pupil decreases in states of emotional
arousal, by action of plasma E, an effect not seen in the preganglionic injury (the
postganglionic neurons remain intact). The explanation for this denervation hyper-
sensitivity is that in the absence of direct innervation, enzyme systems responsible
for the catabolism of neurotransmitters disappear and there is an increase in recep-
tors in postsynaptic membranes, making it possible for bloodborne transmitters to
act on the denervated synapse. Denervated sweat glands are a notable exception to
this principle because they do not secrete sweat to cholinergic stimulation.
Denervation supersensitivity denervation is also observable in the CNS because
of injury or prolonged pharmacological blockade. For example, after prolonged D2
blocker neuroleptic treatment, the tardive dyskinesia observed is attributed to plastic
changes in the postsynaptic membranes [23].
Examples of Spinal Autonomic Reflexes 89

Postsynaptic supersensitivity comprises pre- and postsynaptic mechanisms.


Presynaptically, the denervated autonomic synapses lose the ability to reuptake the
neurotransmitter, so that larger amounts of agonist reach receptor sites, thereby
increasing the effect (presynaptic supersensitivity). Postsynaptically, after a few
hours of denervation, cells show increased synthesis of receptors that translocate
and disperse throughout the cell surface (postsynaptic supersensitivity). As part of
postsynaptic changes after denervation, there is also an increased sensitivity to vari-
ous nonspecific stimuli.
In the author’s laboratory, two strategies were used to examine the neuroendo-
crine and immune consequences of manipulation of sympathetic ganglia (in this
case, the SCG). The first consisted of the “deprivation experiment,” by examin-
ing endocrine and immune sequelae of superior cervical ganglionectomy (SCGx)
1–4 weeks after surgery. The second strategy was to determine the effect of tran-
sient postsynaptic activation that occurs during the early phase of anterograde
degeneration (“Wallerian”) from sympathetic nerve endings in SCGx animals
[24]. This process of postsynaptic activation observed during Wallerian degen-
eration is called “degeneration reaction.” With a latency of 1–3 days, latency that
depends on the distance between the SCG and the territory studied (i.e., the
length of the axonal stump), nerve section is followed by transient hyperactivity
in the denervated territory following the supraliminal release of the neurotrans-
mitter NE from degenerating nerve endings, which continues for about 48 h
before attaining a final and irreversible paralysis. One advantage of this experi-
mental model is that the entire period of Wallerian degeneration in the path of the
SCG can be easily controlled in the conscious animal by the degree of eyelid
retraction (due to activity in the degeneration level of the periorbital muscles). It
should be noted, however, that the continuous transmitter release found during
degeneration is only a rough indication of the effect of the physiological release,
which is phasic and dependent on the stimulation of the afferent pathway.
Therefore, the degeneration reaction only indicates the potential role of the post-
synaptic activation, but not its physiological details (Table 3.1). However,
although it proved to be a model of great heuristic potential, it is unfortunately
largely unnoticed today [24].

 rination, Defecation, and Pupillary and Sexual Responses


U
Are Examples of Spinal Autonomic Reflexes

The function of the urinary bladder is the periodic storage and full voiding of urine
produced continuously by the kidney. This function is based on the myogenic activ-
ity of bladder smooth muscle and on autonomic and somatic neural mechanisms. In
controlling the bladder, prolonged phases of urine collection alternate with short
expulsive periods (urination) [6].
During urine collection, neural activity prevents emptying of the bladder. The
bladder is filled at a rate of about 50 ml/h. The plasticity of the bladder smooth mus-
culature ensures that intravesical pressure is increased only mildly during filling.
90 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Fig. 3.20 Bladder
innervation.
Parasympathetic
innervation stimulating
Mesenteric ganglion
the detrusor derives from
S2–S4, whereas Lumbar
sympathetic (mesenteric spinal cord
ganglion) innervation
inhibits the detrusor and
stimulates the trigone. Ureter
The external sphincter
receives somatic
innervation through the
pudendal nerve. Modified
with permission from Pelvic nerve
Cardinali [1]
Pudendal nerve
2 2
3 3
4 Internal 4
sphincter External
Sacral spinal sphincter Sacral spinal
cord cord

When the bladder is filled up to 150–250 ml, a perceived urgency to urinate


occurs, a phenomenon triggered by brief increases in intravesical pressure. Upon
reaching a content of 250–500 ml, volunteer continence mechanisms are overcome
and the evacuation phase of urination is triggered [6].
In Fig. 3.20, the different neural and muscular components participating in mic-
turition are depicted. Bladder muscle (detrusor muscle) is innervated by parasympa-
thetic fibers from the pelvic nerve. This innervation is essential for normal bladder
emptying. The sympathetic system inhibits the detrusor muscle and stimulates the
bladder internal sphincter (trigone). Sympathetic preganglionic neurons are located
in the intermediolateral column of the upper lumbar cord, and reach the bladder via
the inferior mesenteric plexus [25].
The somatic motor system supplies the external sphincter through the pudendal
nerves, which carry the axons of motor neurons located in segments S2–S4. The
filling level of the urinary bladder is detected by stretch mechanoreceptors located
in the bladder wall, whose afferent fibers run via the pelvic nerve, and the nerve cell
bodies are in the dorsal root ganglia of the corresponding sacral segments.
Sensory information, particularly nociceptive, originates from the trigone and is
carried by fibers traveling parallel to the sympathetic fibers and having the neuronal
bodies in the dorsal ganglia of the upper lumbar region. This afference is important
in the processes of bladder inflammation (pollakiuria, or pain when urinating) [25].
Up to certain limits, increased bladder volume induces relaxation of the detrusor
muscle; thus, intravesical pressure does not increase and urinary continence can be
maintained. This reflex action is mediated by sympathetic fibers, which inhibit the
detrusor muscle via β-adrenoceptors, while stimulating the smooth muscle of the
Examples of Spinal Autonomic Reflexes 91

Fig. 3.21  Spinal and


pontine reflex arches
controlling urination.
Modified with permission
from Cardinali [1]
Pontine
micturition center

Bladder afferent

Detrusor

Pelvic nerve

Pudendal nerve
Sphincter
Somatic pathway
Parasympathetic pathway

trigone and the urethra through α-adrenoceptors. The initial contraction of the blad-
der causes a greater excitation of the mechanoreceptors and, by reflex, a new
increase in the contraction, so that the voiding reflex is self-reinforcing.
Electrical stimulation of the anterior brainstem region triggers the micturition
reflex (Fig. 3.21). Once the bladder has begun to empty, the process accelerates
exponentially (“positive feedback”) because of: (a) increased activation of the
mechanoreceptors of the bladder wall, this time by contraction of the detrusor; (b)
activation of trigone afferents due to the passage of urine; (c) blocking supraspinal
inhibitory influences of the reflex at the spinal cord level; (d) inhibition of sacral
α-motoneurons that control the external bladder sphincter [6, 26].
After spinal cord section, and when the reflex restarts after the spinal shock, this
is due exclusively to the lower arc operation shown in Fig. 3.21. This is called “auto-
matic bladder,” triggering the reflex by stimulation of the corresponding derma-
tome. The lower reflex arch is the single acting pathway in the newborn, and only in
later stages does the upper reflex arc develop [6].
Defecation is under the control of the enteric intrinsic system, sacral parasympa-
thetic innervation, and somatomotor mechanisms (Fig. 3.22) [26]. The role of the
sympathetic system is only minor. Two sphincters close the distal end of the rectum:
(a) the internal anal sphincter, composed of smooth muscle without voluntary con-
trol; (b) the external anal sphincter, composed of striated muscle innervated by
motor neurons of the segments S2 and S4, whose axons travel via the pudendal
nerve (Fig. 3.22). Usually, both anal sphincters are closed. The tonic contraction of
92 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Fig. 3.22 Defecation.
The external anal sphincter
is under voluntary control, Descending
whereas the internal anal colon
sphincter is under
parasympathetic control. Pelvic nerve
The internal sphincter is
relaxed reflexively by
rectal distension. Modified
with permission from
Cardinali [1]

Sigmoid
Rectum Pudendal Sacral
nerve cord

External Internal
sphincter sphincter

the external anal sphincter is due to a reflex with afferences from the perianal mus-
cles and surrounding tissue, especially the anal skin.
When the rectum is filled with intestinal content, peristaltic contractions of the
descending colon relax the internal sphincter, and reflexively increase the contrac-
tion of the external sphincter. The relaxation of the internal sphincter as a reflex
originated in the enteric autonomic system, whereas contraction of the external
sphincter is a reflex that involves somatic afferents that run through the pelvic nerve
to the sacral spinal cord (Fig. 3.22).
This sequence of phenomena is accompanied by the urgency to defecate, a con-
scious sensation triggered by the stretching of the rectal and colonic wall. After
about 30 s to 1 min, the relaxation of the internal sphincter disappears simultane-
ously, and because of the rectal plasticity, the rectus musculature adapts to the new
content. The result is that the need for defecation disappears.
Through these neural mechanisms, a healthy individual can maintain fecal conti-
nence up to a volume of rectal content of approximately 2 L. Cortical mechanisms par-
ticipate in fecal continence: (a) through the excitation of α-motoneurons that innervate
the external sphincter; (b) through the inhibition of the parasympathetic reflex (Fig. 3.22).
Defecation is initiated by a voluntary effort (increased intra-abdominal pres-
sure), the simultaneous supraspinal facilitation of parasympathetic pathways, and
the relaxation of both sphincters. Lesions of the sacral spinal cord eliminate the
defecation reflex. The spinal cord section at the thoracolumbar level causes elimina-
tion of the supraspinal control with maintenance of the reflex, which can be excited
in the paraplegic patient through other mechanisms (e.g., manual dilation of the anal
sphincter), thus ensuring periodic bowel movement [27].
The ANS influences numerous ocular functions [28]. It does this by way of para-
sympathetic innervation from postganglionic fibers that originate from neurons in
Examples of Spinal Autonomic Reflexes 93

Fig. 3.23 Autonomic Third nerve nucleus


control of pupillary
Oculomotor nerve
diameter. Modified
with permission from Ciliary ganglion
Cardinali [1]
Acetylcholine
Midbrain

Superior cervical
ganglion Norepinephrine
Spinal cord

Sympathetic Parasympathetic
stimulation stimulation

the ciliary and pterygopalatine ganglia, and by way of sympathetic innervation from
postganglionic fibers that originate from neurons in the SCG.
Ciliary ganglion neurons project to the ciliary body and the pupillary sphincter
muscle of the iris to control ocular accommodation and pupil constriction respec-
tively. SCG neurons project to the pupillary dilator muscle of the iris to control
pupil dilation (Fig. 3.23). Fibers of circular arrangement and others of radial dispo-
sition constitute the smooth muscle of the iris. The contraction of the first fibers,
whose innervation is parasympathetic, provokes miosis, and that of the second
fibers, innervated by the sympathetic system, causes mydriasis. The parasympa-
thetic innervation of the iris originates in the Edinger–Westphal nucleus. The pre-
ganglionic fibers leave the nucleus, accompanying the third cranial nerve, and
follow the path of the nerve until reaching the ciliary ganglion located behind the
eyeball. From the ciliary ganglion, the postganglionic fibers innervate the ciliary
muscle, which is responsible for the accommodation, and the constrictor muscle of
the iris. Neurotransmission at this level is cholinergic [28].
Ocular blood flow is controlled both via direct autonomic influences on the
vasculature of the optic nerve, choroid, ciliary body, and iris, and via indirect influ-
ences on retinal blood flow. In mammals, this vasculature is innervated by vasodi-
latory fibers from the pterygopalatine ganglion, and by vasoconstrictor fibers from
the SCG. Intraocular pressure is regulated primarily through the balance of aque-
ous humor formation and outflow. Autonomic regulation of ciliary body blood
vessels and the ciliary epithelium is an important determinant of aqueous humor
formation; autonomic regulation of the trabecular meshwork and episcleral blood
vessels is an important determinant of aqueous humor outflow. These tissues are all
innervated by fibers from the pterygopalatine and the SCG. In addition to these
classical autonomic pathways, trigeminal sensory fibers exert local, intrinsic influ-
ences on many of these regions of the eye, and on some neurons within the ciliary
and pterygopalatine ganglia. Regarding sympathetic innervation, a first group of
central fibers originates in the hypothalamus and projects to the intermediolateral
94 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

horn of the first thoracic segments of the spinal cord. Here, they synapse with pre-
ganglionic neurons that send fibers to the SCG, where postganglionic fibers going
to the eyeball originate. These adrenergic fibers innervate the iris dilator and the
levator muscle (Müller muscle) [28].
From the functional point of view, the parasympathetic system constricts the
pupil and the sympathetic system dilates the pupil (Fig. 3.23). Thus, a miotic pupil
can be due to a decrease in sympathetic or an increase in parasympathetic activity
and the opposite is true for a mydriatic pupil. The evaluation function can be per-
formed by observing the pupillary responses to the local application of drugs.
Parasympatholytic (anticholinergic) drugs such as atropine block parasympa-
thetic activity by competing with ACh at the effector cells of the iris sphincter and
ciliary muscle, thus preventing depolarization. Parasympathomimetic (cholinergic)
drugs such as pilocarpine are structurally similar to ACh and can depolarize the
effector cell, thus causing miosis. Sympathomimetic (adrenergic) drugs such as E
stimulate the receptor sites of the dilator muscle cells. A defect in the sympathetic
pathway affects the pupillary dilator muscle and results in Horner’s syndrome
(Chap. 7). A defect in the parasympathetic pathway affects the pupillary sphincter
muscle and results in a larger pupil. Defects in both pathways affect the pupillary
dilator and sphincter muscles [29].
Instillation of cocaine, a sympathomimetic agent, produces mydriasis, whose
absence indicates a lesion at some segment of the sympathetic chain, as in Horner’s
syndrome. Hydroxyamphetamine, however, only causes mydriasis in the case of
preganglionic sympathetic injury because it acts by causing NE depletion from
intact postganglionic synaptic terminals. An exaggerated mydriatic response to that
of the contralateral eye, after instillation of phenylephrine, suggests the existence of
postganglionic sympathetic denervation. Instead, an exaggerated miotic response to
pilocarpine indicates postganglionic parasympathetic denervation.
In humans, the sympathetic, parasympathetic, and somatic nervous divisions
participate in the control of sexual responses of erection, glandular secretion, emis-
sion, and ejaculation [30]. Penile erection can be caused by supraspinal centers, in
response to visual, auditory, or psychological stimuli (psychogenic erection), or by
a spinal reflex (reflex erection). In the latter, the impulses evoked by cutaneous
stimulation of the genital area travel through the pudendal nerves to the sacral spinal
segments S2–S4. The efferent pathway originates in the same segments, and contin-
ues through the parasympathetic pelvic nerves, to produce vasodilation of the arter-
ies and closure of the penile arteriovenous shunts, thereby increasing the blood flow
of the corpora cavernosa to allow erection. Not surprisingly, the vasculature, epithe-
lia, and smooth muscle of all urogenital organs receive adrenergic innervation.
These nerves contain non-adrenergic, noncholinergic neurotransmitters such as
ATP and NPY. Cholinergic nerves increase motility in most urogenital organs. The
major non-adrenergic, noncholinergic transmitters found to influence urogenital
organs include those containing VIP/PACAP, galanin, and NO [31].
The glandular secretion of seminal vesicles, Cowper glands, and the prostate
gland is controlled by the parasympathetic system. The emission of semen and glan-
dular secretions to the urethra depends on the sympathetic activity, which causes
The Enteric ANS as an Individual Entity 95

contraction of the smooth muscle of the vas deferens and the excretory ducts.
Sympathetic efferents are also responsible for closure of the bladder neck to prevent
retrograde seminal flow into the bladder. Ejaculation is caused by a somatic reflex
that causes rhythmic contractions of the bulbocavernosus and ischiocavernosus
muscles innervated by the pudendal nerves. The repeated stimulation of the penis
releases sacral centers from superior inhibitory signals. In situations of anxiety, in
which sympathetic tone is high, premature ejaculation can occur [30].
In women, the physical expression of sexual arousal is related to parasympa-
thetic activity, which, through the pelvic nerves, produces congestion of the clitoris
and vulva and vaginal lubrication, and to somatic activity, via the pudendal nerves,
which causes contraction of the vaginal sphincter and pelvic floor muscles during
orgasm. Genital vasodilation seems mainly mediated by VIP. On the other hand,
sympathetic stimulation induces contractions in the smooth musculature of the fal-
lopian tubes and uterus. In both sexes, afferent somatic impulses are part of the
reflex arcs, and their transmission to higher centers through the spinal cord is essen-
tial for the conscious perception of sexual phenomena and their regulation.
Penile erection or tumescence occurs during the REM sleep stage, although in
adolescents it is not only confined to this stage. This has been proven in humans
from 3 to 79 years of age. During the active sex life, it coincides with LH release
pulses. Although its functional role remains unknown, the presence or absence of
erection during sleep is used for the differential diagnosis between organic and psy-
chogenic impotence. Similarly, clitoral erections and increased vaginal blood flow
are seen in women during REM sleep [30].

The Enteric ANS as an Individual Entity

The enteric ANS is sometimes called the “second brain” because of the diversity of
neuronal cell types and complex, integrated circuits that permit the enteric ANS to
autonomously regulate many processes in the bowel [32]. In humans, the enteric
ANS contains about 500 million neurons, far more than the number of neurons in
the remainder of the peripheral ANS. It is a complex network of neurons and glia
that resides in the myenteric and submucosal plexus of the bowel. The myenteric
(Auerbach) plexus, located between longitudinal and circular muscle, primarily
controls muscle contraction and relaxation. The submucosal (Meissner) plexus,
found between circular muscle and bowel mucosa, regulates fluid secretion and
absorption, modulates blood flow, and responds to stimuli from epithelium and
lumen to support bowel function (Fig. 3.24).
The enteric ANS includes sensory neurons, interneurons, and motor neurons.
Full arches are reflected in the enteric nervous system. The extrinsic innervation of
the gastrointestinal tract given by the sympathetic and parasympathetic nerves acts
on these components [33].
The extrinsic ANS is critically involved in modulating the local reflexes and
secretion, absorption, digestion, motility, and sensation in the gastrointestinal tract,
both during fasting and postprandially. In general, the craniosacral parasympathetic
96 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Myenteric plexus
Circular muscle
deep muscle
plexus
submucosal
plexus

Longitudinal
muscle

Villius plexus Periglandular plexus

Fig. 3.24  Enteric neurons are organized into two nerve plexuses, the myenteric plexus and the
submucosal plexus, with other minor side plexuses including muscle, periglandular layers, and the
villi. The enteric nervous system includes sensory neurons, interneurons, and motor neurons.
There are full arches reflected in the enteric nervous system

nervous system is excitatory (e.g., secretion, peristalsis) and the thoracolumbar


sympathetic supply is mostly inhibitory (e.g., proabsorptive and inhibition of motil-
ity with stimulation of the sphincters). Although intrinsic neural and local paracrine
control can sustain most gastrointestinal functions, the extrinsic autonomic nervous
system provides the highly integrated digestive system with additional output [34].
Enteric glia are important components of the enteric nervous system and form an
extensive network in the mucosa of the gastrointestinal tract. Initially regarded as
passive support cells, it is now clear that they are actively involved as cellular inte-
grators in the control of motility and epithelial barrier function [35]. Enteric glia
form a cellular and molecular bridge among enteric nerves, enteroendocrine cells,
immune cells, and epithelial cells, depending on their location.
The gut wall consists of several concentric layers, an inner mucosa, surrounded
by several layers: submucosa, muscle and serous (Fig. 3.24). Motility of the diges-
tive tract is characterized by the synchronized contraction of the inner and outer
muscle layers, resulting in two types of movement: (a) peristalsis, which sets in
motion the intestinal contents and ensures a cephalo-caudal flow; (b) segmenting,
contributing to mixing food with digestive secretions. This is a myogenic contrac-
tile activity that has its origin in the gut wall itself, particularly in the nerve plexus.
Myogenic origin explains the persistence of rhythmic patterns and motility in iso-
lated denervated digestive segments [36].
The mucosa has three components: specialized epithelial cells lining the lumen;
the underlying lamina propria, a layer of connective tissue containing small blood
and lymphatic vessels, immune cells, and nerve fibers; and the muscularis mucosa,
a thin layer of muscle cells. The submucosa is a layer of connective tissue directly
beneath the mucosa, containing the largest blood vessels and the submucosal plexus
(Meissner). This nerve plexus is particularly important for the control of secretion.
In some areas, submucosal glands and lymphoid tissue are present. The external
muscle is composed of an inner circular and an outer longitudinal layer of smooth
The Enteric ANS as an Individual Entity 97

muscle and is responsible for the motility of the gastrointestinal tract. Among these
muscle layers are the myenteric nerve plexus (Auerbach), a division of the enteric
nervous system that regulates motility. The serosa is an outer sheath of squamous
mesothelial cells and connective tissues, where nerves and larger blood vessels
travel in a bed of connective and adipose tissue [36].
Like any neural arch, there are sensory enteric neurons, interneurons, and motor
neurons. Sensory information in the enteric nervous system comes from changes in
luminal volume or environment. The endocrine and paracrine cells function as aux-
iliary sensors [10].
Reflexes regulate colon epithelial responses through cholinergic interneurons.
Two categories of epithelial and submucosal motor neurons innervate choliner-
gic and VIPergic cells and each uses additional neuroactive substances. The sym-
pathetic and parasympathetic tone modulates ion transport, with a cholinergic
basal secretory influence. The loss of the regulatory mechanisms of sympathetic
nerves in diabetic autonomic neuropathy is associated with the development of a
“diabetic diarrhea” and can be corrected by the administration of α2 adrenocep-
tor agonists [33].
Peptidergic neurons release a specific combination of mediators (e.g., VIP, CCK,
bombesin). These mediators can act either as classical neurotransmitters or alterna-
tively as neuromodulators, with fine tuning of neural circuits in presynaptic sites of
origin neurons or from other neurons. Individual neurotransmitters may have bipha-
sic effects at different concentrations.
Figure 3.25 summarizes the regulation of gastric acid secretion by nerves and
hormones. Secretion of gastric acid between meals is low. The cephalic phase of
secretion (~30% response) is initiated by the sight, smell, taste, and swallowing of
food and these stimuli activate the dorsal motor nucleus of the vagus nerve. In the
body of the stomach, postganglionic nerves release ACh, which activates parietal
cells directly by M3 receptors. ACh also induces His release from enterochromaffin-­
like cells (ECL), which stimulates the secretion of H+ ions by the parietal cells. In
the gastric antrum, vagal stimulation induces the release of gastrin-releasing pep-
tide, from postganglionic fibers, and the release of gastrin, thus indirectly stimulat-
ing the secretion of H+ ions. ACh also inhibits the release of somatostatin D cells in
the corpus and pylorus [33].
The gastric phase (~70% response) secretion is induced by stimuli within the
stomach. Vagal sensory nerves detect gastric distention by food and cause a vagova-
gal reflex whereby ACh is released and promotes the release of H+. Partially digested
proteins and amino acids stimulate gastrin release from G cells in the pylorus. Both
the G and D cells (which release somatostatin) are open-type endocrine cells that
directly detect stomach content. Gastrin stimulates further acid secretion.
Acidification of pylorus stimulates somatostatin release, which inhibits acid secre-
tion by negative feedback [36].
During the intestinal phase of digestion, products of protein entering the small
intestine stimulate gastrin release from the G cells in the duodenum. Many sub-
stances, notably fatty acid, stimulate the secretion of hormones in the small intestine
(secretin, CCK) which inhibits gastric acid secretion.
98 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

VAGUS NERVE LUMEN


ACh

M3-R

CCKB-R Parietal
cell Protons
G

H2-R
Histamine
Peptones, AA Acid

G
G cell D cell ANTRUM
ACh ECL cell

ACh S-R
GRP

ACh GRP-R ACh


BLOODSTREAM
ACh

Fig. 3.25  Regulation of gastric acid secretion by nerves and hormones. During the cephalic phase
of digestion, vagal cholinergic nerves stimulate directly the parietal cells and induce histamine
release from enterochromaffin-like cells (ECLs), which also stimulates parietal cells. Vagal fibers
also release gastrin-releasing peptide (GRP) in the antrum to induce gastrin (G cells), which
through the bloodstream induces histamine release and stimulates the parietal cells. During the
gastric phase of digestion, food in the stomach causes local reflexes that stimulate the secretion of
gastrin. Acidification of the gastric antrum stimulates somatostatin release (D cells), which inhibits
the release of gastrin and thus acid secretion. Vagal stimulation inhibits somatostatin release

The responses are exerted through the enteric ANS, the enteroendocrine cells,
the CNS via the extrinsic innervation of the gastrointestinal tract given by the sym-
pathetic and parasympathetic nerves, and the gut immune and tissue defense sys-
tems. It is apparent that the control of the digestive organs is an integrated function
of all these effectors. The enteroendocrine cell release about 20 different hormones,
together making the gut endocrine system one of the largest endocrine organs in the
body. Influenced functions include satiety, mixing and propulsive activity, release of
digestive enzymes, induction of nutrient transporters, fluid transport, local blood
flow, gastric acid secretion, evacuation, and immune responses [36].
Gut content receptors, including free fatty acid, peptide, and phytochemical
receptors, are primarily located on enteroendocrine cells. Hormones released by
enteroendocrine cells act via both the enteric ANS and the CNS to optimize diges-
tion. Toxic chemicals and pathogens are sensed and then avoided, expelled, or
metabolized. These defensive activities also involve the enteroendocrine cells and
signaling from enteroendocrine cells to the enteric nervous system and the CNS [37].
The various types of epithelial cells and smooth muscle cells are the primary
effectors controlling gastrointestinal transit, absorption, and secretion. They are
The Enteric ANS as an Individual Entity 99

both directly and indirectly regulated by the enteric ANS. The intestinal epithelium
is a single layer of cells that combines the potential for nutrient absorption, water,
and electrolyte secretion (and reabsorption) with sensory and endocrine function
and with its role as a physical barrier to prevent the uncontrolled entry of harmful
substances and microbes into the body. Within the epithelium of the adult small
intestine, many types of differentiated cells can be found: enterocytes, which play a
key role in water and electrolyte secretion in the crypts and absorb various nutrients
at the villus tips; enteroendocrine cells, which secrete many different types of endo-
crine and neurotransmitter-like signaling molecules; goblet cells, which secrete
mucus into the gut lumen; and Paneth cells, which are of an immune nature.
Enterocytes, enteroendocrine cells, and goblet cells are present in the villi, whereas
Paneth cells reside in the crypts together with stem cells, which divide to replenish
these differentiated cell types (Fig. 3.26) [26].
The colon lacks villi, but has crypts that house intestinal stem cells, differen-
tiating mostly to form colonocytes (absorptive cells) and enteroendocrine cells.
Enteric circuits controlling secretion and absorption are mostly located in the
submucous plexus [36]. In the small intestine, together with systemic (sympa-
thetic) pathways, local enteric reflexes act through the activation of secretomotor

Goblet T lymphocytes
cells B lymphocytes
Plasma cells Lamina
Macrophages propia
Mastocytes
Eosinophilis
Epithelial
Intestine cells
Intraepithelium
Villi lymphocytes
Lumen

Peyer’s patch
Immune
system
Mesenteric
lymph node Circulation

Thoracic
duct

Fig. 3.26  The adaptive immune system of the mucosa or gut-associated lymphoid tissue monitors
the contents of the intestinal lumen through a variety of mechanisms used by cells such as lym-
phoid and myeloid lineages. Lymphoid cell aggregates form Peyer’s patches and lymphoid folli-
cles located throughout the intestine. These lymphoid aggregates also play an important role in
immune surveillance, guarding against pathogenic bacteria, viruses, and toxins, and allowing tol-
erance to dietary substances and potentially immunogenic bacteria
100 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

and vasodilator neurons to promote or inhibit secretion as part of digestion and


whole body homeostasis.
As far as intestinal movements are concerned, different enteric ANS circuits
direct various motor patterns depending on the region of the gut and the physiologi-
cal circumstances [36]. Although esophageal and gastric motility are controlled via
intrinsic and extrinsic innervation, propulsive and mixing contractile patterns in the
small intestine and colon mostly rely on the enteric ANS. A basic circuit, including
sensory neurons (intrinsic sensory neurons), orally directed interneurons and excit-
atory motor neurons, anally directed interneurons, and inhibitory motor neurons,
mediates propulsion. A key element of this circuit is its polarity; stimuli that excite
motor pattern generators activate an orally directed (ascending) pathway leading to
smooth muscle contraction. The same stimuli excite an anally directed pathway that
inhibits or relaxes the circular muscle. Both classes of motor neurons receive input
from local sensory neurons and from relevant interneurons; they are the final com-
mon outputs of the monosynaptic and polysynaptic pathways [33].
In Fig. 3.27, the peristaltic reflex of the small intestine is schematized. Enteric sen-
sory nerves detect the chemical or mechanical stimulation of the mucosa, or the stretch-
ing of the muscular layer. The signals are transmitted in an oral or anal direction by
interneurons. Excitatory motor nerves release ACh and substance P, which cause mus-
cle contraction on the oral side of the stimulus. Nervous motor inhibitors release VIP
and NO, which cause muscle relaxation on the anal side of the stimulus [33].
The response to local stretching of the intestinal wall, for example by a bolus of
partially digested food, results from the direct activation of mechanosensitive neurons
in the myenteric plexus. The intrinsic sensory neurons have mechanosensitive elements
in the myenteric plexus and probably in the smooth muscle. These neurons also project
to the mucosa and respond to chemical stimuli indirectly via the release of 5-HT, ATP,
and other mediators from EEC, giving them a polymodal stimulus–response profile. To
control gut motility, enteric nerve circuits act in concert with myogenic control ele-
ments, notably interstitial cells of Cajal and fibroblast-like cells. By acting as intestinal
pacemaker cells, interstitial cells of Cajal are responsible for the generation of intesti-
nal slow waves and other rhythmic contractions of the smooth musculature and for
excitatory and inhibitory transmission from enteric neurons to smooth muscle cells,
and thereby integrating enteric ANS and slow-­wave activity [33, 36].
Cells of the immune system and its wide range of products play an integral role
in the regulation of fluids and are closely interrelated with the enteric nervous sys-
tem and the endocrine paracrine network [37]. The lamina propria is a location for
immunocompetent cells. Most of them (60%) are T lymphocytes with fewer B lym-
phocytes and plasma cells (25–30%), macrophages (8–10%), and mast and poly-
morphonuclear cells (2–5%). Inflammation causes an increase in the number of
immune cells in the gut. Acute bacterial infections result in an increase in polymor-
phonuclear leukocytes, whereas lymphocytes in celiac disease characteristically
increase. In inflammatory bowel disease, there is activation of all components of the
immune system, with an increase in immunoglobulin (Ig) G-secreting cells.
Therefore, the cause of the inflammatory reaction can determine the type of inflam-
matory cells recruited and the range of cytokines released.
The Enteric ANS as an Individual Entity 101

Mechanical muscle stimulation (stretching)

Sensory
nerve

Excitatory Inhibitory
motoneuron Interneuron Interneuron motoneuron

ACh VIP
Substance P NO
Sensory
nerve
Contraction Relaxation

Mechanical or chemical mucose stimulation


LUMEN
Fig. 3.27  Peristaltic reflex of the small intestine. Enteric sensory nerves detect the chemical or
mechanical stimulation of the mucosa or the stretching of the muscular layer. The signals are
transmitted in an oral or anal direction by interneurons. Excitatory motor nerves release ACh and
substance P, which cause muscle contraction on the oral side of the stimulus. Inhibitory motor
nerves release vasoactive intestinal peptide (VIP) and nitric oxide (NO), which cause relaxation of
the muscle on the anal side of the stimulus. The figure was prepared in part using image vectors
from Servier Medical Art (www.servier.com), licensed under the Creative Commons Attribution
3.0 Unported License (http://creativecommons.org/license/by/3.0/)

Cytokines, eicosanoids, and other peptide mediators interact with immune cells,
neurons, and epithelial cells directly or alternately alter ion transport rates or gut
barrier function. The adaptive immune system of the mucosa or gut associated lym-
phoid tissue monitors the contents of the intestinal lumen through a variety of mech-
anisms used by cells both as lymphoid myeloid lineages. Myeloid cells (dendritic
cell populations, specific macrophages) extend through processes of the intestinal
epithelial barrier, which are associated with the luminal environment. Lymphoid
cell aggregates form Peyer’s patches (larger aggregates in the distal small intestine)
and lymphoid follicles located throughout the intestine (Fig. 3.26). These lymphoid
aggregates also play an important role in immune surveillance, guarding against
pathogenic bacteria, viruses, and toxins, and allowing tolerance to dietary sub-
stances and potentially immunogenic bacteria.
102 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

Specialized goblet cells secrete mucus in the intestine. The mucus forms a pro-
tective layer on the epithelial cells and antimicrobial peptides are secreted into the
intestinal lumen. Paneth cells produce and secrete lysozyme and α-defensins that
contribute to the defense. Clover peptides are also secreted in the lumen of the gas-
trointestinal tract. One of their many effects is promote the healing of mucosal
lesions [33, 36].
In Chap. 1, we define microbiome as the set of microorganisms that are normally
located in different places in the human body, in particular the digestive tract [38].
These microbial components aid in the digestion of food, produce vitamins, and
protect against the colonization of other microorganisms that may be pathogenic.
The gut microbiome is highly dynamic, exhibiting daily cyclic fluctuations, which
have repercussions for the host metabolism and provide evidence for the cross-reg-
ulation of prokaryotic and eukaryotic circadian rhythms.
Figure 3.28 depicts the bases of the effect of the microbiome and the way in
which they interact with the neuroimmune and bioenergetic afferents and with cog-
nitive function. An essential pathway is the neural pathway through innervation of
distant organs (e.g., intestine, spleen, liver) and neuroimmune signals that can be
transmitted bidirectionally between the brain and the periphery [39]. Another
important route is the general circulation, where circulating immune signals (cyto-
kines, chemokines) and bioenergetic signals (e.g., endogenous metabolites or
microbial origin) can exert a systemic impact on the brain microenvironment. Such
immune/bioenergetic signals can have an impact on the blood–brain barrier signals
thus spread to the brain, or they can cross the blood–brain barrier, exerting direct
neuronal effects (Fig. 3.28).

 he Overlooked Role of the Local Autonomic Projections


T
in Neuroendocrine Communication

The dominant paradigm of endocrinology in the twentieth century endowed the


innervation of endocrine glands with a merely secondary role compared with the
corresponding trophic hormones. Indeed, both endocrine and immune tissues have
three identifiable neural systems: (a) sympathetic noradrenergic neurons derived
from para- or prevertebral chains in the corresponding metameres; (b) parasympa-
thetic cholinergic neurons whose bodies are in the local parasympathetic ganglia;
(c) peptidergic neurons, most of them located in the innervated organ. Concerning
peptidergic innervation, there is an almost universal presence of substance P, a typi-
cal neurotransmitter of type C sensory fibers. That is, there is both sensory and
motor innervation in endocrine and immune tissues. The sensory aspect is a strong
evidence for the now recognized viscerotopic representation of endocrine and
immune structures in the CNS, as discussed in this chapter.
The thyroid and parathyroid glands are a typical case of this threefold innerva-
tion (Fig. 3.29). These structures are innervated by: (a) noradrenergic neurons
derived from the SCG, or from lower cervical nodes and reaching the thyroid
through the SCG and the external carotid nerve; (b) parasympathetic cholinergic
Local Autonomic Projections in Neuroendocrine Communication 103

Behavior, Neurogenesis,
Cognition synaptic plasticity

Microglia
Neuron
Astrocyte
Sympathetic
afferents
Blood/Brain barrier
Vagall
afferents
en s Blood

Spleen Immune cells

IMMUNE AND BIOENERGETIC SIGNALS

Liver

Cytokines,
chemokines
emok Metabolic signals
(5HT, kynurenines)
renine

Afferent Enteric Hormone signals


Microbial signals
neuron glia (Ghrelin,, GLP
GLP/1)
A, bile acids)
(FFA, a

Intestinal
Immune
Intestine System

MICROBIOME

Fig. 3.28  The basis of the effect of the microbiome on brain function is shown. An essential path-
way is the neural pathway, via innervation of distant organs (e.g., intestine, spleen, liver) and
neuroimmune signals that can be transmitted bidirectionally between the brain and the periphery.
Another important route is through the general circulation, where circulating immune signals
(cytokines, chemokines) and bioenergetic signals (e.g., endogenous metabolites of microbial ori-
gin) can exert a systemic impact on the brain microenvironment. The figure was prepared in part
using image vectors from Servier Medical Art (www.servier.com), licensed under the Creative
Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

pineal gland

vagus nerve

internal
carotid nerve
external superior
nodose carotid laryngeal nerve
ganglion nerve
C1 PARATHYROID
2
spinal chord

Fig. 3.29 Parasympathetic 3 GLAND


4 SCG
and sympathetic 5
innervation of the thyroid/ 6 THYROID
7 MCG GLAND
parathyroid glands in rats.
8 ICG
Reproduced with T1 inferior
permission from parasympathetic laryngeal nerve
Cardinali [40] ganglia
104 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

neurons located in the local ganglia and receiving preganglionic projections from
the vagus nerve; (c) peptidergic neurons, including substance P-containing neurons,
which are present in local ganglia [24].
Sympathetic nerve fibers arriving at the thyroid/parathyroid territory are not only
widely distributed in the vasculature, but also make physical contact with epithelial
cells of the thyroid follicles, the parafollicular C-cells, and the parathyroid cells.
Retrograde neuronal tracing demonstrated that neurons located in the caudal pole of
the SCG that project via the external carotid nerve of the SCG are the only source
of sympathetic innervation to the thyroid–parathyroid complex.
The neuroendocrine relevance of the SCG is underlined by the number of endo-
crine and neuroendocrine structures found in their territory [40]. Among these neu-
roendocrine structures, besides the thyroid–parathyroid glandular complex, the
pineal gland, the medial basal hypothalamus/pituitary complex, and the carotid bod-
ies are found. Sympathetic nerve fibers arising from the SCG are also widely dis-
tributed in the submaxillary/mandibular salivary glands, the oral and laryngeal
cavity, the pial vessels, and eye structures such as the cornea, the iris, the nictitating
membrane, and Müller’s muscles.
Unlike other paravertebral and prevertebral ganglia of the sympathetic chain,
the sympathetic ganglia of the cervical region lack communicating branches and,
consequently, the preganglionic fibers reach the SCG from lower segments of the
sympathetic chain. Postganglionic sympathetic fibers leave the SCG in two ways
(Fig.  3.29). The internal carotid nerve pathway is followed by the postgangli-
onic fibers innervating the intracranial structures, such as the pineal, the median
eminence, the adeno- and neurohypophysis, and the choroid plexus. The external
carotid nerve is the neural path by which the thyroid and the parathyroid glands
and immune structures such as the submaxillary lymph nodes are innervated. In
this case, some of the innervating neurons are located in the middle and/or lower
cervical sympathetic ganglia and send their axons through the SCG and the external
carotid nerve [24].
Concerning the hypothalamic–pituitary–thyroid axis, the relevance of SCG is
best explained in terms of central and peripheral effects of the sympathetic nerve
terminals, as thyrotropin (TSH) release and the thyroid response to exogenous TSH
was inhibited during the increased release of NE from degenerating nerve terminals
shortly after SCGx [24]. Therefore, in the presence of normal or elevated levels of
TSH, NE release from local sympathetic nerves provides a negative signal for the
release of thyroxine. The daily fluctuations in the concentration of thyroxine in
serum and the thyroid content of catecholamines in rats were consistent with this
modulatory function of peripheral NE in the secretion of thyroid acini.
In rats studied 2–4 weeks after SCGx, greater growth (goiter response) was iden-
tified. The effect of promoting goiter was ipsilateral to sympathetic denervation.
The compensatory growth of the remaining lobe after a hemithyroidectomy per-
sisted in the absence of the anterior pituitary gland, an effect that was completely
blocked by an ipsilateral SCGx, indicating its essential neural nature [24].
Preganglionic parasympathetic input to local thyroid neurons derives from the
dorsal motor nucleus of the vagus and, through the nodose ganglion, is conveyed via
Local Autonomic Projections in Neuroendocrine Communication 105

the inferior laryngeal nerves and the thyroid nerves (Fig. 3.29). Before entering the
thyroid gland, the nerves display two types of ganglionic formation known as the
laryngeal ganglion and the thyroid ganglion. Cholinergic and peptidergic nerve
fibers arising from central nuclei and/or coming from local ganglia enter the thy-
roid, innervating blood vessels and endocrine structures. Indeed, parasympathetic
fibers distributed in the thyroid and parathyroid gland were retrogradely traced to
the medulla oblongata labeled the dorsal nucleus of the vagus [24].
In rats subjected to unilateral parasympathetic decentralization by ipsilateral
inferior laryngeal nerve section, a low compensatory growth of the lobe after
hemithyroidectomy was found. In hypophysectomized rats, section of the inferior
laryngeal nerve produced an additional involution of the thyroid gland. Thus, the
intact parasympathetic nerves are needed to maintain adequate trophism of the thy-
roid gland [24].
Concerning thyroid C and parathyroid cells, at the time of the increased NE
release during nerve degeneration shortly after SCGx, both a decrease in the maxi-
mum release of calcitonin and a delay to reach that maximum after stimulation with
an injection of calcium gluconate were observed. A similar inhibitory effect was
seen as far as a hypocalcemic-stimulated parathyroid hormone (PTH) release.
Blocking α-adrenoceptors suppressed the inhibition of C and parathyroid cell
response during post-SCGx degeneration, whereas ß-adrenoceptor blockade did not
affect the release of calcitonin or PTH, although it was effective in partially revers-
ing the activity of α-adrenoceptor blockade. The results point to a significant inhibi-
tory effect of sympathetic nerves on calcitonin and PTH release [24].
To study the effect of regional parasympathectomy on the release of calcitonin
and PTH animals with inferior laryngeal or thyroid nerves, sections were carried
out. After the injection of calcium chloride, a greater increase in serum calcitonin
and lower blood calcium levels than in controls were seen. PTH levels after a hypo-
calcemia challenge were also higher in parasympathectomized rats. Thus, the thy-
roid parasympathetic innervation exerts an inhibitory influence on calcitonin
secretion by C-cells. In summary, these few examples in the thyroid–parathyroid
territory indicate that the nerves supplying the endocrine glands were alternative
routes through which the brain communicates with them [24].
Similar phenomena occur for regional autonomic nerves in the adrenal glands,
the gonads, and pancreatic islets. For example, extrapituitary mechanisms of adre-
nal cortical control, including sympathetic neural activity, have been implicated in
controlling the amplitude of the cortisol awakening response, a diagnostic index of
hypothalamic pituitary adrenal activity in humans. In addition, increases in sympa-
thetic neural tone have been implicated in polycystic ovary syndrome, a leading
cause of female infertility [41, 42]. In the ovary, the superior ovarian nerve inhibits
ovarian estradiol secretion by activation of α2 adrenoceptors [43].
Intracellular glucose signaling pathways control the secretion of glucagon and
insulin by pancreatic islet α- and β-cells respectively [44]. In addition, glucose also
indirectly controls the secretion of these hormones through the regulation of the
ANS that richly innervates this endocrine organ. Both parasympathetic and sympa-
thetic nervous systems also have an impact on postnatal endocrine pancreas
106 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

development and plasticity in adult animals. Defects in these autonomic regulations


impair β-cell mass expansion during the weaning period and β-cell mass adaptation
in adult life. Both branches of the ANS also regulate glucagon secretion. In type 2
diabetes mellitus, impaired glucose-dependent autonomic activity causes the loss of
the cephalic and first phases of insulin secretion, and impaired suppression of glu-
cagon secretion in the postabsorptive phase [45].
Collectively, these examples offer compelling evidence for the capability of
autonomic neural activity to alter the functional sensitivity of endocrine glands. By
using a dual viral tracing technique that enabled simultaneous exploration of the
neural circuits of two organs, the following observations were made [46]:

( 1) There are autonomic neurons exclusively innervating a given endocrine organ.


(2) A left-sided predominance in the supraspinal innervation of the endocrine
glands (adrenal, ovary) studied so far.
(3) Viral co-infection of neurons, i.e., special neuronal populations coexist in
different brain areas that are trans-synaptically connected with both paired
endocrine and non-endocrine organs, endocrine glands and non-endocrine
organs.

The number of common neurons seems to be related to the need to coordinate the
action of different systems. The data on co-infection of neurons suggest that the
CNS might have the capacity to coordinate different organ functions via common
brain neurons, providing supraspinal innervation of the organs [46].
Therefore, the concept that nerves innervating endocrine glands constitute alter-
native pathways through which the brain communicates with the endocrine glands
must be emphasized. The organization of these pathways, as for many other sensory
and motor pathways of the central nervous system, is essentially hierarchical and in
parallel (Fig. 3.30).

 he ANS Contributes to the Maintenance of Healthy


T
Bone Tissue

The skeleton is a specialized and dynamic organ that undergoes continuous regenera-
tion. It consists of highly specialized cells, mineralized and nonmineralized connec-
tive tissue matrix, and spaces that include the bone marrow cavity, vascular canals,
canaliculi, and lacunae. Removal of bone (resorption) is the task of osteoclasts,
whereas formation of new bone is the task of osteoblasts. Tight molecular control of
bone remodeling is vital for the maintenance of appropriate physiology and microar-
chitecture of the bone, providing homeostasis, also at the systemic level. The process
of remodeling is regulated by the rich innervation of the skeleton, as it is the source
of various growth factors, neurotransmitters, and hormones regulating the functions
of the bone [47].
A ubiquitous autonomic innervation of all periosteal surfaces exists and its disruption
may affect bone remodeling control and lead to various bone diseases (e.g., osteogenesis
The ANS Contributes to the Maintenance of Healthy Bone Tissue 107

LIMBIC SYSTEM
(frontal and mesencephalic poles)

Autonomic descendent pathways

HYPOTHALAMUS ICN
Cervical sympathetic
Hypophysiotropic area SON, PVN ganglia

Median Thoracic and lumbar


eminence sympathetic gangalia
ICN

ANTERIOR POSTERIOR ECN


Vagus nerve
HYPOPHYSIS HYPOPHYSIS

ENDOCRINE GLAND
INNERVATION
HORMONE RELEASE

Fig. 3.30  Hierarchical and parallel organization of neurohormonal pathways under endocrine
control. ICN internal carotid nerve, ECN external carotid nerve, SON supraoptic nucleus, PVN
paraventricular nucleus. Reproduced with permission from Cardinali [40]

imperfecta). Patients with neurological disorders exhibit localized osteopenia and bone
fragility, altered fracture healing, and excessive callus formation [48].
An intact ANS contributes to the maintenance of healthy bone tissue. The role of
the ANS in abnormal bone formation and its association with clinical diseases has
been proposed, for example, postmenopausal osteoporosis, adolescent idiopathic
scoliosis, and depression-induced osteoporosis [49]. The long bones of the upper
extremities receive nerve supply from the brachial plexus, which then branches to the
median nerve to innervate the humerus and the ulnar and radian nerves, which supply
the forearm bones. Sympathetic innervation of the lower limbs originates in the lum-
bar plexus, which supplies the femoral and deep saphenous nerves to the femur, and
the tibial, medial, and popliteal nerves to the tibia and fibula. Basivertebral nerves in
the spine supply intraosseous autonomic innervations of the vertebral bodies [50].
In the author’s laboratory, studies were focused on results obtained in the field of
projection of the SCG territory, the first sympathetic ganglion of the paravertebral
chain, which includes the mandibular bone [51]. To assess in an anatomically spe-
cific way the effect of local sympathectomy on bone physiology, we examined the
effect of unilateral SCGx on growth, bone mineral content, and bone mineral
108 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

density of the ipsi- and contralateral hemimandibles. Total bone mineral content of
the hemimandibular bones decreased on the side ipsilateral to the SCGx. Bone min-
eral density (i.e., the bone mineral content/bone mineral area ratio) was also signifi-
cantly lower in the hemimandible ipsilateral to the SCGx.
The ANS is one of the factors modifying tooth eruption. Teeth are innervated by
unmyelinated sympathetic axons originating in the ipsilateral SCG, and by unmyelin-
ated and small myelinated sensory axons, most of them terminal branches of larger
parent axons in the trigeminal nerve. The sympathetic nerve endings contain NE,
whereas sensory dental axons contain SP-like immunoreactivity. Other neuropeptides
are also present in dental nerves, such as VIP and NPY. Adrenergic nerves end at the
odontoblast/predentin border, in the predentin adjacent to the odontoblast processes, and
as free endings in the middle part of the predentin. In situations of degenerative auto-
nomic neuropathy, an overall marked reduction in pulpal innervation, with an absence
of large nerve bundles and the subodontoblastic plexus, has been reported in humans.
We examined the effect of a unilateral SCGx on the eruption rate of ipsilateral
and contralateral rat incisors in two experimental situations: (a) without any further
manipulation; (b) after the ipsi- or contralateral lower rat incisor had been cut out of

unilateral sympathetic denervation

ipsilateral unaffected mandible unaffected


mandible BMD thickness & growth tooth growth

unilateral mechanical stress

ipsilateral ipsilateral mandible tooth growth


mandible BMD thickness & growth

+unilateral sympathetic denervation

mandible further increase


BMD of mandible BMD

mandible further increase of mandible


thickness & growth thickness & growth

tooth growth tooth growth

(unimpeded incisor eruption) (impeded incisor eruption)

Fig. 3.31  Diagram summarizing the effect of unilateral sympathetic denervation on rat mandibu-
lar bone. The effect of a unilateral superior cervical ganglionectomy on bone mineral density
(BMD), morphometric assessment of mandible thickness, and the growth and eruption rate of
ipsilateral and contralateral incisors were analyzed in two experimental situations: (1) without any
further manipulation; (2) after a unilateral mechanical stress given by cutting the ipsi- or contralat-
eral lower incisor out of occlusion
References 109

occlusion (Fig. 3.31). In a first experiment, the eruption rate of ipsilaterally dener-


vated incisors was similar to that of contralaterally innervated incisors, when
assessed for up to 28 days after surgery. In a second experiment, under conditions
of unilateral unimpeded eruption of incisors performed ipsilaterally or contralater-
ally to a unilateral SCGx, a significantly lower eruption rate of denervated incisors
on the impeded eruption side and a significantly higher eruption rate of denervated
incisors on the unimpeded side were observed (Fig. 3.31).
The results indicated that incisor eruption is not modified by local sympathetic
denervation unless the contralateral lower rat incisor was cut out of occlusion.
When this was done, sympathetically denervated incisors exhibited higher eruption
rates on the unimpeded eruption side and lower eruption rates on the impeded side.
Therefore, normal presynaptic neural activity, and not merely an augmented trans-
mitter release, seems to be a prerequisite for maintaining a minimal normal unim-
peded incisor eruption and for maintaining the unimpeded eruption to attain
abnormally high velocities under conditions of stimulated incisor growth. This
dual role of the ANS on organ activity is found in several other tissues, such as
endocrine and immune tissues. For example, a normal sympathetic output main-
tains basal cell proliferation in lymph nodes and curtails overstimulation following
an antigen challenge [52].

References
1. Cardinali DP. Neurociencia Aplicada. Sus Fundamentos. Editorial Médica Panamericana:
Buenos Aires; 2007.
2. Goldstein DS. Chapter 2 – Differential responses of components of the autonomic nervous
system. In: Buijs RM, Swaab D, editors. Handbook of clinical neurology, autonomic nervous
system. New York: Elsevier; 2013. p. 13–22.
3. Waxman SG. Chapter 20. The autonomic nervous system. In: Clinical neuroanatomy. 27th ed.
New York: McGraw-Hill; 2013.
4. Barrett KE, Barman SM, Boitano S, Brooks HL. Autonomic nervous system. In: Ganong’s
review of medical physiology. 25th ed. New York: McGraw-Hill Education; 2016.
5. Fernandez-Gil MA, Palacios-Bote R, Leo-Barahona M, Mora-Encinas JP. Anatomy of the
brainstem: a gaze into the stem of life. Semin Ultrasound CT MR. 2010;31:196–219.
6. De Groat WC, Griffiths D, Yoshimura N. Neural control of the lower urinary tract. Compr
Physiol. 2015;5:327–96.
7. Westfall TC, Westfall DP. Neurotransmission: the autonomic and somatic motor nervous sys-
tems. In: Brunton LL, Chabner BA, Knollmann BC, editors. Goodman & Gilman's: the phar-
macological basis of therapeutics. 12th ed. New York: McGraw-Hill Education; 2011. p. 1–58.
8. Ray A, Chakraborti A, Gulati K. Current trends in nitric oxide research. Cell Mol Biol.
2007;53:3–14.
9. Wehrwein EA, Joyner MJ. Chapter 8 – Regulation of blood pressure by the arterial baroreflex
and autonomic nervous system. In: Buijs RM, Swaab D, editors. Handbook of clinical neurol-
ogy, autonomic nervous system. Elsevier; 2013, p 89–102.
10. LeBlond RF, Brown DD, Suneja M, Szot JF. The nervous system. In: DeGowins diagnostic
examination. 10th ed. New York: McGraw-Hill Education; 2015.
11. Kibble JD, Halsey CR. Neurophysiology. In: Medical physiology: the big picture. New York:
McGraw-Hill Education; 2015.
110 3  First Level: Peripheral Sympathetic and Parasympathetic Nervous System

12. Burnstock G. Chapter 3 – Cotransmission in the autonomic nervous system. In: Buijs RM,
Swabb D, editors. Handbook of clinical neurology. Autonomic nervous system. New York:
Elsevier; 2013, p 23–35.
13. Herring N. Autonomic control of the heart: going beyond the classical neurotransmitters. Exp
Physiol. 2015;100:354–8.
14. Hahn MK. Chapter 8 – Antidepressant-sensitive norepinephrine transporters: structure and
regulation. In: Biaggioni I, Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic
nervous system. Third ed. San Diego: Academic; 2012. p. 49–50.
15. Sherrington C. The integrative action of the nervous system. Cambridge: Cambridge University
Press; 1906.
16. Craig AD. Chapter 9 – Cooling, pain, and other feelings from the body in relation to the
autonomic nervous system. In: Buijs RM, Swabb D, editors. Handbook of clinical neurology,
autonomic nervous system. New York: Elsevier; 2013. p. 103–9.
17. Benarroch EE. Central autonomic network. Functional organization and clinical correlations.
Leander: Futura Publishing; 1997.
18. Longhurst JC, Fu LW. Chapter 35 – Cardiac and other visceral afferents. In: Biaggioni I,
Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous system. Third ed.
San Diego: Academic; 2012. p. 171–6.
19. Zhuo M. Contribution of synaptic plasticity in the insular cortex to chronic pain. Neuroscience.
2016;338:220–9.
20. Verriotis M, Chang P, Fitzgerald M, Fabrizi L. The development of the nociceptive brain.
Neuroscience. 2016;338:207–19.
21. Gonzalez Burgos G, Biali FI, Cardinali DP. Picrotoxin-sensitive receptors mediate

γ-aminobutyric acid-induced modulation of synaptic plasticity in rat superior cervical gan-
glion. Brain Res. 1997;751:148–51.
22. Landa ME, Gonzalez Burgos G, Cardinali DP. In vitro effects of thyroxine on cholinergic neuro-
transmission in rat sympathetic superior cervical ganglion. Neuroendocrinology. 1991;54:552–8.
23. Shireen E. Experimental treatment of antipsychotic-induced movement disorders. J Exp

Pharmacol. 2016;8:1–10.
24. Romeo HE, Cardinali DP. The autonomic nervous system of the cervical region as a channel
of neuroendocrine communication. Front Neuroendocrinol 1991;12:278–97.
25. Seth JH, Panicker JN, Fowler CJ. Chapter 10 – The neurological organization of micturition.
In: Buijs RM, Swabb D, editors. Handbook of clinical neurology, autonomic nervous system.
New York: Elsevier; 2013. p. 111–7.
26. Ranson RN, Saffrey MJ. Neurogenic mechanisms in bladder and bowel ageing. Biogerontology.
2015;16:265–84.
27. Low PA, Engstrom JW. Disorders of the autonomic nervous system. In: Kasper D, Fauci A,
Hauser S, Longo D, Jameson JL, Loscalzo J, editors. Harrison's principles of internal medi-
cine. 19th ed. New York: McGraw-Hill Education; 2015.
28. McDougal DH, Gamlin PD. Autonomic control of the eye. Compr Physiol. 2015;5:439–73.
29. Joos KM, Melson MR. Chapter 49 – Control of the pupil. In: Biaggioni I, Burnstock G,
Low PA, Paton JFR, editors. Primer on the autonomic nervous system. Third ed. San Diego:
Academic; 2012. p. 239–42.
30. Clement P, Giuliano F. Anatomy and physiology of genital organs – men. Handb Clin Neurol.
2015;130:19–37.
31. Jobling P. Autonomic control of the urogenital tract. Auton Neurosci. 2011;165:113–26.
32. Avetisyan M, Schill EM, Heuckeroth RO. Building a second brain in the bowel. J Clin Invest.
2015;125:899–907.
33. Kibble JD, Halsey CR. Gastrointestinal physiology. In: Medical physiology: the big picture.
New York: McGraw-Hill Education; 2015.
34. Camilleri M. Chapter 42 – Gastrointestinal function. In: Biaggioni I, Burnstock G, Low PA,
Paton JFR, editors. Primer on the autonomic nervous system. Third ed. San Diego: Academic;
2012. p. 205–9.
References 111

35. Ochoa-Cortes F, Turco F, Linan-Rico A, Soghomonyan S, Whitaker E, Wehner S, Cuomo R,


Christofi FL. Enteric glial cells: a new frontier in neurogastroenterology and clinical target for
inflammatory bowel diseases. Inflamm Bowel Dis. 2016;22:433–49.
36. Barrett KE. Gastrointestinal physiology. Second ed. New York: McGraw-Hill Education; 2014.
37. Berg CJ, Kaunitz JD. Gut chemosensing: implications for disease pathogenesis. F1000Res.
2016;5:2424.
38. Erny D, Hrabe de Angelis AL, Prinz M. Communicating systems in the body: how microbiota
and microglia cooperate. Immunology. 2017;150:7–15.
39. Mayer EA, Tillisch K, Gupta A. Gut/brain axis and the microbiota. J Clin Invest.

2015;125:926–38.
40. Cardinali DP. Ma Vie en Noir. Fifty years with melatonin and the stone of madness. Switzerland:
Springer; 2016.
41. Lansdown A, Rees DA. The sympathetic nervous system in polycystic ovary syndrome: a
novel therapeutic target? Clin Endocrinol. 2012;77:791–801.
42. Engeland WC. Chapter 4 – Sensitization of endocrine organs to anterior pituitary hormones by
the autonomic nervous system. In: Buijs RM, Swaab D, editors. Handbook of clinical neurol-
ogy, autonomic nervous system. New York: Elsevier; 2013. p. 37–44.
43. Uchida S, Kagitani F. Autonomic nervous regulation of ovarian function by noxious somatic
afferent stimulation. J Physiol Sci. 2015;65:1–9.
44. Rodriguez-Diaz R, Caicedo A. Neural control of the endocrine pancreas. Best Pract Res Clin
Endocrinol Metab. 2014;28:745–56.
45. Thorens B. Neural regulation of pancreatic islet cell mass and function. Diabetes Obes Metab.
2014;16(Suppl 1):87–95.
46. Gerendai I, Toth IE, Boldogkoi Z, Halasz B. Recent findings on the organization of central
nervous system structures involved in the innervation of endocrine glands and other organs;
observations obtained by the transneuronal viral double-labeling technique. Endocrine.
2009;36:179–88.
47. Niedzwiedzki T, Filipowska J. Bone remodeling in the context of cellular and systemic regula-
tion: the role of osteocytes and the nervous system. J Mol Endocrinol. 2015;55:R23–36.
48. Binks S, Dobson R. Risk factors, epidemiology and treatment strategies for metabolic bone
disease in patients with neurological disease. Curr Osteoporos Rep. 2016;14:199–210.
49. Ji-Ye H, Xin-Feng Z, Lei-Sheng J. Chapter 14 – Autonomic control of bone formation: its
clinical relevance. In: Buijs RM, Swaab D, editors. Handbook of clinical neurology, autonomic
nervous system. New York: Elsevier; 2013. p. 161–71.
50. Elefteriou F, Campbell JP. Chapter 53 – Autonomic innervation of the skeleton. In: Biaggioni
I, Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous system. Third
ed. San Diego: Academic; 2012. p. 257–9.
51. Boggio V, Ladizesky MG, Cutrera RA, Cardinali DP. Autonomic neural signals in bone: physi-
ological implications for mandible and dental growth. Life Sci. 2004;75:383–95.
52. Cardinali DP, Cutrera RA, Esquifino AI. Psychoimmune neuroendocrine integrative mecha-
nisms revisited. Biol Signals Recept. 2000;9:215–30.
Second Level: The Brainstem
4

Abstract
The sympathetic and parasympathetic divisions of the ANS are integrated and
regulated by a hierarchy of central structures. This central autonomic neural net-
work has been identified and mapped in recent years and functional neuroimag-
ing data largely support it. Such a network involves reciprocal connections, both
direct and indirect, between the two efferent systems and with more cranial neu-
ral clusters located in the brainstem, hypothalamus, limbic system, and the insu-
lar cortex. As examples of that organization, this Chapter discusses the neural
mechanisms that control arterial blood pressure, heart rate, breathing, immunity,
and gastrointestinal function and how they vary in the three body configurations
(wakefulness, NREM sleep, REM sleep).

Keywords
Brainstem • Cardiorespiratory homeostasis • Carotid body • Cerebellum
• Gastrointestinal function • Immunity • Migrating motor complex • Neural
regulation of cardiovascular function • Neural regulatory system of breathing
• Nucleus tractus solitarium • Parabrachial nucleus • Reticular formation

Objectives
After studying this chapter, you should be able to:

• Describe the location of forebrain and brainstem neurons that are compo-
nents of central autonomic pathways.
• Describe the neural mechanisms that control arterial BP and heart rate, includ-
ing the receptors, afferent and efferent pathways, central integrating pathways,
and the effector mechanisms involved, and how they vary in the three body
configurations (wakefulness, NREM sleep, REM sleep) during a 24-h cycle.
• Identify the location and functions of the dorsal and ventral groups of
respiratory neurons, the pneumotaxic center, and the apneustic center in

© Springer International Publishing AG 2018 113


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1_4
114 4  Second Level: The Brainstem

the brainstem and how they vary in the three body configurations (wakeful-
ness, NREM sleep, REM sleep) during a 24-h cycle.
• Describe the participation of the cerebellum in the autonomic posture.
• Identify the roles and mechanisms of innate, acquired, humoral, and cel-
lular immunity and how they vary in the three body configurations (wake-
fulness, NREM sleep, REM sleep) during a 24-h cycle.
• Describe the functional significance of the gastrointestinal system, and its
roles in nutrient assimilation, excretion, and immunity and how they vary
in the three body configurations (wakefulness, NREM sleep, REM sleep)
during a 24-h cycle.

 t the Brainstem, Various Complex Autonomic Responses


A
Are Coordinated

The sympathetic and parasympathetic divisions of the ANS are integrated and regulated
by the hierarchy of central structures plotted in Fig. 1.2. This central autonomic neural
network has been clearly identified and mapped in recent years and functional neuroim-
aging data largely support it. Such a network involves reciprocal connections, both direct
and indirect, between the two efferent systems and with more cranial neural clusters
located in the brainstem, hypothalamus, limbic system, and brain (e.g. insular cortex).
The brainstem is the region between the spinal cord and the diencephalon. It
comprises the medulla oblongata, the pons, and the midbrain. Many sensory and
motor nuclei and the different components of reticular formation are in the brain-
stem and lesions affecting this area have profound motor, sensory, and consciousness-­
related consequences, which is why a discussion of its organization is necessary for
an understanding of the physiology of the ANS.
The brainstem is a site of somatic and autonomic motoneurons and different
types of sensory neurons that are organized into anatomically identifiable columns
corresponding to the cranial nerves [1]. The cranial nerves have three defined func-
tions: (a) to provide somatosensory and motor innervation to the neck and head; (b)
to provide innervation to the sense organs; (c) to provide parasympathetic pregan-
glionic innervation to autonomic ganglia that control visceral function.
As in the spinal cord, somatic motor efferent neurons, autonomic efferent neu-
rons (preganglionic) and second-order sensory afferent neurons coexist in the brain-
stem. Second-order sensory neurons receive afferent fibers from somatic or visceral
primary sensory neurons located in ganglia outside the brainstem or in the sense
organs (the only exception to this rule is the mesencephalic nucleus of the V pair,
which contains primary sensory neurons).
What is particular in the brainstem is that, in the case of motor and sensory, somatic
and visceral neurons of the cranial nerves are subdivided into anatomically segregated
functional groups. This subdivision consists of two organizational principles:

1. There are three types of motor neurons in the brainstem: somatic motoneurons,
special visceral motoneurons, and general visceral motoneurons. Somatic
At the Brainstem, Various Complex Autonomic Responses Are Coordinated 115

motoneurons innervate the muscles of the face and neck derived from the myo-
tome, with the same origin as the rest of the skeletal musculature. They send
their fibers through the III, IV, VI, and XII cranial nerve pairs (voluntary ocular
musculature and tongue). Special visceral motoneurons also innervate the stri-
ated musculature of the face and neck, but in this case, the muscles are derived
from the gill arches (mastication, facial expression, larynx, pharynx). They send
their fibers through V, VII, IX, X, and XI cranial nerve pairs. General visceral
motoneurons provide parasympathetic preganglionic autonomic innervation.
They send their fibers through the III, VII, IX and X cranial nerve pairs.
2. There are four types of second-order sensory neurons in the brainstem: general
somatic sensitivity, special somatic sensitivity, general visceral sensitivity, and
special visceral sensitivity respectively. General somatic sensitivity comprises
the sensitivity of face skin and oral and pharyngeal mucous membranes, and
proprioception (V, VII, IX and X cranial nerve pairs). The special somatic sensi-
tivity comprises that of the inner ear (VIII cranial nerve pair). The general vis-
ceral sensitivity comprises that coming from thoracic and abdominal organs
(IX and X cranial nerve pairs). The special visceral sensitivity originates in the
gustatory corpuscles (VII, IX, X cranial nerve pairs).

These seven functionally differentiable neuronal groups are grouped in columns


along the brainstem, as represented in Fig. 4.1:

Afferent: Motor:
Visceral Visceral Special somatic afferent: VIII
Somatic Somatic
General somatic afferent:
V, VII, IX, X

General and special visceral


afferent: VII, IX, X
Sulcus

General visceral motor:


III, VII, IX, X

Special visceral motor:


V, VII, IX, X, XI

General somatic motor:


III, IV, VI, XII

Fig. 4.1  Nuclei of cranial nerves and column organization of the brainstem. Modified with
­permission from Cardinali [3]
116 4  Second Level: The Brainstem

• The somatic motor column, which contains motor neurons that innervate the
extraocular and tongue muscles (in rostrocaudal order).
• The special visceral motor column containing motor neurons that supply the
muscles of the larynx, pharynx, face, and jaw. They are, in descending order: a)
motor nucleus of the V cranial nerve pair (mastication); (b) motor nucleus of the
VII cranial nerve pair (facial expression); (c) ambiguous nucleus (IX, X cranial
nerve pairs; speech, swallowing); (d) nucleus of the spinal accessory nerve (XI
cranial nerve pair).
• The general visceral motor column, which contains the neuronal bodies of para-
sympathetic preganglionic neurons: (a) Edinger–Westphal nucleus (III), pregan-
glionic to ciliary ganglion; (b) upper salivary nucleus (VII), preganglionic to the
sphenopalatine ganglion (lacrimal gland) and submaxillary ganglion (sublingual
and submaxillary glands); (c) inferior salivary nucleus (IX), preganglionic to the
otic ganglion (parotid gland); (d) dorsal motor nucleus of the X pair: parasympa-
thetic preganglionic to the thoracic and abdominal organs.
• The general visceral afferent columns and the special visceral afferent columns
form the nucleus tractus solitarium (NTS), which comprises two parts: (a) ros-
tral, which is the relay site for taste and general visceral afferents from the diges-
tive tract; (b) caudal, which receives the afferents of the carotid body and the
general visceral afferent from the bronchi and lungs. The perikarya of these pri-
mary sensory neurons are in ganglia associated with the VII, IX and X cranial
nerve pairs, and their central extensions form the solitary tract, ending in the
secondary neurons that form the solitary nucleus. From here, the portion that
mediates sensitivity projects to the thalamus, whereas that corresponding to car-
diorespiratory regulation establishes direct contact with the reticular and indirect
formation with the limbic system (through the parabrachial nucleus, PBN).
• The special somatic afferent column contains the vestibular and cochlear second-
ary relay neurons (VIII cranial nerve pair).
• The general somatic afferent column contains secondary neurons of oral and
oropharyngeal somatic sensitivities. It consists of three divisions separated ros-
trocaudally into: (a) the mesencephalic nucleus of the V cranial nerve pair (pro-
prioception of facial and mandibular muscles, the only case of primary sensory
neurons in a brainstem) receives afferents from mucous and muscles; (b) the
main nucleus of the V cranial nerve pair receives afferents from mucous and
muscles; (c) the spinal nucleus of the V cranial nerve pair.

In summary, there are three basic principles in the organization of the cranial nerves:

1 . Most motor nuclei are linked to individual cranial nerves.


2. Relay nuclei in the sensory pathways receive fibers from various cranial nerve
pairs, e.g., the NTS receives taste information brought by the VII, IX, and X
pairs, following the general principle that somatotopic sensory location ends in
the same central area, regardless of the route it follows.
3. Neurons with similar functional properties occupy the same positions in the
brainstem.
At the Brainstem, Various Complex Autonomic Responses Are Coordinated 117

The NTS, located in the dorsal part of the medulla oblongata, and secondarily the
NPB, constitute the main centers of central relay of visceral sensory information
(Fig. 4.2). The main afferent inputs of the NTS comprise fibers from the cardiovas-
cular, respiratory, gastrointestinal, and neuroendocrine–immune systems, along
with gustatory collaterals and somatoesthetic sensitivity.
The NTS is the primary integrative center for cardiovascular control and other
autonomic functions in the CNS. The NTS has long been identified as a site where
the first synapse of the baroreceptor reflex is located. Therefore, the NTS, in addition
to other key central nuclei in the hypothalamus and other forebrain regions, play
important roles in mediating cardiovascular responses to acute stresses. The NTS not
only integrates convergent information, but itself is the site of substantial modula-
tion. Evidence demonstrated several neurotransmitters or neuromodulators such as
Glu, NE, E, ACh, 5-HT, NO, angiotensin II, arginine vasopressin (AVP), β-endorphin,
enkephalins, NPY, adenosine and insulin [2].
By means of neural circuitry mapping techniques a viscerotopic map was traced
in the NTS and in the neuronal groups of the dorsal nucleus of the X cranial nerve
pair. These data support the notion that vagal information in the gut–brain axis
maintain specific pathways from the gastrointestinal tract through individualized
vagal branches. The viscerotopic organization of the gut–brain afferent loop corre-
lates with a parallel map in the efferent limb of the brain–gut axis forming a network

Oral cavity
VII Tongue NTS
IX Tonsils Thalamus
Lymph nodes Parvicellular

Dorsal
Heart
Parabrachial nucleus
Lungs Ventral

Pharynx
Larynx Caudal Hypothalamus
Thyroid gland Amygdala
Parathyroid gland BST
X Thymus Insula
Heart
Lungs Ambiguous &
Esophagus Dorsomedial n.
Stomach Intermediolateral
Adrenal gland column of the
Intestine Ventrolateral spinal cord
Liver medulla

Fig. 4.2  Visceral sensory input to the nucleus tractus solitarius (NTS) brain relevant to the control
of food intake and energy balance. All along the alimentary canal, various mechano- and chemo-
sensors are located that transmit food- and nutrition-related signals via primary visceral afferents
in the trigeminal (V), facial (VII), glossopharyngeal (IX), and vagus nerve (X) to the brainstem
118 4  Second Level: The Brainstem

ideally suited to mediate cephalic, gastric, hepatic, and intestinal reflexes accompa-
nying the food intake.
The NTS is subdivided into several subnuclei, which play different functional
roles according to the visceral afferents they receive (Fig. 4.2). Thus, cardiovascular
afferences terminate predominantly in the dorsal subnucleus, pulmonary afferents
in the ventral subnucleus, and gustatory input in the parvicellular subnucleus,
whereas the caudal commissural subnucleus receives afferences from all visceral
components. Projections from the NTS transmit a wide range of visceral informa-
tion to higher nuclei, and establish circuits of reciprocal connection with regions of
the reticular formation (particularly PBN and periaqueductal gray matter), hypo-
thalamus, amygdala, limbic system, and insular cortex.
The descending efferent pathways of these circuits innervate, directly or through
synapses in the NTS, the parasympathetic preganglionic neurons and the sympa-
thetic intermediolateral columns. The descending autonomic pathways mainly
travel ipsilaterally in the anterior portion of the lateral cord. Using neuroimaging
(functional magnetic resonance imaging, fMRI; positron emission tomography,
PET) the asymmetry and lateralization of the ANS function was verified.
The PBN plays a key role in the central autonomic network, as an intermediate
between the brainstem reflex control and behavioral control systems at the dience-
phalic and telencephalic levels (Fig. 4.2). The most relevant parts of the PBN for con-
trolling food intake and energy homeostasis are the medial and lateral subdivisions.
The neurons located in the upper lateral nucleus of the lateral subdivision of PBN are
particularly important for the control of food intake, and send dense projections to the
ventromedial nucleus (VMN) of the hypothalamus related to the control of satiety. The
upper lateral nucleus of the PBN has a large concentration of CCK-expressing neurons
that are related to the induction of satiety and are activated by circulating leptin.
The central projections on the preganglionic autonomic neurons come mainly
from the hypothalamus and various brainstem nuclei (dorsal raphe nucleus, DRN,
locus coeruleus, LC, ventrolateral medulla reticular formation). The role of these
central pathways is to maintain a tonic state of excitability of preganglionic auto-
nomic neurons, modulate segmental reflexes, and generate organized patterns of
activity in different functional groups of preganglionic neurons.
In the viscera, there are numerous receptors, such as osmoreceptors, barorecep-
tors, glucoreceptors, etc., which respond to changes in the internal environment.
Their afferents participate in different autonomic reflexes of homeostatic impor-
tance. These receptors are the basis of interoception, and presumably they can be
the physical substrate of what from a psychological point of view is called the
preconscious.
A key question is whether nuclei such as the NTS or PBN are predominantly
relay nuclei, or do they transform incoming visceral information? It is noteworthy
that contrary to popular anatomical terminology, there is no need for relay nuclei
because axon potentials do not need to be boosted by the synapse. A chemical syn-
apse on its own simply adds an unavoidable delay.
Neurons form synapses in a nucleus to transform or change the incoming signal.
The information obtained in the nuclei of the dorsal column (Goll and Burdach
At the Brainstem, Various Complex Autonomic Responses Are Coordinated 119

nuclei) in the somatosensory system indicates the existence of several transforma-


tions. On the one hand, there is differential convergence that results in a distorted
viscerotopic representation of the body. In addition, because of the lateral inhibition,
the discrimination of the stimuli is accentuated. As in other sensory territories, stimu-
lus in the center activates NTS neurons whereas stimulus in the surround, through
inhibitory feedback, inhibits the same NTS neurons. Another function is cortical
gating: corollary discharge selectively gates input based on motor output [3].
The existence of the gating control exerted by superior mesencephalic, dience-
phalic, and cortical structures on the afferent information that arrives from receptors
is exemplified by pain. The main function of this input control (central analgesia) is
to suppress the irrelevant information originated in the periphery, allowing only that
of sufficient intensity to have an adaptive meaning to pass. As seen in Figs. 4.3 and
4.4, the convergence of visceral and cutaneous afferents on the same second-order
neural groups results in visceral pain being experienced on a portion of the cutaneous
surface (referred pain). This referred pain is verified in the portion of the cutaneous
surface of embryological origin similar to the affected viscera (dermatome) [4].

Modulatory
descendent
input

Skin

Muscle

Joint

Viscera

Fig. 4.3  The main function of this afferent control (central analgesia) is to suppress the irrelevant
information originating at the periphery, only allowing that of sufficient intensity to have an adap-
tive meaning to pass. Modified with permission from Cardinali [3]
120 4  Second Level: The Brainstem

Viscerotopic

Somatotopic Skin

Viscera

Preganglionic
ANS neuron

Spinothalamic tract
Motoneuron

Fig. 4.4  The convergence of visceral and cutaneous afferents on the same second-order neural
groups results in visceral pain being experienced on a portion of the cutaneous surface (referred
pain). Modified with permission from Cardinali [3]

 onoaminergic Systems in the Brainstem Modulate


M
24-h Rhythms in Physiological Function

The reticular formation is composed of neurons that do not correspond to the differ-
ent functional columns of the brainstem mentioned above. It is convenient from an
anatomical point of view to analyze the reticular formation in the medial–lateral
sense, thus distinguishing (a) the raphe nuclei at both sides of the medial line; (b) a
magnocellular region; (c) a parvicellular region.
Pons reticular formation and midbrain reticular formation have different functions.
Most neurons in the reticular formation display a great diversity of connections, dif-
fusely distributed, both with upper centers and toward the spinal cord (distribution “in
a spider web,” Fig. 4.5). Based on the neurotransmitter employed, the following neu-
ral groups are distinguished, most of them with locations in the brainstem:

1. Noradrenergic system: located in the LC and contiguous nuclei, it provides


innervation to the entire brain, including the spinal cord (Fig. 4.6). Except for a
small intrahypothalamic group of neurons, this is the only source of central nor-
adrenergic innervation. As discussed in Chap. 2, its role is fundamental in the
maintenance of wakefulness and its alteration is linked to emotional illnesses
such as depression, their spinal projection participating in muscular tone control
and gait generation. The different NE receptors have been analyzed in Chap. 3.
Monoaminergic Systems in the Brainstem Modulate 24-h Rhythms 121

“Spider web” distribution

Neuronal body

Fig. 4.5  “Spider web” neuron in a rat brainstem. Modified with permission from Cardinali [3]

Cingulate Cortex

Neocortex
Thalamus

Hypoth.

Locus
Olfactory bulb coeruleus Cerebellum

Fig. 4.6  Principal noradrenergic projections of the reticular formation. Modified with permission
from Cardinali [3]
122 4  Second Level: The Brainstem

2. Dopaminergic system: several cell groups are in the midbrain (e.g., the substantia
nigra pars compacta and the VTA, Fig. 4.7). They provide dopaminergic innervation
to the basal ganglia, hypothalamus, limbic system, and neocortex. Four major dopa-
minergic systems are found in the CNS: (a) nigrostriatal; (b) mesolimbic; (c) hypo-
thalamic tuberoinfundibular; (d) retinal. The nigrostriatal projection (neuronal bodies
in the substantia nigra pars compacta) participates in the regulation of function of the
basal ganglia, whereas the mesolimbic projection (neuronal bodies in the VTA, or
A10) is involved in emotional states, psychiatric diseases, and drug addiction. The
tuberoinfundibular system (neuronal bodies in the arcuate nucleus, ARC, of the
hypothalamus), participates in the control of prolactin (PRL) secretion. The retinal
dopaminergic neurons are a subgroup of amacrine interneurons. Six major types of
dopaminergic receptors have been identified and their different isoforms have been
cloned. These types are called D1, D2a, D2b, D3, D4, and D5. They all have the
seven-peptide hydrophobic sequences that indicate their association with G proteins.
Functionally, they can be differentiated by their nature into (a) excitatory (class D1,
comprising D1 and D5) associated with a stimulating G protein and with the activa-
tion of an adenylate cyclase; (b) inhibitory (class D2, comprising D2, D3, and D4)
associated with an inhibitory G protein and inhibition of an adenylate cyclase.

Frontal
cortex Striatum

Septum
VTA

Substantia
nigra
Olfactory bulb
Arcuate
nucleus

Fig. 4.7  Principal dopaminergic projections of the reticular formation. Modified with permission
from Cardinali [3]
Monoaminergic Systems in the Brainstem Modulate 24-h Rhythms 123

3. Serotoninergic system: comprising the DRN (a continuation of the periaqueductal


gray matter; Fig. 4.8). DRN are the origin of almost all serotonergic innervation
of the brain. The serotonergic receptors (5-HT-1, 5-HT-2, 5-HT-3, 5-HT-4, 5-HT-
5, 5-HT-6, and 5-HT-7) are classified by the gene superfamily corresponding to:
(a) The superfamily of receptors associated with G protein, including 5-HT1A,
5-HT1D, 5-HT1E, and 5-HT1F (all inhibit adenylate cyclase), the 5-HT2
receptor subfamily (5-HT2A, 5-HT2B, and 5-HT2C subtypes, all stimulate
the synthesis of phosphoinositides and phospholipase C), and 5-HT-4, 5-HT-­
5, 5-HT-6, and 5HT-7 receptors (which stimulate adenylate cyclase). These
receptors, especially 5-HT-1 (5-HT1D), are linked to the normal and patho-
logical regulation of cerebral flow. 5-HT1A agonists have anxiolytic action
(b) The superfamily of receptors that directly control channels, corresponding
to the 5-HT 3 receptor.
4. Cholinergic system: there is a dense cholinergic innervation of the thalamus, stria-
tum, limbic structures, and cerebral cortex (Fig. 4.9). Except for the striatum, which
is mainly intrinsic (from local interneurons), in most other regions cholinergic inner-
vation is extrinsic. The medial–septal area (with the diagonal band of Broca) is the
origin of cholinergic innervation of the hippocampus (septohippocampal neurons).
The nucleus basalis of Meynert is the origin of the cholinergic innervation of the
neocortex and amygdala (Fig. 4.9). The group of cholinergic neurons of the

Caudate Putamen
nucleus

Neocortex
Hypothalamus
Cerebellum
Olfactory
bulb Raphe
nuclei
Amygdala
Hippocampus

Spinal cord

Fig. 4.8  Principal serotonergic projections of the reticular formation. Modified with permission
from Cardinali [3]
124 4  Second Level: The Brainstem

Parietal cortex

Frontal cortex

Occipital
cortex

Striatum Thalamus

Cerebellum

Medioseptal
nuclei

Meynert’s nucleus Brain


stem Laterodorsal
Amygdala &
Temporal Hippocampus tegmental &
lobe peduculopontine
nuclei

Fig. 4.9  Cholinergic projections in the human brain. Modified with permission from Cardinali [3]

pontomesencephalic reticular formation (PPT/LDT) are the source of cholinergic


thalamic innervation (Fig. 4.9). The neurons of these nuclei participate in the waking
reaction, and in the onset of REM sleep (they activate REM-on cells, Chap. 2). The
alteration of the aforementioned cholinergic systems coexists with the alterations of
the memory and learning, such as those observed in Alzheimer’s disease (AD).
5. Histaminergic system: although it has been linked to this system with monoami-
nergic neurons “in a spider web,” its location is not mesencephalic, but dience-
phalic, in the tuberomammillary nucleus (TMN) of the posterior hypothalamus
(Fig. 4.10). Its receptors are a class of G protein-coupled receptors comprising
four groups: H1 (associated with the circadian cycle, itching, systemic vasodila-
tation, and bronchoconstriction; H2 (tachycardia, stimulation of gastric acid
secretion, smooth muscle relaxation, inhibition of antibody synthesis, T cell pro-
liferation, and cytokine production); H3 (presynaptic autoreceptor, decreasing
ACh, 5-HT and NE in CNS); H4 (which mediates mast cell chemotaxis).

As already mentioned, in most cases, these systems do not participate in point-­to-­


point communication, but rather their diffuse distribution points to a general permissive
role for other brain processes. This anatomical arrangement offers an ideal anatomical
substrate for diffuse functions of alertness, emotionality, and neurovegetative control.
In addition to the role discussed in Chap. 2 in the regulation of sleep/wake rhythm, the
functions of reticular formation include:
24-h Rhythms in Cardiovascular Control 125

Cerebral cortex
a b Lateral ventricles

Thalamus
Fornix

Third
Cerebellum ventricle Fornix
Tuberomammillary
nucleus

To brain stem
Hypothalamus
and spinal cord
Tuberomammillary
nucleus

Fig. 4.10  Histaminergic projections originating in the tuberomammillary nucleus (TMN). (a)
Distribution in a sagittal view. (b) Hypothalamic location of the TMN. Modified with permission
from Cardinali [3]

• Activation of the CNS to trigger specific behaviors and control of alertness.


• Modulation of medullary reflexes and muscle tone, through the reticulospinal tracts.
• Modulation of locomotion (descending pathways of the LC).
• Modulation of the ventromedial and dorsolateral spinal motor neurons (LC, DRN).
• Participation in respiratory and cardiovascular control.
• Participation in the control of different visceral responses, such as urination,
vomiting or defecation.
• Modulation of the flow of nociceptive information in the posterior horn of the
medulla, through reticulospinal projections, mainly serotonergic.

Rather than an apparent reticular structure with little order, the monoaminergic
areas of the brainstem represent a modular somatotopic organization like other brain
areas, such as the neocortex. For example, the periaqueductal serotonergic DRN
neurons have columns composed of afferent, output neurons and interneuronal cir-
cuits organized viscerotopically. Thus, important functions linked to these areas,
such as defensive reactions, analgesia, and autonomic regulation, are integrated in
overlapping longitudinal columns, and different types of aversive or painful stimuli
trigger specific somatic, vegetative, and nociceptive programs.

24-h Rhythms in Cardiovascular Control

A schematic representation of the extrinsic and intrinsic innervation of the heart is


depicted in Fig. 4.11. The neural regulation of the cardiovascular system, controlled
by arterial mechanoreceptors, operates as a reflex arc (Fig. 4.12). Elevations or falls
in BP cause a proportionally greater or lesser deformation of the arterial walls,
which are encoded by baroceptors to a greater or lesser frequency of action potential
firing. This information is carried to the CNS by afferent nerve fibers: (a) the aortic
126 4  Second Level: The Brainstem

Upper
centers

Dorsal Medulla
ganglion
neurons

Spinal cord neurons

Cardiac extrinsic
ganglia

Efferent
Local
sympathetic
neurons
neurons

Cardiac intrinsic ganglia

Local
circuit
Efferent
neurons
parasympathetic
neurons

Afferent
neurons

HEART

Fig. 4.11  Schematic representation of the extrinsic and intrinsic innervation of the heart

depressor nerve, which travels along the X cranial nerve pair; (b) the sinus nerve
(Hering’s nerve), which travels along the IX cranial nerve pair [5].
Primary afferent axons from the baroreceptors project to the caudal region of the
NTS, where they synapse onto second-order neurons, which in turn send excitatory
glutamatergic projections onto inhibitory neurons within the region of the caudal
ventrolateral medulla. These inhibitory neurons synapse directly onto excitatory
neurons within the rostral ventrolateral medulla and serve to inhibit the spontaneous
24-h Rhythms in Cardiovascular Control 127

∆ BP NTS AVP
ADN oxytocin
Mechano - AN cVLM
HN
receptors DMN rVLM

NE/E
Angiotensin
renin

Vagus n. SNS Aldosterone

ANP
HR SV VC TPR
Volemia
CO VR

∆ BP

Fig. 4.12  Schematic representation of neurohumoral responses triggered by the stimulation of vascu-
lar mechanoreceptors by changes in blood pressure (BP) at the three body configurations in a 24-h
cycle. ADN aortic depressor nerve, HN Hering’s nerve, rVLM rostral ventrolateral medulla, cVLM
caudal ventrolateral medulla, AN ambiguus nucleus. DMN dorsomedial nucleus of vagus, SNS sympa-
thetic nervous system, ANP atrial natriuretic peptide, HR heart rate, SV systolic volume, VC venous
capacitance, TPR total peripheral resistance, CO cardiac output, VR venous return. The figure was
prepared in part using image vectors from Servier Medical Art (www.servier.com), licensed under the
Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

activity of premotor sympathetic neurons (Fig. 4.12). Glutamatergic neurons from


the NTS also project to the X cranial nerve pair nuclei that in turn project to the
heart via the X cranial nerve pair. The vagal output originates from preganglionic
neurons located in the dorsal motor nucleus of the vagus and in the ventrolateral
portion of the ambiguous nucleus (AN) in the medulla.
Studies in awake human subjects, in which fMRI was coupled with microelec-
trode recordings of muscle sympathetic nerve activity, has confirmed the operation
of these medullary nuclei in humans [6]. Specifically, information on changes in the
rostral ventrolateral medulla indicated that it is implicated in elevated sympathetic
outflow associated with several cardiovascular diseases including hypertension and
heart failure related to physical inactivity [7].
128 4  Second Level: The Brainstem

In the heart, the vagal postganglionic fibers cause bradycardia, when the initial
stimulus is an increase in BP, and tachycardia, when the stimulus is a decrease in BP
(Fig. 4.12). Other glutamatergic neurons in the NTS project to the caudal ventrolat-
eral medulla, exciting it or not, depending on the increase or decrease in BP.
GABAergic neurons at the caudal ventrolateral medulla project to the rostral ventro-
lateral area, inhibiting it (if the initial stimulus is a BP elevation) or not (when BP
decreases). The rostral ventrolateral area is the site of sympathetic premotor neu-
rons that project to the intermediolateral column of the spinal cord, the site of pre-
ganglionic neurons to the cardiac intrinsic and extrinsic ganglia (Fig. 4.11), and to
the sympathetic peripheral ganglia innervating resistance and capacitance vessels.
This sympathetic activity is thus inhibited during increases in BP or not inhibited
during transitional decreases in BP [8].
Thus, when the triggering reflex stimulus is an increase in BP, there is, reflex-
ively, a vagal activation and inhibition of the sympathetic output, with a consequent
reduction in heart rate, systolic volume, and total peripheral resistance, and an
increase in venous capacitance. This decreases the venous return to the heart, deter-
mining a decrease in BP and its return to baseline (Fig. 4.12). On the other hand,
when the triggering stimulus is a decrease in BP, the vagus nerve is not activated and
the sympathetic input is not inhibited. This causes an increase in heart rate, systolic
volume, and total peripheral resistance, and a decrease in venous capacitance.
Consequently, there is an increase in the venous return, which helps to increase
cardiac output further (Fig. 4.12). These responses, triggered neurally, are extremely
fast, correcting, in seconds, BP swings up or down the control values.
Orthostatic pooling of blood begins almost immediately upon the change from the
supine to the upright posture. The main sensory receptors involved in orthostatic
cardiovascular reflex adjustment are the arterial baroreceptors located in the carotid
sinuses and aortic arch and mechanoreceptors located in the heart and lungs [9]. A
decrease in BP, as occurs on assumption of the upright posture, removes this tonic
inhibition with a resultant decrease in vagal outflow and an increase in sympathetic
activity causing an increase in heart rate, cardiac contractility and vasomotor tone.
Central modulation of vasomotor outflow is reinforced by local vasoconstrictor
mechanisms, such as the veno-arteriolar axon reflex and a myogenic response. The
veno-arteriolar axon reflex is triggered when venous pressure exceeds 25 mmHg,
which results in vasoconstriction of the corresponding arteriole and is reported to
elicit up to 30–45% of the total vasoconstriction in the legs in the upright posture [9].
The sympathetic and parasympathetic components of the ANS play a crucial role
in maintaining cardiovascular homeostasis and enabling the body to respond to phys-
iological stressors. Neurogenic mechanisms are not only essential for maintaining
and regulating arterial BP, but also play a crucial role in regulating the distribution of
blood flow between and within vascular beds. The sympathetic component of the
ANS plays the predominant role in regulating vascular tone and whole-body hemo-
dynamics via its effects on both resistance and capacitance vessels. By contrast, the
overall contribution of the parasympathetic nervous system to the regulation of vas-
cular tone and hemodynamics is small compared with its primary regulatory role in
mediating negative chronotropic and inotropic effects on the heart [10].
24-h Rhythms in Cardiovascular Control 129

The vascular endothelium plays an essential role in the regulation of blood vessel
tone and cellular activity, helping to maintain a healthy vessel [11]. Endothelial cells
produce several important vasoactive substances including NO, prostacyclin, endo-
thelium-derived hyperpolarizing factor, endothelin, vasoactive prostanoids, and ROS.
These factors, and other endothelium-derived substances, also modulate local throm-
botic and inflammatory pathways influencing the progression of atherosclerosis and
its complications. Exposure to risk factors for atherosclerosis and the presence of
circulatory diseases, including atherosclerosis and heart failure, leads to endothelial
activation, associated with reduced NO bioavailability and expression of proinflam-
matory cytokines, chemokines, selectins, and adhesion molecules. This enhances
vasoconstrictor tone and promotes a proatherogenic milieu.
Several hormonal mechanisms are also involved in BP control, including the
release of catecholamines, angiotensin II, aldosterone, AVP, oxytocin, and atrial
natriuretic peptide (ANP), which act by supporting the maintenance of baseline BP,
intensifying and prolonging the cardiovascular responses for minutes or even hours,
making BP control more effective, especially in situations of prolonged elevations
or falls of BP (hemorrhage, dehydration, drug reactions, etc.; Fig. 4.12). The actions
of signaling molecules that are not classically viewed as such, e.g., cytokines and
ROS, must also be considered [12].
Circadian clock genes are expressed in the heart and aorta and these genes and
approximately 4–6% of the cardiac protein genes showed circadian rhythms in tran-
scription [13]. Ex vivo experiments demonstrate that varied functions of the heart
and aorta are dependent on the time in which tissues are collected. In murine knock-­
out models of circadian genes, suppression of Bmal1 in cardiomyocytes results in
an abnormal electrocardiogram (ECG) with RR and prolonged QRS intervals. The
hearts of Bmal1 knockout mice were more susceptible to arrhythmia. Other studies
have revealed that removal of Bmal1 in endothelial cells or vascular smooth muscle
cells alters the diurnal variation of BP. These findings are consistent with the pres-
ence and importance of circadian genes in the cardiovascular system [13, 14].
The alteration of normal day–night cycles, such as jet lag or shift work, leads to the
desynchronization between the central and peripheral clocks and the deregulation of the
clock genes (Chap. 8). Restoration of a normal daytime rhythm rescues from these
changes, suggesting that maintaining a normal rhythm is crucial for cardiovascular health.
Cardiovascular function changes significantly in the three body configurations
during the 24-h cycle [15]. BP decreases during NREM sleep and becomes variable
in REM sleep. During REM sleep, transient BP increases of up to 40 mmHg occur
that coincide with the phasic events of this stage of sleep in conjunction with vaso-
constriction in the skeletal muscles. Pulmonary artery pressure remains stable.
Variation of sleep-related BP can be described by a square wave function with
changes in the onset and end of sleep and relatively constant values during sleep.
This fall in BP during sleep is important for cardiovascular health [15].
The values of systolic pressure drop 15 mmHg or more during sleep and are heav-
ily influenced by the S Process described in Chap. 2. That is, the BP drop accompanies
the presence of sleep, regardless of the time of day at which it occurs. An initial fall in
BP due to postural change and darkness (~ 7 mmHg) is followed by a period of
130 4  Second Level: The Brainstem

instability when sleep is unstable (stage N1 of sleep) and by an abrupt drop once sta-
ble sleep is achieved (N2–N3 stages of sleep; ~ 7 mmHg). Within each sleep phase the
BP is constant, in the NREM sleep, the values are lower than in wakefulness, and in
REM sleep they are similar to those of relaxed wakefulness. Although there are tran-
sient increases in association with phasic events of REM sleep, the major disturbance
of BP during sleep is in the awakening. BP at the end of sleep shows a rise largely due
to postural changes. The magnitude of the changes is greater when awakening occurs
at the N2 sleep stage compared with awakening in the REM stage [15].
In the case of the heart rate the closest approximation is to a sinusoidal pattern
related to the central temperature and compatible with the influence of the C process
(circadian), i.e., the changes occur regardless of sleep. The lowest point of the oscil-
lation is in the middle of the night and remains, albeit attenuated, in individuals who
are deprived of sleep. As for BP, at the beginning of sleep, there is a fall in the heart
rate with two components (preparation for sleep; when sleep becomes stable). There
is a close relationship between heart rate and metabolic heat production because of
its circadian dependence. Heart rate is higher during the REM sleep phase, with
transient tachycardia in relation to REM phasic events (Fig. 4.13).

WAKEFULNESS NREM SLEEP REM SLEEP

MCC MCC MCC


INS INS INS
AMYG AMYG AMYG

BS BS BS

a a a
b b b
c c c

HR +++ + ++

SNS +++ + ++
PNS + +++ +

Fig. 4.13  Modulation of cardiac activity at the three body configurations in a 24-h cycle. Letters
designate reflex loops (a: baroreflex; b: chemoreflex; c: respiration). The brain stem centers (BS) and
central autonomic network including midcingulate cortex (MCC), insula (INS), amygdala (AMYG)
are depicted. Relative changes in the heart rate (HR), sympathetic nervous system. In red increases and
in blue decreases in activity are shown. Redrawn from Chouchou and Desseilles [16]. The figure was
prepared in part using image vectors from Servier Medical Art (www.servier.com), licensed under the
Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)
24-h Rhythms in Cardiovascular Control 131

Heart rate variability (HRV) analysis combined with brain imaging has identified
a close connectivity between autonomic cardiac modulation and activity in brain
areas such as the amygdala and insular cortex during REM sleep, but no connectivity
between the brain and cardiac activity during NREM sleep [16]. There is also evi-
dence for an association between HRV and the intensity and emotionality of dreams.
Brief awakenings are a characteristic of normal sleep. They occur with high fre-
quency (more than 15–20 per night) and are a normal situation in which an increase
in heart rate (≥8 beats per minute), BP (<15 mmHg) and peripheral vasoconstriction
are observed. Awakenings occur in both NREM sleep and REM sleep. There are two
components of awakenings. The first component is a transient peak of heart rate and
BP that occurs within 3–6 s of the event. The second component is dependent on the
previous wake period. Most awakenings are brief and there is no such second compo-
nent. For many normal individuals, the cardiovascular activation response occurs at a
level of excitation below the occurrence of α changes in the EEG (cortical awaken-
ing). There are large individual differences in cardiovascular response to awakenings
and interindividual variability of systolic BP response in an awakening can reach
about 15 mmHg. Healthy individuals with large activating cardiovascular responses
are at increased risk for cardiovascular disease because of frequent awakenings [17].
In NREM sleep, there is a period of relative autonomic stability with vagal pre-
dominance and increased gain of baroreceptors. A sinusoidal modulation of heart
rate exists because of the coupling of cardiovascular and respiratory regulation cen-
ters, resulting in respiratory sinus arrhythmia. During inspiration, the frequency
briefly accelerates to accommodate the increased venous return, resulting in
increased cardiac output, whereas during expiration, a progressive decrease in heart
rate occurs. This normal sinus heart rate variability, particularly during NREM
sleep, is generally indicative of a state of heart health; thus, the absence of sinus
arrhythmia has been associated with heart disease and advanced age [18].
Cardiac- and respiratory-related rhythms are observed in nerve activity supply-
ing the heart and blood vessels [19]. These rhythms arise from phasic inputs related
to cardiac/pulse or ventilation-related afferent activity and/or a common cardiore-
spiratory CNS network. Much of this heart rate fluctuation is linked to the phase of
respiration (i.e., respiratory sinus arrhythmia) and its loss is a prognostic indicator
of morbidity and mortality. Another possible origin of sympathetic rhythms is inde-
pendent CNS oscillators. Separate oscillators, which are able to couple, may drive
activity to different sympathetic nerves, and sympathetic neurons regulating the
same target may be influenced by populations of weakly coupled or uncoupled
oscillators [19]. Two major hypotheses have been proposed to account for cardiac-
and respiratory-related rhythms in sympathetic discharges. The classic view holds
that these rhythms are imposed upon sympathetic discharge by “external” inputs.
The observation of a nonrespiratory and noncardiac-related sympathetic rhythm
suggests that sympathetic rhythms might not arise exclusively from phasic inputs to
tonic sympathetic tone generating networks [19]. Loss of vagal tone in cardiovascu-
lar diseases can be demonstrated by the diminished change in heart rate on admin-
istration of a vagolytic drug such as atropine and by the loss of respiratory sinus
arrhythmia. The burst of cardiac vagal activity originates centrally at the level of the
preganglionic neurons in the AN, which are inhibited during inspiration, but excited
132 4  Second Level: The Brainstem

during postinspiration. The changes in BP during respiration are inversely related to


the respiratory effort increases in BP, resulting in a decrease in respiratory rate. This
effect is potentiated during NREM sleep in which small reductions in BP lead to
increases in respiratory rate. These pauses and increases of respiratory rate serve as
compensatory mechanisms to normalize BP and their disappearance in infants pre-
disposes to a sudden death.
Sympathetic nerve activity is relatively stable during NREM sleep, and cardio-
vascular activity is reduced to more than half of that observed in the vigil at the N3
NREM sleep stage. NREM sleep, with relative hypotension, bradycardia, and
reduced cardiac output and systemic vascular resistance, provides a “daily vaca-
tion” for heart activity. Bradycardia is mainly of a vagal origin, whereas hypoten-
sion is mainly attributable to a reduction in sympathetic vasomotor tone.
In the transition from NREM to REM sleep, abrupt vagal discharges can lead to
pauses in cardiac rhythm or to asystolia. As discussed in Chap. 2, REM sleep occurs
at 90-min intervals and several essential homeostatic mechanisms are disrupted dur-
ing this period. The increase in limbic activity during REM sleep leads to significant
sudden increases in cardiac sympathetic nerve activity at the level of the coronary
vessels (Fig. 4.13). The baroreceptor gain is reduced and the heart rate fluctuates
markedly, with marked episodes of tachycardia and bradycardia. Vagal tone is gen-
erally suppressed during REM sleep with very irregular breathing patterns that can
lead to hypoxemia, particularly in patients with pulmonary or cardiac disease.
Except for the diaphragm and cricopharyngeal sphincter there is hypotonia of the
muscles of the airways. During sleep apnea, respiratory activity can be stopped for
central or peripheral reasons several times each night, with adverse consequences
for normal cardiorespiratory activity.
Changes in coronary blood flow occur at sleep states and during transitions
between sleep states. In monkeys, nighttime coronary flow increases sporadically
up to 100% during sleep. These periodic oscillations in blood flow are not associ-
ated with alterations in cardiac work or BP. Concomitant with sudden increases in
heart rate during REM sleep, abrupt increases in coronary pressure and correspond-
ing decreases in coronary vascular resistance were observed. These phenomena are
seen predominantly during REM sleep in phase with POG events in EEG and ocular
movements. There were no significant changes in mean BP. As the heart rate is
elevated during sudden increases in coronary flow, increased cardiac metabolic
activity was assumed to be the cause of coronary vasodilation. These sudden
increases in coronary blood flow appear to be the result of increased adrenergic
discharge as they were suppressed by removal of the corresponding sympathetic
ganglia.
During severe stenosis of the coronary artery a decrease in coronary arterial
blood flow is observed during REM sleep instead of the increases found in normal
subjects. In these patients, there is a strong correlation between the magnitude of
the increase in heart rate and the decrease in coronary flow. The link between
REM sleep and the occurrence of myocardial ischemia in coronary patients is
consistent.
24-h Rhythms in Respiratory Control 133

24-h Rhythms in Respiratory Control

The spontaneous rhythm alternating between inspiration and expiration is produced


by the automatic generation of a basal pattern in the brainstem. This spontaneous
and stereotyped rhythm is modified by metabolic changes such as changes in pH or
partial gas concentrations in the blood, or by mechanical changes (e.g., postural).
Respiratory rate and tidal volume are modifiable to allow modifications of the inter-
nal environment. Interruptions of breathing (apnea) can be generated to allow pho-
nation or swallowing.
The regulatory system of breathing includes: (a) a control center located in the
CNS at the level of the brainstem, where the neural activity that triggers breathing
starts; (b) effector muscles that produce ventilation (the respiratory muscles, espe-
cially the diaphragm); (c) a set of sensors located in the lungs and central and
peripheral chemoreceptors that regulate respiratory activity [20].
The respiratory centers are divided into four major groups, two groups in the
medulla and two in the pons (Fig. 4.14). The two groups in the medulla are the dor-
sal respiratory group and the ventral respiratory group. The two groups in the pons
are the pneumotaxic center also known as the pontine respiratory group, and the
apneustic center. The exact location of the apneustic center is not yet defined. The
inspiratory center (dorsal respiratory group) is in the dorsal portion of medulla, in
the NTS.
The expiratory center (ventral respiratory group) is in the AN, in the anterolateral
part of the medulla (Fig. 4.14). The AN consists of a rostrally to caudally extending
column of neurons expressing respiratory-related activity, with subregions contain-
ing motoneurons that innervate the muscles of the larynx and pharynx that are not
considered part of the ventral respiratory group. It generally causes expiration, but
can cause either expiration or inspiration depending upon which neuron in the group
is stimulated.
The expiratory center sends an inhibitory impulse to the apneustic center, which
is presumably located in the lower part of the pons and which releases stimulatory
impulses to the inspiratory center causing inspiration. It receives an inhibitory
impulse from the pneumotaxic center and from stretch receptors of the lung via the
X cranial nerve pair and discharges inhibitory impulse to the expiratory center. The
pneumotaxic center, located in the upper part of the pons, in the PBN, controls both
rate and pattern of breathing and limits inspiration (Fig. 4.14).
Both the dorsal respiratory group and the other sub-nuclei of the NTS are the
primary sites for vagal afferent projection of the lungs and afferents and chemore-
ceptors of the carotid and aortic baroreceptors. The NTS, including the dorsal respi-
ratory group, is a key site integrating sensory information from the lung, in addition
to information on the levels of arterial PCO2, PO2, and pH.
The ventral respiratory group and dorsal respiratory group contain both bulbos-
pinal respiratory premotoneurons (i.e., neurons that project to spinal motoneurons,
which in turn innervate the respective respiratory pump and abdominal muscles of
breathing), and propriobulbar neurons (i.e., neurons that project to, and influence
134 4  Second Level: The Brainstem

Facial Pneumotaxic
nucleus center

Bötzinger
complex

Pre-Bötzinger
complex
VRG DRG
Rostral VRG

Caudal VRG

Nucleus ambiguus

Phrenic
nucleus

Fig. 4.14  The respiratory centers are divided into four major groups. The two groups in the
medulla are the dorsal respiratory group (DRG) and the ventral respiratory group (VRG). The two
groups in the pons are the pneumotaxic center and the apneustic center (whose exact location is not
yet defined). The inspiratory center (DRG) is in the dorsal portion of the medulla, in the NTS. The
expiratory center (VRG) is in the ambiguous nucleus. For other details, see the text

the activity of, other medullary respiratory neurons, but do not themselves project to
motoneurons). The hypoglossal, trigeminal, and facial motor nuclei also innervate
muscles important to pharyngeal motor control and the maintenance of upper air-
way patency.
In mammals, the Bötzinger complex is a group of neurons located in the rostral
ventrolateral medulla and ventral respiratory column. In the medulla, this group is
located caudally to the facial nucleus and ventrally to the AN. The Bötzinger com-
plex plays an important role in controlling breathing and responding to hypoxia. It
consists primarily of glycinergic neurons, which inhibit respiratory activity.
24-h Rhythms in Respiratory Control 135

The Bötzinger complex has projections to the NTS, phrenic pre-motor neurons
in the medulla, phrenic motor neurons in the cervical spinal cord, the dorsal respira-
tory group, and the ventral respiratory group (Fig. 4.14). Bötzinger complex neu-
rons are intrinsic pacemakers that are important to the generation of the basic
respiratory rhythm and the expression of rhythmic neuronal activity elsewhere in
the respiratory network. Respiratory rhythm-generating pre-Bötzinger complex
neurons coexpress μ-opioid and neurokinin-1 receptors (i.e., the receptors for sub-
stance P) that slow and increase the respiratory rate respectively. The presence of
μ-opioid receptors in pre-Bötzinger complex neurons explains the respiratory rate
depression that follows the administration of opioid drugs. During inspiration, the
central respiratory drive potential is transmitted to phrenic and intercostal motoneu-
rons via monosynaptic connections from inspiratory premotor neurons of the dorsal
respiratory group.
Central sensors that detect changes in CO2 are located on the ventral surface near
the entrance of the VIII and XI cranial nerve pairs. These chemoreceptors respond
to the local application of CO2 or acids and are inhibited by anesthetics and local
cold. These chemoreceptors are not in direct contact with blood, but are bathed in
CSF and respond to changes in both arterial PCO2 and CSF pH.
Carbon dioxide diffuses easily through the blood–brain barrier, but H+ does not.
Thus, the stimulus produced by increased ventilation is the increase in PCO2, which,
after crossing the blood–brain barrier, causes a fall in pH in brain tissue. Alterations
of PaCO2 are rapidly transmitted to the CSF, which has little CO3H2 as a buffer.
After an acute change in arterial PaCO2, there is an even greater change in PCO2 in
the CSF. The response time constant is about 60 s.
Peripheral sensors are located in the carotid and aortic bodies [21]. They measure
PO2, PCO2, and arterial pH. They are sensitive to the decrease in PaO2 that is measured
directly and induces hyperventilation. Denervation of peripheral chemoreceptors
leaves hypoxia without its fundamental homeostatic regulatory mechanism and because
of hypoxia, CNS depression occurs. Increases in PaCO2 and the decrease in arterial
blood pH stimulate these receptors less, but amplify their response to hypoxemia.
The carotid body, a highly vascular tissue, receives afferent innervation from the
carotid sinus nerve, which is a branch of the glossopharyngeal nerve. The increased
sensory activity of the carotid body is maintained during the entire period of hypoxia
with little adaptation. Thus, the exquisite sensitivity and the rapid response to a
wide range of hypoxic intensities with little or no adaptation make the carotid body
a unique oxygen-sensing organ in comparison with other tissues [22].
Autonomic nerves play an important role in regulating the functions of the air-
ways, including airway smooth muscle tone, mucus secretion, and blood flow.
Afferent nerves in the airway are important with regard to airway defenses (cough),
inducing reflex effects, and through the release of neuropeptides (neurogenic inflam-
mation). Cholinergic nerves are the major bronchoconstrictor pathway through the
activation of muscarinic receptors on airway smooth muscle. By contrast, adrener-
gic nerves have little direct control of airway smooth muscle, circulating E being
more important in adrenergic regulation. A neural bronchodilator pathway is medi-
ated by release of NO. Several neuropeptides are expressed in airway nerves and
136 4  Second Level: The Brainstem

play a co-transmitter role in concert with the classical autonomic transmitters [23].
Autonomic nerve function is regulated primarily through reflexes initiated upon
bronchopulmonary vagal afferent nerves [24].
During NREM sleep, there is a decreased chemosensitivity and apneas at the
onset of the sleep phase [25]. During REM sleep, the decrease and variability of
chemosensitivity, with a greater propensity to apneas and the decrease in the neural
discharge to the respiratory muscles (except the diaphragm), are accentuated.
During wakefulness, the respiratory control is exerted by three mechanisms:

• Metabolic mechanisms, which ensure arterial O2 and CO2 homeostasis by central


and peripheral chemoreceptors.
• Voluntary control, which allows the adaptation of ventilation to phonation and
other demands.
• Tonic depolarization of the spinal motor neurons of the respiratory muscles.

Respiratory changes during sleep reflect the inhibition of some of these control
mechanisms (Fig. 4.15). Metabolic control (index of a predominant parasympathetic

AWAKE SLOW WAVE SLEEP REM SLEEP

Eupnea / hyperpnea Hypopnea / Apnea Hyperpnea / Apnea

Respiratory
pattern

Respiratory
Voluntary and oscillator
Obstruction
metabolic control desynchronization

Central
component
Strong tonic Reduced tonic Augmented tonic and
respiratory drive respiratory drive phasic respiratory drive

Normal Decreased Normal Collapsed


High muscle Absent
airway muscle tone airway airway
tone muscle tone

Peripheral
component

Forced expiration

Fig. 4.15  Balance between the different components of respiratory control in the three body con-
figurations in a 24-h cycle. When the voluntary control disappears and the peripheral component
is reduced, there is a tendency toward hypopnea or apnea in sleep (slow-wave sleep). This is exac-
erbated in REM sleep because of the motor and autonomic disconnection. Hence, the apnea pre-
dominates in the latter part of the night. Modified with permission from Cardinali [3]
24-h Rhythms in Respiratory Control 137

mechanism) prevails in slow-wave sleep and there is a decrease in this control during
REM sleep. The predominance of REM sleep in the last stages of the night explains
the higher incidence of episodes of apnea in the second part of the night [25].
The initiation of sleep and up to stage N2 of slow-wave sleep is accompanied by
an unstable respiratory rhythm with successive episodes of hypo- and hyperventila-
tion called “periodic ventilation.” During stage N3, ventilation becomes regular in
terms of respiratory rate and amplitude. Respiratory rate and depth are relatively
constant, this being a stable period from the respiratory point of view.
Respiratory rhythm during REM sleep is characterized by being faster and
mostly irregular, with apneic episodes and hypoventilation. The responsible mecha-
nism is central, neural, to which is added the muscular hypotonia, which has a dou-
ble influence: on the one hand, it diminishes the force of the expansion of the rib
cage, and on the other hand it increases the resistance of the superior airway to the
passage of the air. The diaphragm maintains irregular activity, but does not partici-
pate in the generalized atonia of REM sleep, because it lacks a significant number
of muscle spindles.
Cardiorespiratory homeostasis involves the regulation of two motor systems, one
that supplies the somatic (i.e., diaphragmatic, intercostal, abdominal, and upper air-
way) musculature and another the autonomic innervation of heart and vasculature.
As seen, the activity of respiratory neurons varies greatly in the stages of sleep, as
does the regularity of the heart rhythm. During REM sleep, there is tachycardia,
polypnea, sweating, and dramatic elevations in arterial BP secondary to the intense
autonomic activity that occurs in this period. Maintaining the perfusion of vital
organs through the control of adequate BP is essential for homeostasis. Respiratory
mechanisms are recruited to support cardiovascular action, helping the venous
return and altering the heart rate by reflex.
The integration of the cardiorespiratory function during sleep requires the par-
ticipation, in addition to the brainstem, of central areas of the autonomic hierarchy.
This has been documented by imaging studies using PET. In REM sleep, the pref-
erential activation of limbic and paralimbic regions of the anterior brain is demon-
strated in comparison with wakefulness or NREM sleep. The serotonergic neurons
of the DRN play an important role in vascular control. These neurons are damaged
in heart failure, probably because of altered perfusion and hypoxia accompanying
altered breathing.
The orbitofrontal cortex, parts of the formation of the hippocampus, the hypo-
thalamus, and other structures of the CNS participate in regulation of the cardio-
respiratory pattern [26]. The central nucleus of the amygdala is strategically
positioned to regulate cardiac and respiratory functions in affective behavior, as
it broadly projects to the brainstem (NTS, PBN, the dorsal motor nucleus of the
vagus, periaqueductal gray). Many are involved in mediating the transient rise in
arterial BP caused by the cold or the Valsalva maneuver. These structures are
severely damaged, both in patients with heart failure and in obstructive sleep
apneas.
The insular cortex deserves special attention among the cortical areas that express
the regulatory action on cardiovascular control in the sleep and waking states
138 4  Second Level: The Brainstem

(Fig. 4.13). This area modulates the sympathetic activity (mainly the right insula)
and the parasympathetic activity (mainly the left insula). There is marked damage
of the insula under conditions of respiratory disorders and heart failure.

The Cerebellum and the Autonomic Posture

The cerebellum, the largest subcortical center for motor control, has been demon-
strated in the last decade to be involved in the regulation of nonsomatic functions,
such as respiration, feeding behavior, cognition, and working memory [27–29].
For example, activation of fastigial nuclear neurons predominantly increases ven-
tilation via elevation of the respiratory frequency and/or tidal volume. Ablation of the
fastigial nucleus did not significantly alter eupneic breathing, but did markedly atten-
uate the respiratory response to medium and severe hypercapnia, and hypoxia. The
fastigial nucleus contains respiratory-modulated neurons and about 25% of these
neurons do not show their respiratory-related phasic activity until exposed to hyper-
capnia. The fastigial nucleus also contains CO2/H+-chemosensitive sites that contrib-
uted to the respiratory response to hypercapnia. The involvement of the cerebellum
in the control of cardiovascular and respiratory activity in the three body configura-
tions of a 24-h cycle, i.e., wakefulness, NREM, and REM sleep, is thus significant.
Although it is not yet clear through which pathways such cerebellar nonsomatic
functions are mediated, the direct bidirectional connections between the cerebellum
and the hypothalamus are probably involved [30]. The direct hypothalamocerebel-
lar projections originate from the widespread hypothalamic nuclei/areas and termi-
nate in both the cerebellar cortex as multilayered fibers and the cerebellar nuclei.
Immunohistochemistry studies have indicated that some of these projecting fibers
are histaminergic. It has been suggested that through their excitatory effects on cer-
ebellar cortical and nuclear cells mediated by metabotropic histamine H2 and/or H1
receptors, the hypothalamocerebellar histaminergic fibers participate in the cerebel-
lar modulation of somatic motor and nonmotor responses. The histaminergic affer-
ent system of the cerebellum, having been considered an essential component of the
direct hypothalamocerebellar circuits, originates from the tuberomammillary
nucleus in the hypothalamus [31].
The direct cerebellar–hypothalamic projections arise from all the three cerebellar
nuclei, the fastigial nucleus, the interpositus nucleus, and the dentate nucleus
(Fig. 4.16), and terminate to extensive regions of the hypothalamus, such as the lateral
hypothalamic area (LHA) and the posterior and dorsal hypothalamic areas, in addition
to the dorsal medial nucleus (DMN) and the paraventricular nucleus (PVN; Fig. 4.17)
[32]. Neurophysiological and neuroimaging studies have demonstrated that these con-
nections are involved in feeding, cardiovascular, osmotic, respiratory, micturition,
immune, emotion, and other nonsomatic regulation. For example, electrophysiologi-
cal data suggest that via the direct cerebellohypothalamic projections, the cerebellar
outputs may reach, converge, and be integrated with some critical feeding signals,
including gastric vagal afferents, CCK, leptin, and glycemia on single hypothalamic
neurons [33]. Hypothalamic orexin neuronal projections to the cerebellum
The Cerebellum and the Autonomic Posture 139

Fig. 4.16 Cerebellar Vermis


nuclei. Modified with
permission from
Cardinali [3]

Dentate
nucleus

Globose
Emboliform
Fastigial nucleus nucleus
nucleus

MOTOR PROJECTIONS AUTONOMIC PROJECTIONS

Medial motor Spinocerebellum


brain stem pathways

Lateral motor
brain stem pathways LHA

Posterior
& dorsal
hypo-
Premotor & thalamus
motor cortex
VMA

Cerebro- DMN
cerebellum
PVN

Vestibular
nuclei
Vestibulocerebellum

Fig. 4.17  Motor and autonomic projections of the cerebellum. Modified with permission from
Cardinali [3]

co-ordinate vestibulo-cerebellar motor and autonomic functions associated with feed-


ing behavior [34]. Furthermore, functional imaging studies provide substantial evi-
dence that hunger, satiation, and thirst are accompanied by a cerebellar activation.
Concerning the immune function, lesions of fastigial nuclei enhance cellular and
humoral immunity, whereas lesions of the interpositus nucleus inhibit the immune
function. The positive and negative immunoregulations by the two different cerebellar
nuclei strongly show that the cerebellum involves an elaborate and critical modulation
of immune homeostasis [35]. Collectively, these observations provide support for the
hypothesis that the cerebellum is an essential modulator and coordinator for integrat-
ing motor, visceral, and behavioral responses participating in the autonomic posture.
140 4  Second Level: The Brainstem

In rodents, the reversible inactivation of the vermis during the consolidation or


the reconsolidation period hampers the retention of the fear memory trace. This is a
typical example of cerebellar participation in emotional learning conclusively dem-
onstrated for motor control. Imaging experiments show that in humans the cerebel-
lum is also activated during mental recall of emotional personal episodes and during
learning of a conditioned or unconditioned association involving emotions
(Fig.  4.18). The vermis participates in fear learning and memory mechanisms
related to the expression of autonomic and motor responses of emotions. In humans,
the cerebellar hemispheres are also involved at a higher emotional level.
To understand the cerebellar role in autonomic function it is necessary to briefly
review the cerebellum function as derived from classical motor control experiments.
Cerebellar neurons are distributed in the cortex, and in three pairs of nuclei located
in the interior of the cerebellar hemispheres: (a) the fastigial nucleus; (b) the inter-
positus nucleus (globular plus emboliform nuclei); (c) the dentate nucleus
(Fig. 4.16). Despite representing 10% of the brain weight, the cerebellum contains
50% of the brain neurons. The structure of the cerebellum is very systematized and
is similar throughout the organ, suggesting a common basic function, modified by
the type of information that each zone receives.
To perform its function of motor coordination, the cerebellum acts as a compara-
tor of intention with the motor activity performed. The basic mechanism of function
of the part of the cerebellum related to motor execution is to supervise the different
executive stages of the motor plan. For this it uses:

(a) Information about the motor plan derived from CNS structures, from the pri-
mary motor cortex to motor neurons. This information is called “internal
feedback.”
(b) (b) Information on the periphery via sensory pathways originating in the skin,
muscles, and joints (called “external feedback”). Correction of the plan is car-
ried out by the cerebellum by projection to the neuronal groups constituting the
descending motor systems.

This comparison of the “plan” with the “execution of the plan” allows the cerebel-
lum to appreciate deviations and to correct them, not by direct action on the motoneu-
rons, but by indirect influence through the descending motor pathways. An equivalent
occurs for several autonomic behaviors. The cerebellum receives information from
the limbic system on the autonomic strategy selected as adequate, and presents similar
phenomena to the internal and external feedback indicated above. The cerebellar data
are consistent with a dampening or coordinating role for the cerebellum in the pres-
ence of significant changes in BP that could be similar to motor coordination.
The function of the cerebellum is modified by experience, hence its importance in
motor and autonomic learning. For this motor function, the cerebellum receives infor-
mation from three sources: (a) the periphery; (b) the brainstem; (c) the cerebral cortex.
The pathways that enter the cerebellum send collaterals to the cerebellar nuclei and
the cerebellar cortex. For its autonomic function, the main afferent is that of the corti-
cal and subcortical areas of the limbic system (Fig. 4.17). The exit of the cerebellar
Auditory cortex Medial geniculate body Inferior colliculus

Anterior insula Pons

Pregenual anterior Cerebellar cortex


cingulate cortex

Anterior mid-
Cerebellar nuclei
cingulate cortex
The Cerebellum and the Autonomic Posture

Supplementary
Motor area

Supramarginal gyrus
Parieto-temporo-
occipital cortex Hypothalamus
Superior temporal
gyrus AUTONOMIC POSTURE

Fig. 4.18  A diagram describing the generation of autonomic responses following a loud tone presentation. Auditory signals influence auditory regions and the
anterior insula, whose activity reflects stimulus saliency. Signal generated in the anterior insula modulate autonomic activity through direct projections to the
hypothalamus and to other cortical areas within the salience network. Signals generated in the auditory cortex also influence cerebellar nuclei through pontine
projections. The cerebellar cortex, in turn, influences autonomic activity through projections to the hypothalamus
141
142 4  Second Level: The Brainstem

cortex always passes through the cerebellar nuclei (except for that corresponding to
the flocculonodular lobe, which is sent to the vestibular nuclei). This particularity
allows the cerebellar nuclei to perform the integration of the input information with
the elaboration of the cerebellar cortex. The cerebellar entrance and exit routes are
projected by the three pairs of cerebellar peduncles: upper, middle, and lower.
Each portion of the cerebellar cortex projects to a given group of cerebellar
nuclei (Fig. 4.16) following the distribution:

• The vermis projects to the fastigial nucleus


• The middle portion of the cerebellar hemisphere projects to the interposed nucleus
• The lateral portion of the cerebellar hemisphere projects to the dentate nucleus
• The flocculonodular lobe projects to the vestibular nuclei

In Fig. 4.17, the three functional divisions of the cerebellum and their motor and
autonomic outputs are schematized: (a) vestibulocerebellum (flocculonodular lobe);
(b) spinocerebellum, composed of the vermis and the intermediate portion of the
cerebellar hemispheres; (c) the cerebrocerebellum, composed of the lateral portion
of the cerebellar hemispheres.
The vestibulocerebellum is the phylogenetically oldest part of the cerebellum. Its
function is the control of body posture and ocular movements. It receives vestibular
information directly from the labyrinth, and through the vestibular nuclei. The exit
is through the vestibular nuclei. Its autonomic participation is important in BP
responses to rapid postural changes.
The spinocerebellum receives information from the spinal cord through the spi-
nocerebellar bundles, and from the auditory, visual, vestibular, and ANS projec-
tions. These projections are organized somatotopically, indicating the comparative
function of the motor plan and execution that fulfills the cerebellum. The projec-
tions that reach the spinocerebellum from the spinal cord are direct (through the
ventral and dorsal spinocerebellar bundles) and indirect (from the mesencephalic
relay nuclei, such as the inferior olive). Lesions of the vermis block the retention of
the fear memory trace (Fig. 4.18) and in humans the cerebellum is activated during
the mental recall of emotional personal episodes and during the learning of a condi-
tioned or unconditioned association involving emotion. Thus, the spinocerebellum
contributes substantially to the autonomic posture (Chap. 3; Figs. 1.5 and 4.18).
The cerebrocerebellum participates in motor planning. It is the center of a com-
plex feedback system, which modulates cortical commands. The output of the cere-
brocerebellum oversees the dentate nucleus, which projects to the lateral ventral
nucleus of the thalamus, and from here to the premotor and motor cortex and to the
LHA, DMN, and PVN areas of the hypothalamus.
The histological structure of the cerebellum is homogeneous, regardless of the func-
tion that the region fulfills. The cerebellar cortex is divided into three layers: molecular,
Purkinje, and granule cells. The fundamental cellular elements present in the cerebel-
lum are the GABAergic Purkinje cells, which are the only neurons to exit the cerebellar
cortex and granule cells, their excitatory interneurons originating from the parallel
fibers (Fig. 4.19). The different afferences that reach the cerebellum are made in one of
two ways: (a) mossy fibers, which constitute the most important contribution and
The Cerebellum and the Autonomic Posture 143

Parallel
fiber

Granule
cell Purkinje cell

Climbing fiber

Mossy
fiber

Output

Cerebellar nucleus

Fig. 4.19  The histological structure of the cerebellum is homogeneous regardless of the function
that the region fulfills. The principal cell elements are: (a) the GABAergic Purkinje cells, which are
the only projection neurons of the cerebellar cortex; (b) granule cells, which are excitatory inter-
neurons that give rise to the parallel fibers. The different afferences that reach the cerebellum are
made in one of two ways: (a) mossy fibers, which constitute the most important contribution and
which comprise all the entrances to the cerebellum except for the inferior olive; (b) climbing fibers,
originating in the lower olive. Both pathways send a stimulating collateral to the cerebellar nuclei
before proceeding to the cerebellar cortex. Modified with permission from Cardinali [3]

which comprise all the entrances to the cerebellum except for the inferior olive; (b)
climbing fibers, originating in the inferior olive. Both entry routes send a stimulatory
collateral to the cerebellar nuclei before proceeding to the cerebellar cortex (Fig. 4.19).
Mossy fibers exert an indirect stimulatory influence on Purkinje cells through the
granule cells, on which they synapse. The granule cells give rise to the parallel fibers,
which activate the dendrites of the Purkinje cells, at the level of the molecular layer.
The arrangement of the parallel fibers in relation to the dendritic trees of the Purkinje
cells resembles that of old telephone wires in their contact with the posts that support
them. Each Purkinje cell receives information from about 200,000 parallel fibers,
and each parallel fiber contacts thousands of Purkinje cells aligned perpendicularly.
The climbing fibers, on the other hand, have a different distribution. They origi-
nate in the inferior olive, and ascend through one of the Purkinje cells, making
several synaptic contacts with it. This link is almost individual, with only one single
climbing fiber per Purkinje cell.
The synaptic connection between the climbing fiber and the Purkinje cell is the
strongest detected in the CNS: only an action potential in the climbing fiber can
144 4  Second Level: The Brainstem

produce a giant EPSP in the Purkinje cell, which discharge in response a salvo of
action potentials, constituting the so-called “complex spikes.” The terminals of the
mossy fibers, on the other hand, behave as common excitatory synapses of the CNS,
requiring the spatial and temporal summation of EPSP for the discharge of an action
potential in the Purkinje cell (“simple spikes”).
The mossy fibers discharge spontaneously at high frequencies (50–100/s). They are
the main control element of the Purkinje cells, and during the movement or by sensory
stimuli induce in the Purkinje cells discharges of simple spikes with a high frequency.
Only during the motor or autonomic learning process are complex spikes observed
in the Purkinje cells, evidence for the activation of the climbing fibers. This discharge
modulates the posterior response of Purkinje cells to the mossy fibers (post-tetanic
inhibition). Based on this observation, a role of importance for climbing fibers has
been postulated in the procedure memory (Chap. 6).
Other afferents of the cerebellar cortex are the “spider web” monoaminergic pro-
jections originating in the mesencephalon and diencephalon. For example, the hista-
minergic afferent system of the cerebellum, having been considered an essential
component of the direct hypothalamocerebellar circuits, originates from the TMN in
the hypothalamus. Unlike the mossy fibers and climbing fibers, the histaminergic
afferent fibers, a third type of cerebellar afferents, extend fine varicose fibers through-
out the cerebellar cortex and nuclei. Histamine directly excites the Purkinje cells and
granule cells in the cerebellar cortex, and the cerebellar nuclear neurons. Therefore,
the histaminergic afferents modulate these dominant components in parallel in the
cerebellar circuitry and consequently influence the final output of the cerebellum.
The hypothalamocerebellar histaminergic fibers/projections, bridging the autonomic
centers to somatic structures, play a critical role in the somatic–ANS integration.
The use of PET and fMRI to identify brain structures in humans involved in
cognitive functions indicated the important role of the cerebellum in language and
cognition. An evolutionary fact of importance is that the cerebellum, along with the
prefrontal cortex, has expanded significantly in humans relative to other hominids.
Until recently, it was assumed that this development was linked with, and resulted
in, motor skills of Homo sapiens. Many studies indicate that the human cerebellum
is activated in the absence of movement, when the subject performs cognitive and
verbal functions or experiences emotions.
In the evolution of the human cerebellum, the part that has developed most is the
cerebrocerebellum, whose nucleus is the dentate nucleus. In the dentate nucleus, an
older part is identifiable, present in the lower primates, in addition to a new portion, typi-
cal of humans. The new portions of the dentate are connected, via the thalamus, with the
association areas of the cerebral cortex involved in language and the limbic system.

24-h Rhythms in the Immune Response

The way in which the nervous system communicates with the immune system is
twofold: (a) via the endocrine system; (b) through the ANS, both sympathetic and
parasympathetic divisions, which supplies innervation to the lymph nodes, thymus,
spleen, and bone marrow. This interaction varies significantly in the three body
configurations during a 24-h cycle (Fig. 4.20).
24-h Rhythms in the Immune Response
145

Fig. 4.20  Neuroendocrine–immune communication varies significantly among the three body configurations during a 24-h cycle. The figure was prepared in
part using image vectors from Servier Medical Art (www.servier.com), licensed under the Creative Commons Attribution 3.0 Unported License (http://creative-
commons.org/license/by/3.0/)
146 4  Second Level: The Brainstem

In turn, because of immune reaction significant changes are verifiable in neuronal


activity. Various groups of hypothalamic neurons react to humoral signals (cytokines)
produced by immunocompetent cells, such IL-1, IL-6, TNF-α, and IFN-γ. Besides
participating in the normal humoral regulation of NREM sleep (Chap. 2), these sig-
nals produce the “illness behavior” accompanying the acute or chronic infection (loss
of appetite or anorexia, depressed motor activity, loss of interest in daily activities), in
addition to fever and the activation of the pituitary adrenal axis (Fig. 4.21).

Cerebral
vessel IL-1
IL-1 IL-1

2
4
IL-1 8

PG 3 Paraventricular
5 nucleus
5
OVLT Catecholamines
6
5 AP

5 NTS
ME
5 IL-1

6
1

IL-1 Vagus
nerve
1 Direct action on anterior pituitary and adrenal 1 7
2 BBB transport
a. Diffusion through altered BBB
b. Specific carriers IL-1

3 Prostaglandin-mediated function
4 Induction of intermediaries in BBB
5 Action on OCV
Visceral endings
6 Activation of catecholaminergic nuclei carotid bodies
7 Vagal stimulation Synthesis
8 Synthesis in CNS
OCV

Fig. 4.21  The different mechanisms by which an increase in cytokines (e.g., IL-1) results in cen-
tral changes (e.g., illness behavior, which includes increased central temperature). Prostaglandins
are intermediates in the effects of IL-1 on the circumventricular organs. Modified with permission
from Cardinali [3]
24-h Rhythms in the Immune Response 147

To deal with this subject, we first discuss briefly the bases of the immune response.
Then, we analyze the data linking sleep with immune responsiveness. A definite
immune configuration of wakefulness and NREM sleep can be achieved. However,
no configuration is known for the REM sleep itself, because of the methodological
problems of monitoring immune responses at short intervals.
The immune response consists of a first phase of antigen recognition, before the
later cell activation to produce molecules to eliminate it. This activation, which
constitutes the second phase of the immune response, is a set of processes that are
finely regulated. Indeed, an uncontrolled activation of leukocytes because of the
failure of the regulation could lead to the onset of illness or death of the individual.
The third and final phase is the destruction of the non-self, and involves the genera-
tion of inflammation and oxidation that allows the elimination of the pathogen.
The immune response comprises innate or nonspecific mechanisms and acquired
or specific mechanisms. Various pathogens, such as bacteria, viruses, fungi, and
parasites, trigger an immune response. Pathogens are recognized by a pattern recog-
nition receptor that gives rise to various signaling pathways to initiate activation of
adaptive innate immune response.
The unspecific response develops and acts immediately to deal with any foreign
agent that has managed to pass the natural barriers of the body. The innate response
defends the body from external pathogens and against any of its own cells that have
become dangerous, i.e., cancer cells. This response, which is fast because it is trig-
gered within seconds and lasts a few hours, is carried out by several cells and solu-
ble factors.
Innate immunity does not entail immune memory: it responds in the same form
and intensity to subsequent infections. It recognizes groups of pathogens, not the
subtle differences between them. Its cellular components are macrophages, neutro-
phils, basophils, eosinophils, natural killer (NK) cells, and cytokines (e.g., TNF-α,
IL-1β, IL-6, IL-8).
Phagocytes ingest and destroy infectious agents and NK cells bind to tumor and
virus-infected tissues to program them for destruction by apoptosis. In the innate
response, immediate mechanisms of action (within minutes) are followed by other
induced responses (lasting between 4 and 96 h). These mechanisms do not provide
lasting immunity protection.
The acquired immunity is specific and with immunological memory, mediated
by cells (T and B lymphocytes), cytokines and antibodies (Fig. 4.17). It comprises
humoral immunity (antibodies that recognize extracellular pathogens or foreign
molecules and make them sensitive to macrophage destruction) and cellular immu-
nity (cytotoxic T lymphocytes, CD8+, which recognize and destroy infected cells).
T helper (Th) lymphocytes (CD4+) coordinate the innate, cellular, and humoral
responses by means of numerous cytokines (Figs. 4.22 and 4.23). The acquired and
adaptive immunity possesses a system for self-regulation designed to avoid the
response of activation against the antigens to extend undesirably in time and space.
Lymphocytes can recognize, thanks to their specific receptors, millions of differ-
ent antigenic molecules, distinguishing even those that have a great structural simi-
larity, and they do it with great specificity. When the lymphocytes (B-cells in the
bone marrow and T-cells in the thymus) proliferate, a whole series of genetic
148 4  Second Level: The Brainstem

IgG
Blood marrow IgA
IgD
IgE
B cell
CD8+ cytotoxic T
Thymus lymphocyte

T cell
Immunosurveillance
of HLA-1
negative cells
CD4+ T
(cancer, viral
lymphocyte
infection)
Lymphoid precursor NK cell Th 1 Th 2

IL-12 IL-4

Plasma dendritic cell


Th 0
IL-4, IL-5
Stem cell IFN-γ
Extracellular
Intracellular
microorganisms
microorganisms
IL-12
AG presentation

Myeloid dendritic cell


TR

IL-1, IL-6 Regulatory T cells


Microbial phagocytosis

Monocyte / Macrophage

Microbial phagocytosis,
Secretion of inflammatory
products
Granulocyte

Fig. 4.22  Schematic model of intercellular interactions of immune cells in acquired immu-
nity. Stem cells differentiate into T cells, dendritic antigen-presenting cells, NK cells, macro-
phages, granulocytes, or B cells. Foreign antigen is processed by the dendritic cells, and the
peptide fragments are presented to T lymphocytes. Activation of CD4+ T cells leads to Th
lymphocytes. Activation of CD8+ T cells leads to the induction of cytotoxic T lymphocytes.
For the production of antibodies against the same antigen, the antigen binds within the recep-
tor complex and induces B cell maturation into plasma B cells secreting Ig. The figure was
prepared in part using image vectors from Servier Medical Art (www.servier.com), licensed
under the Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/
license/by/3.0/)

combinations allow the expression of millions of possible different receptors on


their membrane. These receptors (T cell receptors, TCRs, in T lymphocytes;
Fig. 4.23) and B cell receptors (BCR) in B lymphocytes (which are class M immu-
noglobulins) allow recognition in a very specific way of the millions of different
antigens that can be contacted throughout life. These receptors discriminate between
the self and the nonself, deciding whether to tolerate, in the first case, or destroy it,
in the second.
Lymphocytes have memory, which enables them to remember, when they recog-
nize an antigen, if it is the first time they have met it, or if there has already been a
previous interaction. Like episodic memory that is highly dependent on slow-wave
sleep (Chap. 6), the immune memory is also maximal in NREM sleep.
24-h Rhythms in the Immune Response 149

Th cells
CD4+
T lymphocyte
ANTIGEN
PRESENTATION

ACTIVATION EXPANSION DIFFERENTIATION

CD8+
T lymphocyte

MHC-II/peptide complex

MHC-I/peptide complex CTL

CD4 coreceptor Cytokines released by Th cells T cell receptor


CD8 coreceptor (TCR)
Cytokines released by CTL

Fig. 4.23  Antigen presentation and activation, expansion, and differentiation phases of T cells.
The figure was prepared in part using image vectors from Servier Medical Art (www.servier.com),
licensed under the Creative Commons Attribution 3.0 Unported License (http://creativecommons.
org/license/by/3.0/)

T helper (Th) lymphocytes constitute a fundamental component of acquired


immunity and their phenotypes determine the type of response (Fig. 4.22). Based on
the cytokine environment, expression of transcription factors and the cytokine
secretion pattern, Th lymphocytes can be differentiated into four phenotypes: Th1,
Th2, and Th17 (effector phenotypes; Fig. 4.24), and T regulatory (Treg), which
regulate any excessive response of effector lineages.
Th1 cells play a key role in the development of the inflammatory process through
the production of proinflammatory cytokines (IFN-γ, IL-1, IL-6, IL-12, IL-18).
Th17 cells are a subset of CD4+ T cells, which are highly relevant in inflammatory
processes. They express the transcription factor RORγt and produce IL-17 involved in
autoimmunity and in eliminating extracellular pathogens. The current concept is that
the inflammatory response involves Th1/Th17 control.
Th2 cells produce cytokines of anti-inflammatory activity (IL-4, IL-5, IL-10,
IL-13) and are responsible for the anti-inflammatory response.
Treg cells are a subpopulation of CD4+ (mainly CD25 +) lymphocytes express-
ing the transcription factor Foxp3 that controls the functions of effector cells.
The profile of the sympathetic predominance characteristic of prolonged wake-
fulness includes the increase in innate immunity, as revealed by an augmented
number of NK cells, a reduced proliferation of T lymphocytes, increased apopto-
sis of T lymphocytes and memory cells and a reduced number of “naive” cells.
Thymic involution and increased production of autoantibodies are characteristic
of the sympathetic predominance. Th2 cytokines (IL-4, IL10, IL-13) increase.
150 4  Second Level: The Brainstem

Group Cell products Effector cell Pathogens

IFN-γ Intracellular
Th1 bacteria
IL - 2
Fungi
IL-12 R Virus
Macrophages
Dendritic cells

IL-17A
Th17 IL-17F
Extracellular
IL-21
bacteria
IL-22
IL-23 R Fungi
Neutrophils

IL-4 Parasytes
Th2
IL-10
IL-13
IL-4 R IL-5
Eosinophils
Basophils

Fig. 4.24  Based on the cytokine environment, transcription factor expression, and cytokine secre-
tion pattern, Th lymphocytes can be differentiated into three effector phenotypes: Th1, Th2, and
Th17. The figure was prepared in part using image vectors from Servier Medical Art (www.servier.
com), licensed under the Creative Commons Attribution 3.0 Unported License (http://creativecom-
mons.org/license/by/3.0/)

The sympathetic innervation of lymphoid organs largely explains these effects


(Fig. 4.25). A crucial role of peripheral innervation in regulating cell behavior and
response to the microenvironment has recently emerged. In the hematopoietic
system, the ANS regulates stem cell niche homeostasis and regeneration and fine-
tunes the inflammatory response [36].
The ANS is activated in response to infection or injury and its products have
major effects on immune function and inflammation [37, 38]. The parasympathetic
nervous system also modulates inflammation by acting as an anti-inflammatory
neural circuit. The vagus nerve senses peripheral inflammation and transmits action
potentials to the brainstem, the area postrema, and the NTS.
Excessive activation of the immune system is prevented by anti-inflammatory
mediators such as corticosteroids and anti-inflammatory cytokines. Moreover, the
brain not only senses peripheral inflammation through vagal afferent nerve fibers,
including those originating in the carotid bodies [39], but also provides an inte-
grated response dampening the immune system through vagal efferents [37]. This
24-h Rhythms in the Immune Response 151

Sympathetic
nerve ending
Plasma E
IL-12
TNF-α
IFN-γ NE
IFN-γ
NE
Ag NK
IL-12
IL-2, IL-3
Monocyte IFN-γ, TNF-β, NK
Th1
γ

Tc
-
IL

IFN

GM-CSF
-2

Th1 cytokines
Cellular immunity
Th0 IL-2, IFN-γ, IL3,
TNF-β, GM-CSF Humoral immunity

IL-4 Th2 cytokines


IL-3, IL-4 B
Monocyte Th2 IL-5, IL-10, Ac
IL-13

Ag IL-10
NE NE
NE

Plasma E

Sympathetic
nerve ending

Fig. 4.25  The sympathetic nervous system promotes Th2 responses (humoral immunity) and
inhibits Th1 responses (cellular immunity). Ag antigen, Ac antibodies, R receptor. The figure was
prepared in part using image vectors from Servier Medical Art (www.servier.com), licensed under
the Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/
by/3.0/)

so-called anti-inflammatory pathway has been introduced as a second system by


which the immune system is modulated by ANS. In sepsis, the anti-inflammatory
effect is mediated by the modulation of splenic macrophages, whereas in the gut,
vagal nerve fibers synapse with enteric cholinergic neurons interacting with resident
intestinal macrophages [40]. In the gastrointestinal tract, the microbiome contrib-
utes significantly to the activation and inhibition of autonomic control and has an
impact on the set of the neuroinflammatory inhibitory reflex mediated by the cho-
linergic nervous system [41].
Valuable information on autonomic nerve regulation of immune function has
derived from studies in the thymus and submaxillary lymph nodes. The thymus
plays a critical role in establishing and maintaining the peripheral T-cell pool. It
does so by providing a microenvironment within which T-cell precursors differenti-
ate and undergo selection processes to create a functional population of major his-
tocompatibility complex-restricted, self-tolerant T cells. These cells are central to
adaptive immunity. Thymic T-cell development is influenced by sympathetic norad-
renergic signaling [42].
As for thyroid and parathyroid glands (Chap. 3), submaxillary lymph nodes receive
sympathetic innervation from the neurons located in the SCG, whereas their
152 4  Second Level: The Brainstem

parasympathetic innervation derives from the lingual nerve, branch of the facial nerve,
and reaches the submaxillary glands via the chorda tympani. Based on this, an experi-
mental model was developed comprising the submaxillary lymph nodes and the ipsi-
lateral local manipulation of the sympathetic nerves and/or the ipsilateral manipulation
of regional parasympathetic nerves via the chorda tympani [43]. In the submaxillary
lymph model, reactive immune homeostasis was studied by subjecting unilaterally
denervated rats to different types of stress. An inhibitory effect of the ipsilateral sym-
pathetic nerve ablation on the sympathetic-driven immunosuppression and an
increased response to stress after parasympathetic denervation were observed. In
addition, T cells express choline acetyltransferase, the ACh-synthesizing enzyme and
immunological T cell activation enhances ACh synthesis, suggesting that lympho-
cytic cholinergic activity might be related to immunological activity [44].
The sympathetic system selectively inhibits Th-1 responses, while favoring Th-2
responses, whereas the parasympathetic system has the opposite effect (Fig. 4.26).
NE and E, through activation of β-adrenoceptors, suppresses the production of Th-1
cytokines, such as IL-12, TNF-α, and IFN-γ by the antigen-presenting cells and
Th-1 cells, while promoting the production of Th-2 cytokines (e.g., IL-10).
Sympathetic activation also shifts toward a Th-2 response via the increased produc-
tion of glucocorticoids.

Th 1 cell
Parasympathetically
promoted

+ IL-12
Th 0 cell IL-2, IFNγ
- IL-4

+ IL-12
IL-2
IFNγ, IL-4 Th 2 cell
- IFNγ

Sympathetically
promoted

IL-4, IL-5

Fig. 4.26  Differentiation of T helper cells in Th1 and Th2 cells. Repeated stimulation in the pres-
ence of IL-12 produced by macrophages causes differentiation to Th1 cells that produce IL-2 and
interferon (IFN) γ, both cytokines that are very effective at increasing immune responses involving
macrophages and other phagocytes and cellular immunity. Stimulation in the presence of IL-4
promotes the development of Th2 cells, effective in producing cytokines acting on mast cell and
eosinophil responses and on humoral type immunity. Modified with permission from Cardinali [3]
24-h Rhythms in the Immune Response 153

The first half of sleep represents a proinflammatory state characterized by the


downregulation of the two reaction systems, the hypothalamic–pituitary–adrenal
axis, and the sympathetic nervous system. Mediators that serve cell growth, differ-
entiation, and restoration, such as growth hormone (GH), prolactin (PRL), and mel-
atonin, show increased blood levels during early sleep (Fig. 4.27). During this
period, plasma leptin, which is released by adipocytes as a signal indicative of the
size of the lipid deposits, also increases.
Melatonin, GH, PRL, and leptin are synergistic and support the activation,
proliferation and differentiation of immune cells, with the production of proin-
flammatory cytokines (IL-1, IL-12, TNF-α) and Th1 cytokines (IFN-γ). In con-
trast, cortisol and catecholamines generally suppress these immune functions in
an anti-­inflammatory manner. In rat submaxillary lymph nodes, a daily rhythm of
lymph cell proliferation and response is found with maxima during the resting
phase of the 24-h cycle.

Antibodies
nREM
sleep

GH, PRL,
melatonin,
Antigen leptin B lymphocytes

IL-12

Dendritic cell Th cells

T lymphocytes

ANTIGEN IMMUNOLOGICAL RESPONSE TO RE-


PRESENTATION MEMORY EXPOSURE TO THE
ANTIGEN

Fig. 4.27  Endocrine changes during early sleep are critical for immune memory-related phenom-
ena. These include the interaction between antigen-presenting cells and T cells, which increases
IL-12 production, the shift of the Th1/Th2 cytokine balance toward Th1 predominance, the migra-
tion of virgin T cells to lymph nodes, and the proliferation of Th cells. The figure was prepared in
part using image vectors from Servier Medical Art (www.servier.com), licensed under the Creative
Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)
154 4  Second Level: The Brainstem

There is growing evidence that indicates a significant diurnal rhythm for all
blood cell types in humans [45]. Circulating lymphocytes show peak levels during
the night, whereas granulocytes and NK/dendritic cells peak during the late after-
noon. In addition, circulating granulocyte rhythm is severely affected by prolonged
wakefulness, with higher circulating levels, lower amplitude, and loss of
rhythmicity.
Additional data indicates a shift in the Th1/Th2 cytokine balance toward Th1
predominance during nocturnal sleep in humans, tracked by IFN-γ/IL-4-producing
CD4+ cells. There is a shift toward Th1 cytokines (cell-mediated immunity) during
early sleep, when NREM sleep is predominant, favoring the cellular aspects of
adaptive immune responses over anti-inflammatory activity, or Th2 cytokines
(humoral immunity). The effect is reversed during the late part of sleep, when REM
sleep is dominant [46].
There were two possible mechanisms by which the CNS is regulating circadian
activity in immune organs. One could involve purely neuroendocrine signals as cir-
culating melatonin or glucocorticoids. The other is neural, involving the local auto-
nomic nerves. In several studies, it was verified that the daily changes in cell
proliferation in the submaxillary lymph nodes were linked in part to a circadian sig-
nal reaching the tissue through local sympathetic innervation [47]. In addition, in
pinealectomized rats, cell proliferation in lymph nodes was reduced by half, main-
taining its daily maximum at midday. Melatonin administration restored the levels of
cell proliferation in the submaxillary lymph nodes of both pinealectomized and
SCGx rats. A significant effect of melatonin in maintaining the normal diurnal rhyth-
micity of neurotransmitter synthesis and release in various neuroimmune territories
of the sympathetic nervous system was also reported [48]. Accumulating evidence
indicates that melatonin exerts a biphasic immune effect: in basal conditions or a
depressed immune response, melatonin increases immune activity; in conditions of
augmented immune and inflammatory reaction, melatonin decreases it [48].
The endocrine milieu during early sleep is critical for several phenomena linked
to immunological memory: the interaction between antigen-presenting cells and T
cells to increase production of IL-12, shifting the balance of Th1/Th2 to Th1 domi-
nance, the migration of naïve T cells to the lymph nodes, and the proliferation of Th
cells (Figs. 4.27 and 4.28). In other words, imitating the role that slow-wave sleep
plays in the neural mechanisms of episodic memory (Chap. 6), the endocrine milieu
during early sleep promotes the initiation of the Th1 immune response, leading to
the formation of long-lasting immunological memory.
Sleep facilitates extravasation of T cells and their redistribution to lymph nodes
(Fig. 4.28). Sleep improves IL-12 production by dendritic cells, which are precur-
sors of mature medullary antigen-presenting cells and monocytes. IL-12 is a key
cytokine for the induction of Th1-type adaptive immune responses. Production of
the main anti-inflammatory cytokine IL-10 by monocytes is reduced, whereas that
of IL-17 (which stimulates the growth and differentiation of T cells) increases in
sleep. The increase in cortisol, E, and NE in wakefulness produces a strong anti-
inflammatory effect that disrupts the sleep proinflammatory response. IL-10 is pro-
duced by stimulated monocytes and reaches peak levels in the morning [46].
24-h Rhythms in the Immune Response 155

“NAIVE” T
CELLS

NREM sleep
Rapid cell
WAKEFULNESS
exchange between
Cortisol - induced
circulation and
redistribution to bone
lymph nodes
marrow

2300 0700 2000 h

Fig. 4.28  In slow-wave sleep, the virgin lymphocytes enter the lymphoid tissue. In wakefulness,
they pass to the blood by the action of cortisol and are found in bone marrow. The figure was pre-
pared in part using image vectors from Servier Medical Art (www.servier.com), licensed under the
Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

Slow-wave sleep is the right time to initiate immune responses because of the
proinflammatory neuroimmune environment described above. If the sleep-related
production of proinflammatory cytokines were to occur during wakefulness, it
would cause discomfort, fatigue, immobility, pain, etc., adaptatively incompatible
with the requirements of the waking state. The encounter with the pathogen usually
occurs during waking and is adaptatively controlled by the innate immune defense,
predominant during wakefulness, which protects the peripheral tissues and spleen
before the slow process of adaptive immunity in the lymph nodes. Cytotoxic effec-
tor cells are at their maximum during the waking period [49, 50].
The predominance of the sympathetic system in wakefulness mobilizes a subset of
white blood cells (“stress” leukocytes) with potent cytotoxic effect (Fig. 4.29). They
are a phylogenetically primitive group of immune cells with a high level of expression
of β-adrenoceptor that multiply in peripheral blood during wakefulness. “Stress” leu-
kocytes are mature cytotoxic cells with a long replicative history and a very short
156 4  Second Level: The Brainstem

Stress Leukocytes

WAKEFULNESS
Epinephrine-induced
redistribution from
marginal pool)

Cell number

2300
300 0700 2000 h

Fig. 4.29  In wakefulness, the ANS mobilizes a subset of leukocytes (“stress leucocytes ”), with
potent cytotoxic effect. The figure was prepared in part using image vectors from Servier Medical
Art (www.servier.com), licensed under the Creative Commons Attribution 3.0 Unported License
(http://creativecommons.org/license/by/3.0/)

telomere. They release toxic substances (granzyme, perforin, IFN-γ, Fas ligand apop-
tosis inducer) that eliminate infected or malignant cells. “Stress” leukocytes reside in
the marginal pool and are attached to the endothelium of post-­capillary venules. They
are mobilized within seconds after stimulation of β2-adrenoceptor by E and do not
recirculate (Fig. 4.29). They have a short half-life and do not need costimulation,
which explains their immediate effector potential. Proinflammatory neutrophils,
monocytes, and dendritic cells are mobilized after sympathetic activation via
α-adrenoceptors. Overall, these effects explain the “adrenergic leukocytosis,” which is
part of immunosurveillance for rapid responses during wakefulness [50].
Sleep manipulation strongly affects distinct immune parameters, including leuko-
cyte numbers, cytokine production, and cytotoxic activity of immune cells. Whereas
short-term total sleep deprivation of only a single night seems to primarily compromise
adaptive immune functions, as has been revealed in several human studies relying on
vaccination as an experimental model of acute infection, prolonged sleep restriction
induces an immunological condition mainly characterized by small but distinct
increases in inflammatory markers [51]. This is often referred to as a “low-grade
inflammation,” and it is of major clinical relevance because it has been associated with
an increased risk for developing, for example, diabetes mellitus and cardiovascular
diseases. Increases in inflammatory markers have also been observed in habitual short
24-h Rhythms in the Immune Response 157

sleepers and in patients with primary insomnia and may represent a mechanism that
also mediates increased susceptibility to clinical signs of common cold infections.
One night of total sleep deprivation strikingly decreases the number of dendritic
cells producing IL-12, the main inducer of the Th1 response to approximately 40%
of normal sleep levels. Regarding other cytokines, IL-2 is increased in sleeping
subjects, compared with waking conditions, further supporting the notion that sleep
facilitates Th1 immunity. The acute enhancing effect of sleep on lipopolysaccharide-­
stimulated monocytic TNF-α production adds to the notion that nocturnal sleep
favors an immune defense to a microbial challenge [52].
In humans, the number of circulating T cells shows a circadian rhythm with peak
counts during the night and a steep decline in the morning. Sleep per se appears to
counter this rhythm by acutely reducing the number of total T cells in circulation. The
T cell population, however, is rather heterogeneous, comprising various subpopula-
tions with distinctive features and functions and different circadian rhythms. When
eight different T cell subsets (naïve, central memory, effector memory, and effector
CD4+ and CD8+ T cells) over a 24-h period under conditions of sustained wakeful-
ness were compared with a regular sleep–wake cycle in 14 healthy young men, sleep
reduced the number of all T cell subsets during the nighttime [46]. The changes are
comparable with changes seen for example after vaccination and are therefore likely
to be of physiological relevance [53]. The changes in immune response after a short
period of sleep deprivation in humans are summarized in Table 4.1.
Reduced sleep duration for a longer period (e.g., 10 days) caused increased IL-6
levels, which are correlated with increased pain sensitivity. Corroborating this find-
ing, cross-sectional analyses indicated that shorter sleep is associated with higher
levels of inflammatory markers, such as C-reactive protein, and IL-6. Interestingly,
it has been shown that a 2-h nap during the daytime is able to reverse the effects of
one night of sleep deprivation, particularly the increased IL-6 and cortisol levels,
and consequently improve alertness and performance in the sleep-deprived subjects,
indicating that short naps counterbalance sleep deprivation by restoring the immune
and hormonal milieu necessary for proper immune response [54].

Table 4.1  Changes in Sleep deprivation


immune parameters after short
Circulating cells
sleep deprivation in humans
Th cells, cytotoxic cells, activated T cells, ↑
NK cells
Monocytes, dendritic cell precursors ↓
B lymphocytes ↔
Cytokines
IL-2, IL-7, IL-12, IL-18, TNF-α, IFN-γ/ ↓
IL-4 ratio
IL-10, IL-4, IL-6, TNF-α, IFN-γ ↑
Other
T cell proliferation ↓
Regulatory T cell activity ↓
NK cell activity ↑
Complement system ↓
Response to vaccines ↓
158 4  Second Level: The Brainstem

Extended periods of partial or total sleep deprivation are associated with altera-
tions in many aspects of immunity, such as increased counts of white blood cells,
granulocytes, and monocytes. The analysis of cytokine activity indicates the activa-
tion of innate immunity after extended sleep deprivation. Although during normal
sleep, proinflammatory cytokines such as IL-6 and IFN-α do not change, it is note-
worthy that both are involved in the regulation of early innate immune response and
both increase after sleep deprivation. As we have discussed in Chap. 2, various cyto-
kines influence the quality and timing of sleep. The most frequently studied in this
respect are IL-1, IL-6, and TNF-α. They are somnogenic and induce fatigue, besides
being proinflammatory. It is interesting that these cytokines augment in sleep disor-
ders that cause excessive daytime sleepiness, such as sleep apnea or narcolepsy.
Hence, a reciprocal interaction between sleep and the immune system occurs in
humans. On one hand, sleep pressure (or sleep demand) and NREM are increased
during an experimental immune challenge, suggesting that inflammatory mediators
released during the immune response might be modulators of both physiological
and pathological sleep, by acting on neurotransmission systems. To date, the effects
of IL-1β and TNF-α effects (produced either in periphery or the CNS) have been
extensively studied and are implicated in the modulation of sleep response in both
animals and humans (Fig. 4.30). On the other hand, sleep is required for a proper

Acute sleep deprivation Chronic sleep deprivation


↓ Adaptive immunity ↑ Innate immunity/inflammation
↓ Lymphocytes ↑ IL-6, IL-1β, TNF-α
↓Vaccine antibody response ↑ Leukocyte trafficking

GH, PRL, Melatonin, Leptin


Cortisol IL-1β , TNF-α

Lymph nodes
Thymus

Spleen
Infectious/inflammatory condition:
↑ Sleep pressure
↑ NREM sleep
↓ REM sleep

Fig. 4.30  Interactions between acute or chronic sleep deprivation and inflammatory processes
24-h Rhythms in the Immune Response 159

immune response, possibly because of the endocrine milieu elicited by NREM dur-
ing nighttime sleep. Hormones such as GH, PRL, leptin, and melatonin peak during
NREM sleep and have facilitatory effects in the adaptive immune response, particu-
larly on Th1 immunity. Short or prolonged periods of sleep loss have a negative
impact on NREM sleep and disrupt this hormonal pattern, preventing these immu-
nosupportive actions and leading to immune suppression. During an ongoing
immune response, pro-inflammatory and Th1 cytokines feed back to the brain to
enhance NREM sleep [46].
Therefore, it is possible to speculate that NREM sleep acts as an adjuvant to the
optimal immune response, creating a hormonal milieu that favors type 1 cytokines
and Th1 immunity. In turn, the minimum sleep requirement (or sleep pressure)
increased during an infection or immune response, possibly resulting in immune
suppression if this minimum requirement is not reached. These outcomes create a
vicious cycle between sleep loss, immune modulation, and infectious–inflammatory
diseases (Fig. 4.31).
As stated, enough data support the association between NREM sleep and
immune responsiveness. None is known concerning REM sleep itself, because
of the methodological problems of monitoring immune responses at short
intervals.

GH, PRL, Melatonin, Leptin.


Parasympathetic predominance.

24
22 2

20 4

18 6

16 8

14 10
12

ACTH, Cortisol.
Sympathetic predominance.

Fig. 4.31  During the first half of the night there is a proinflammatory state, with downregulation
of the two reaction systems, the hypothalamic-pituitary–adrenal axis and the sympathetic nervous
system. During the second half of the night, cortisol and catecholamines generally suppress
immune functions and act in an anti-inflammatory manner
160 4  Second Level: The Brainstem

24-h Rhythms in Gastrointestinal Function

Symptoms of several gastrointestinal tract conditions, e.g., gastroesophageal reflux dis-


order, peptic ulcer disease, biliary colic, and cyclic vomiting syndrome, primarily mani-
fest or are worse nocturnally, as are the serious events of peptic ulcer perforation and
esophageal, gastric, and congestive gastropathic variceal hemorrhage [55–58]. All these
emphasize the necessity to analyze the changes in enteric ANS function in the two phys-
iological stages of sleep, NREM and REM sleep, compared with wakefulness.
The changes in gastrointestinal functioning during sleep are different depending on
the organ studied and its function within the gastrointestinal system. For example,
spontaneous esophageal function is markedly reduced during sleep because there is no
need for this organ to function except in the event of food entering the upper esopha-
geal area. This rarely occurs without volitional swallowing, which is diminished sig-
nificantly during sleep. On the other hand, rectal motor activity persists during sleep,
which appears to be a mechanism necessary to preserve continence during sleep.
All activity of the gastrointestinal tract is periodic, affecting motility, gall bladder
function, the gastrointestinal blood flow, gastric pancreatic and bile secretions, the
rate of nutrient absorption, and many other physiological events (Fig. 4.32). This

GASTROINTESTINAL TRACT

Acid secretion Production GI immune


by the stomach of digestive system
enzymes

Mucosal Absorption Transport of


Motility barrier nutrients
protection

Fig. 4.32  Circadian and ultradian variations are present in most functions of the gastrointestinal
tract. The figure was prepared in part using image vectors from Servier Medical Art (www.servier.
com), licensed under the Creative Commons Attribution 3.0 Unported License (http://creativecom-
mons.org/license/by/3.0/)
24-h Rhythms in Gastrointestinal Function 161

periodicity is the consequence of autonomous rhythms of different frequencies, both


ultradian and circadian, and of masking effects of meal schedule synchronizers.
To discern the relative roles of the circadian pacemaker and the effects of sleep,
it is necessary to study the changes in circadian patterns in the absence of the effects
of sleep, such as during sleep deprivation or when it occurs in atypical hours. A
summary of the data obtained regarding the digestive function is shown in Table 4.2.
In addition to the circadian aspects, ultradian rhythms of 90–120 min deserve
special attention, and are characteristically found in most of the functions of the
digestive tract. This period coincides with that of the alternating phases of NREM–
REM overnight sleep phenomenon and continues during the day in cycles of activ-
ity–rest known as the basic rest–activity cycle, i.e., cycles at the alert level that
occur approximately every 90–120 min [59]. As discussed in Chap. 2, this fre-
quency value is a harmonic of the 24-h cycle.
Digestion begins in the mouth with the process of chewing, salivation, and swal-
lowing. These initial digestive processes are related to the waking state, with little
information about their circadian links. The swallowing rate decreases from 25/h dur-
ing waking to 5/h during sleep. Swallows usually follow sleep awakenings [55, 56].
Swallowing must overcome the pressure of the upper esophageal sphincter for the
bolus to enter the esophagus (Fig. 4.33). The upper esophageal sphincter pressure
decreases during N3 sleep to 25% of wakefulness. During sleep, the airway is pro-
tected from aspiration by the cricopharyngeal muscle. This muscle maintains a pres-
sure barrier in the proximal esophagus to inhibit food aspiration. Like the diaphragm
and the muscles in middle ear, the cricopharyngeal muscle differs from the rest of the
striated muscles to keep the tone throughout all stages of sleep. The upper esopha-
geal sphincter is tonically contracted, and a pressure between 40 and 80 mmHg usu-
ally exists because of the function of this sphincter. Swallowing induces a reflex
relaxation to allow the positioning of food and liquids in the upper esophagus, where
the normal peristaltic mechanism transports these materials into the stomach.
During wakefulness, gastroesophageal reflux may occur in normal people. This
occurrence is primarily postprandial in normal subjects and is associated with mul-
tiple episodes of reflux that are neutralized relatively rapidly (in less than 5 min;
Fig. 4.34). Swallowing triggers the primary esophageal peristalsis, which decreases
in the deepest stage of slow-wave sleep and REM sleep. The inferior esophageal
sphincter decreases in tone during sleep, but remains above intragastric pressure.

Table 4.2  Circadian and homeostatic influence on digestive function


Circadian influence Homeostatic influence
Digestive function (C process) (S process)
Salival output Possible on pH Yes
Swallowing/esophageal motility No Yes
Gastric acid secretion Yes Controversial
Gastric motility Possible Yes
Intestinal absorption Yes Unknown
Intestinal motility Possible Yes
Colon motility Possible Yes
Rectus–anal function Possible Yes
162 4  Second Level: The Brainstem

Swallowing
Rest Postswallowing
100 mmHg

0
0
5 sec Swallowing
Rest Postswallowing
100 mmHg

0
0
5 sec
Swallowing
Rest Postswallowing
100 mmHg

0
0
5 sec

Fig. 4.33  Swallowing starts primary esophageal peristalsis. It must overcome the pressure of the
upper esophageal sphincter as the bolus enters the esophagus. The figure was prepared in part
using image vectors from Servier Medical Art (www.servier.com), licensed under the Creative
Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

A drop of 5–30 s in inferior esophageal sphincter pressure with sudden increases in


intra-abdominal pressure cause gastroesophageal reflux during sleep.
Gastroesophageal reflux is exacerbated during sleep [58]. The clearance of acid
into the esophagus occurs in two distinct phases: (a) a first phase (volume clearance)
is caused by swallowing, leading to primary esophageal contractions; (b) a second
phase (acid clearance) neutralizes any acidic residues of pH <4 (the most important
function of saliva). Both phases are inhibited during sleep.
During sleep, prolongation in acid clearance is due to several factors. First, and
perhaps most importantly, there is a delay in the conscious response to acid in the
esophagus during sleep. Studies have documented that an arousal response almost
invariably precedes the initiation of swallowing. An inverse relationship has been
described between the acid clearance time and the amount of time the individual
spends awake during the acid clearance interval. If an individual responds with an
awakening and subsequent swallowing when acid is infused in the distal esophagus
during sleep, clearance is substantially faster than if the individual has a prolonged
latency to the initial arousal response [58].
24-h Rhythms in Gastrointestinal Function 163

Inferior esophageal sphincter (IES)

IES IES

Sphincter closed

Transient
opening of IES

Fig. 4.34  The inferior esophageal sphincter (IES) decreases in tone during sleep, but remains
above intragastric pressure. A drop in IES pressure with sudden changes in intra-abdominal pres-
sure cause gastroesophageal reflux during sleep. The figure was prepared in part using image vec-
tors from Servier Medical Art (www.servier.com), licensed under the Creative Commons
Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

Another important aspect of complete neutralization of the acid in distal esopha-


gus is the salivary flow. Saliva is essential for the neutralization of the acidic esopha-
gus. Salivary flow stops completely with the onset of sleep, which would therefore
substantially retard the acid neutralization.
Although swallowing frequency diminishes during sleep, esophagic peristalsis is
maintained. The frequency of primary contractions in the esophagus (peristaltic
contractions preceded by a swallow) diminish progressively from stage N1 to stage
N3 sleep. Secondary peristaltic contractions (spontaneous contractions) showed a
similar decline from waking to stage N3 sleep, but showed a significant recovery
during REM sleep.
The inferior esophageal sphincter prevents reflux of the stomach contents into
the esophagus. In the upright position, postprandial gastric distention causes brief
relaxation of the inferior esophageal sphincter with transient reflux that is quickly
cleared (Fig. 4.34). Several factors assist in the rapid clearance or neutralization of
stomach contents: (a) volume clearance: two or three swallows induce clearance of
reflux material; (b) acid neutralization: saliva itself buffers the acidity of the refluxed
material. These mechanisms quickly return the distal esophagus to pH 5.5–6.5 (nor-
mal esophageal pH) during wakefulness.
Nocturnal gastroesophageal reflux is common, with up to 10% of the population
reporting symptoms of nocturnal reflux in survey studies. As noted, nocturnal gastro-
esophageal reflux is potentially more injurious than a diurnal one because acid clearance
164 4  Second Level: The Brainstem

mechanisms are impaired during sleep. When episodes of gastroesophageal reflux occur
during sleep, acid contact time is prolonged. During sleep, salivary flow virtually stops
and the frequency of swallowing is very decreased. The clearance of refluxed material
occurs only after arousal from sleep. Most nocturnal gastroesophageal reflux episodes
occur during prolonged waking or after arousal from stage N2 sleep [58].
The stomach plays an important role in food storage and its controlled emptying
into the small intestine. The main functions of the stomach are: (a) acidification of
food; (b) flow control of the bolus into the duodenum. Ingested food accumulates in
the upper stomach, the fundus, and once processed, is driven into the antrum and
hence passes to the duodenum through the pyloric sphincter in discrete amounts.
The gastric emptying rate is influenced by the nature of food: a liquid is discharged
faster than a solid. Furthermore, other characteristics of the food such as liquid
osmolality, acidity, caloric content, etc., can affect gastric retention time [55, 58].
The rate of gastric emptying has a rhythm and changes with the time of day, vary-
ing considerably between meals. For example, breakfast and dinner have very differ-
ent gastric retention. In the morning, the gastric emptying rate is higher than at night.
By nasogastric recording, there is a basal rate of production of HCl, with a maxi-
mum around the beginning of the night, both in normal healthy individuals and in
patients with duodenal ulcers. Ulcer patients produce a greater amount of acid than
normal, but the basic pattern of daily secretion tends to be the same. In general, gas-
tric acidity increases at night, between 22:00 and 02:00 h, and decreases in the morn-
ing. This rhythm persists in the absence of sleep, indicating their circadian nature.
The electrical activity of the stomach also has a very strong daily rhythm [55, 58].
Patients with duodenal ulcers lose rhythms of acid secretion, with overproduc-
tion of acid throughout the day and night. Overproduction of HCl is linked to
chronic Helicobacter pylori infection. Helicobacter pylori is a bacterium that lives
in the mucous layer of the stomach, where the enzyme urease is active with the
conversion of urea into CO2 and ammonia, which is a buffer of luminal acid that
protects the microorganism. H. pylori also secretes proteins that modulate the
immune response and directly alter cell signaling pathways of the mucosa. More
than half of the world’s population is infected with H. pylori. In most cases, the
infection, although chronic, is mild and causes no symptoms. In some individuals,
however, infection leads to increased acid secretion and symptomatic inflammation
causes ulceration of the stomach or duodenum. Almost all duodenal peptic ulcers
and about half of gastric peptic ulcers show H. pylori infection [55, 58].
The fall of somatostatin induced by H. pylori increases the parietal cell mass and
overproduction of acid without a circadian pattern. H. pylori infection interferes
with the release of clover protein, increasing the risk of mucosal damage. Trefoil
protein 1 (clover) is a protein from the digestive tract encoded by the TFF1 gene that
shows circadian rhythmicity. The release of clover protein is reduced by sleep depri-
vation, indicating a relationship with both circadian and homeostatic processes.
Melatonin is a strong inhibitor of acid secretion and increases the release of gastrin,
resulting in stimulation of the lower esophageal sphincter activity (Fig. 4.35). Both
actions protect the esophagus, minimizing contact with reflux. Furthermore, melatonin
reverses inflammatory lesions and reduces lipid peroxidation produced by contact with
24-h Rhythms in Gastrointestinal Function 165

EE
Melatonin

GR
Parietal
Ghrelin cell

ECL
G

Somatostatin D

Fig. 4.35  Food-entrained oscillators (FEO) in the stomach. FEO activity precedes and promotes
food and eating behavior. A location of FEO in the stomach is the oxyntic glandular cells (GR)
co-expressing ghrelin and circadian clock proteins. Entero-endocrine cells (EE) release melatonin,
which is a strong inhibitor of acid secretion antagonizing the effect of histamine and increasing the
release of gastrin, resulting in the stimulation of lower esophageal sphincter activity. D somatostatin-­
releasing cell, ECL enterochromaffin-like cells. See also Fig. 3.25. The figure was prepared in part
using image vectors from Servier Medical Art (www.servier.com), licensed under the Creative
Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

gastric juice. Melatonergic enteroendocrine cells produce about 500 times more melato-
nin than the pineal gland. There is an entero-hepatic circulation for melatonin [60, 61].
The origin of rhythmicity in gastric secretion of HCl does not appear to be endo-
crine as it is not coincident with the parallel rhythms of plasma gastrin (a major
hormone that stimulates gastric secretion). Nerve stimulation is more important
than endogenous circadian rhythmicity as rhythmicity in gastric secretion of HCl
disappears in vagotomized patients.
Pepsin, a protease that degrades food proteins in the stomach, is secreted by cells
of the gastric mucosa as an inactive pepsinogen, passing to the active form, pepsin,
by the action of HCl and subsequently also by the action of activated pepsin itself.
Like HCl, protease pepsin activity follows a daily rate, with a maximum after din-
ner. As the presence of HCl and very acidic pH are essential requirements for the
166 4  Second Level: The Brainstem

action of pepsin, the coupling rate of secretion of the protease with the rhythm of
HCl and pH optimizes the action on the substrates present in the stomach [55, 58].
The nocturnal secretion of gastric acid and pepsin is in principle beneficial for
the body, as it helps to clean up the debris of food that may occur after meals. In a
healthy organism, the rhythms of secretion of HCl and pepsin activity are in phase
with the production of mucus and bicarbonate by providing a natural defense. The
pathological alteration of the phase relationship of these rhythms increases the vul-
nerability of the gastric epithelium, leading to ulcerogenesis.
Gastrointestinal tract motility is characterized by rhythmic waves of membrane
depolarization and contraction of the muscle fiber, with a shape and frequency char-
acteristics of each portion of the gastrointestinal tract. In the stomach, for example,
the depolarization consists of a spike action potential followed by a plateau phase
before repolarization. In the small intestine, in contrast, the action potentials are in
spikes, and in the colon the spikes are prolonged (haustrations).
Like many other variables of the digestive tract, such as motility, the stomach
presents ultradian fluctuations in its rate of gastric secretion. These ultradian
rhythms are closely related to the migrating motor complex (MMC) during the
interdigestive period. During the interdigestive phase, increases in gastric acid and
in pepsin production occur every 90–120 min, preceding the appearance of the
MMC in the duodenum [55, 58].
The MMC (Fig. 4.36) consists of four phases: phase 1, rest; phase 2, irregular con-
tractions; phase 3, 9–12 regular contractions/min; phase 4, transition between phases

Gastrontestinal motor activity

The MMC consists of four


phases:
- Phase 1: rest
- Phase 2: irregular
contractions
- Phase 3: 9-12 regular
contractions / min
- Phase 4: transition between
phases 3 and 1
The cycle time is 90 min.

Fig. 4.36  The four phases of migratory motor complexes. The figure was prepared in part using
image vectors from Servier Medical Art (www.servier.com), licensed under the Creative Commons
Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)
24-h Rhythms in Gastrointestinal Function 167

3 and 1. The cycle time is about 90 min, such as NREM/REM sleep alternation. Food
ingestion establishes a pattern of vigorous contraction throughout the distal stomach
and small bowel. If no food enters the stomach, the MMC cycle has a period of about
90 min (Fig. 4.36).
Gastric motor function is controlled by a gastric pacemaker located in the smooth
muscle of the greater curvature of the stomach. By electrical recordings it was shown
that the amplitude of gastric engine cycle decreases in NREM sleep and is restored in
REM sleep. In normal subjects, acid secretion, water secretion, and the fractional rate
of emptying all showed significant decrements during sleep. There did not appear to
be any differences between REM and NREM sleep, but these measures demonstrated
a significant difference between presleep waking and REM sleep. Data obtained by
the use of radionuclide emptying assessments suggest that this difference might be
circadian, rather than sleep-dependent. Some studies have shown a marked delay in
the gastric emptying of solids in the evening, compared with the morning.
Motility of the small intestine is characterized by a high frequency of targeting
movements against peristalsis. Asynchronous contraction of the circular muscle
layers and outer longitudinal internal (segmentation) help the mixing of food with
digestive secretions and improve absorption by allowing a constant renewal of
food contact area of intestinal epithelium. During the interdigestive periods, intes-
tinal motility does not stop, but continues as a characteristic rhythmic pattern.
The MMCs are generated in many ultradian pacemakers distributed along the
digestive tract and coordinated by the enteric nervous system. If the small intestine
is cut into several pieces and subsequently subjected to reanastomosis, each frag-
ment starts a separate cycle and the coupling of the MMC reappears within 45–60
days, coinciding with the intrinsic regeneration of nerves. Control of the rhythms of
the 24-h propagation velocity of the MMC resides in the CNS, as vagotomy results
in the suppression of these rhythms [55, 58].
In the mouse colon, a functional circadian clock and a subset of rhythmically
expressed genes have a direct impact on colonic motility, such as the contractile
response of colonic tissue to ACh, stool output, and intracolonic pressure changes.
These effects are attenuated in mice with disrupted clock function [62]. Clock-­
controlled transcription of genes, such as choline acetyltransferase and neuronal
NO synthase, leads to the rhythmic release of ACh and NO, which initiate diverse
biochemical, cellular, and physiological processes within the colonic circular mus-
cle, which may in turn, through a cascade of second-order messengers and various
signaling pathways, lead to enhanced colonic motility (Fig. 4.37). We discussed in
Chap. 3 how ACh and NO participate in causing peristalsis.
The propagation velocity generated by the MMC is reduced by more than 50%
during sleep by the suppression of phase 2. The spread of contractions decreases in
proportion to the depth of sleep, and is absent during N3 sleep. The frequency of the
spread and the pressure in the colon during REM sleep is like those found in NREM
N2 sleep. It can be concluded then that the MMCs are the basic rhythmic pattern of
intestinal motility, of endogenous origin, with meals having an effect on them (food
entrained oscillators). In addition, MMCs also appear to be associated with changes
in secretory phenomena such as gastric, pancreatic, and biliary secretion.
168 4  Second Level: The Brainstem

Cytoplasm
PER CRY Night

Afternoon
CLOCK BMAL 1
Morning
(CACGTG)
Nucleus

Myenteric
plexus
Acetylcholine
Nitric oxide

Fig. 4.37  Circadian regulation of colonic motility. The rhythmic expression of clock genes within
the neurons of the myenteric plexus modulate colonic motility through clock-controlled transcrip-
tion of genes such as acetylcholine (ACh) transferase and neuronal nitric oxide synthase (nNOS).
Transcription of ACh and nNOS leads to the rhythmic release of ACh and nitric oxide (NO),
which initiates diverse biochemical, cellular, and physiological processes within the colonic
­circular muscle. The figure was prepared in part using image vectors from Servier Medical Art
(www.servier.com), licensed under the Creative Commons Attribution 3.0 Unported License
(http://creativecommons.org/license/by/3.0/)

In humans, the MMCs are interrupted 10–20 min after food intake, staying in a
kind of permanent phase III. The duration of this interruption appears to be propor-
tional to the caloric content of the food and its chemical composition, the most pro-
longed interruption with fat and sugar-rich foods, and less with protein-rich foods.
The mediators of this disruption are fundamentally hormonal, as hormones released
during the digestive process (secretin, gastrin, CCK, pancreatic polypeptide) can
inhibit the MMC cycle. On the contrary, the interruption is not affected by vagotomy,
splenectomy, or celiac and SCGx or a combination of procedures. Differing from
enteral feeding, parenteral nutrition does not interrupt the MMC cycle, indicating
that it is the presence of nutrients directly into the digestive tract that interrupts this
process. Endocrine activity of the small intestine is periodic on some regulatory pep-
tides, and is manifested by the existence of cycles in the release of pancreatic poly-
peptide, motilin, somatostatin, and secretin in blood circulation [55, 58].
24-h Rhythms in Gastrointestinal Function 169

The MMC plays an important role in maintaining the intestinal mucosa, provid-
ing adequate evacuation of debris and ensuring a cephalocaudal flow of the intesti-
nal content. The association of cycles in motility with increases in gastric acid and
pepsin in the stomach and bicarbonate secretion, amylase, bile acids, trypsin, and
lipase in the exocrine pancreas, and bile, promote antibacterial action and cleaning
of the digestive tract. In fact, when the MMC is disturbed, pathological changes
associated with bacterial infections, myotonic dystrophy, intestinal pseudo-­
obstruction, etc., usually occur.
Pancreatic juice originates primarily in the acini of the exocrine pancreas and is
modified by ductus cells before being secreted at the duodenum with a rich content
of bicarbonate and digestive enzymes. Pancreatic juice has an alkaline pH because
of its high bicarbonate content, and neutralizes the acid gastric juices from the stom-
ach, raising the duodenal pH to values of 6–7. In addition, pancreatic juice is rich in
digestive enzymes: trypsin, chymotrypsin, elastase, carboxypeptidase, lipase, and
amylase. Pancreatic secretion presents both ultradian and circadian rhythms.
Ultradian rhythms are in phase with MMCs originating in the duodenum. There is a
direct relationship between the rate of secretion of pancreatic juice and intestinal
changes in contractility at each phase of the MMC: during phase I, resting, the basal
secretion rate remains low; during phase II, pancreatic secretion increases in syn-
chrony with muscle contractions; during phase III, an intense muscle contraction,
pancreatic secretion reaches maximum values. There is also a large overlap between
the periods of peak pancreatic secreting and gastric pH changes.
In contrast to what happens with the ultradian rhythms, circadian oscillations in
pancreatic secretion are of small amplitude. Overall maximum pancreatic juice
secretion occurs overnight, and in this period the greatest amount of amylase is also
produced.
In the case of bile secretion in the duodenum, gallbladder emptying shows, like
the gastric and pancreatic secretions, both ultradian and circadian rhythms. Ultradian
rhythms are like those discussed above and in fasting individuals they match the
duodenal MMC. Peak secretion of bile precedes maximum duodenal motility and
pancreatic secretion by about 30–40 min.
The gallbladder also shows a circadian rhythm in the rate of secretion of bile, with
a maximum during the early hours of the morning, between 07:00 and 09:00 h, coin-
ciding with the maximal rate of formation of bile salts by the liver and in the activity
of the key enzymes in the bile formation, e.g., cholesterol 7-hydroxylase. Thus, the
increased rate of synthesis of bile salts coincides with the release of bile into the
duodenum; it has been postulated that the increased activity of the 7a-hydroxylase
would be initiated by the decrease in the content of bile salts in the liver.
Cholesterol concentrations in bile secretion and those of other lipids follow a cir-
cadian rhythm, with high nocturnal values. The index of cholesterol saturation (litho-
genic index, as calculated from the concentrations of cholesterol, phospholipids, and
bile salts) also shows daily fluctuations, with higher values in the early morning than
in the afternoon.
Several experimental studies have shown that a large number of digestive
enzymes present biological rhythms. An example is the intestinal disaccharidases,
170 4  Second Level: The Brainstem

which have a synchronization with the rhythm of food, but not with the light–dark
periods. Regardless of the changes that may occur in enzyme activity as a function
of time, absorption processes can also be modulated rhythmically. The expression
of glucose transporters shows circadian rhythms. In rats fed ad libitum, glucose
absorption is low during the day and high overnight. This rhythmic pattern is inde-
pendent of the darkness and light cycle and is synchronized by the feeding schedule.
Periodicity is anticipatory and held for a few days of food deprivation and during
discontinuous parenteral nutrition.
A period of REM sleep is associated with increased pressure on the colon and
frequency of contractions at similar levels to N2 sleep. Arousals stimulate an
increase in the spread of contractions in all segments of the colon. Intestinal and
colonic activity are generally decreased during sleep. Rectal motor activity increases
during sleep, but propulsion is retrograde. This and the fact that the anal sphincter
tone (although reduced) remains higher than the rectal tone prevent anal leakage
during sleep [55, 58].
Irritable bowel syndrome is a homeostatic and circadian dysfunction of small
and large intestinal rhythms that entails a constellation of symptoms, including
alternating constipation and diarrhea. It affects 8–25% of the Western population,
with a predominance in women of 2–3 times more than in men. Exaggerated
abdominal pain, altered intestinal function, inflammation of the mucosa, exagger-
ated response to stress, and an increase of proinflammatory cytokines are signs of
irritable bowel syndrome. Stress plays a key role in the triggering and exacerbation
of symptoms [55, 56].
Several studies indicate alterations of the autonomic tone in irritable bowel syn-
drome that results in abnormal central processing of pain with allodynia as a major
sign. Allodynia refers to central pain sensitization following normally nonpainful,
often repetitive, stimulation. Allodynia leads to the triggering of a pain response
from stimuli that do not normally provoke pain. Allodynia is different from hyper-
algesia: an extreme, exaggerated reaction to a stimulus that is normally painful.
Polysomnography studies have indicated that patients with irritable bowel syn-
drome have an increased autonomic sympathetic activity during REM sleep.
Neuroplasticity changes such as the learning of an abnormal visceral response, are
consolidated in REM sleep (procedural memory, Chap. 6). Hence, alterations of the
autonomic tone seen during REM sleep could be consolidated as a memory of
anomalous visceral pain (allodynia).
Irritable bowel syndrome is predominant in people who undertake rotating shift
work. In several studies, logistic regression analysis indicated an association
between this pathology and rotational shift work, even after controlling for other
confounding factors, such as poor sleep quality (chronodisruption). Greater abdom-
inal pain was observed in the groups with rotating shifts, which could be due to the
interruption of the circadian factors that modulate the visceral sensitivity.
Inflammatory bowel disease comprises the chronic inflammation of the large
intestine, and sometimes of the small intestine, whose main forms are Crohn’s dis-
ease and ulcerative colitis. The main difference between the two is the location and
nature of the inflammatory changes. Disruption of the circadian regulation of the
References 171

intestinal immune system is a possible aggravating cause. Circadian clock gene


disruption can affect the intestinal permeability of epithelial cells and the phenom-
enon initiates the inflammatory cascade observed in patients during exacerbation of
the disease [55, 56].
There is evidence that sleep disorders can affect inflammatory processes in the
colon. Both acute and chronic sleep deprivation aggravates colon inflammation in
murine models of colitis.
Concerning sleep, its disturbances coincide with digestive discomfort. NREM
sleep decreases more than REM sleep. During REM sleep, neuronal reprogramming
of the enteric nervous system occurs (neuroplasticity changes) and these changes
can lead to allodynia and visceral hyperalgesia.

References
1. Fernandez-Gil MA, Palacios-Bote R, Leo-Barahona M, Mora-Encinas JP. Anatomy of the
brainstem: a gaze into the stem of life. Semin Ultrasound CT MR. 2010;31:196–219.
2. Tseng CJ, Cheng PW, Tung CS. Chapter 29 – Pharmacology of the nucleous tractus solitarii.
In: Biaggioni I, Burnstock G, Low PA, Paton JFR, editors. Primer on the Autonomic Nervous
System. 3rd ed. San Diego: Academic; 2012. p. 141–4.
3. Cardinali DP. Neurociencia Aplicada. Sus Fundamentos. Buenos Aires: Editorial Médica
Panamericana; 2007.
4. Verriotis M, Chang P, Fitzgerald M, Fabrizi L. The development of the nociceptive brain.
Neuroscience. 2016;338:207–19.
5. Wehrwein EA, Joyner MJ. Chapter 8 – Regulation of blood pressure by the arterial baroreflex
and autonomic nervous system. In: Buijs RM, Swaab D, editors. Handbook of clinical neurol-
ogy. Autonomic nervous system. New York: Elsevier; 2013. p. 89–102.
6. Macefield VG, Henderson LA. Chapter 3 – Imaging of brainstem sites involved in cardio-
vascular control. In: Biaggioni I, Burnstock G, Low PA, Paton JFR, editors. Primer on the
autonomic nervous system. 3rd ed. San Diego: Academic; 2012. p. 13–6.
7. Mischel NA, Subramanian M, Dombrowski MD, Llewellyn-Smith IJ, Mueller PJ. (In)
activity-­
related neuroplasticity in brainstem control of sympathetic outflow: unraveling
underlying molecular, cellular, and anatomical mechanisms. Am J Physiol Heart Circ Physiol
2015;309:H235–43.
8. Gordan R, Gwathmey JK, Xie LH. Autonomic and endocrine control of cardiovascular func-
tion. World J Cardiol. 2015;7:204–14.
9. Wieling W, Groothuis JT. Chapter 39 – Physiology of upright posture. In: Biaggioni I,
Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous system. 3rd ed.
San Diego: Academic; 2012. p. 193–5.
10. Lombard JH, Cowley J. Chapter 38 – Neural control of blood vessels. In: Biaggioni I,
Burnstock G, Low PA, JFR P, editors. Primer on the autonomic nervous system. 3rd ed. San
Diego: Academic; 2012. p. 187–91.
11. Halcox JPJ. Chapter 66 – Endothelial dysfunction. In: Biaggioni I, Burnstock G, Low PA,
Paton JFR, editors. Primer on the autonomic nervous system. 3rd ed. San Diego: Academic;
2012. p. 319–24.
12. Smith PM, Ferguson AV. Recent advances in central cardiovascular control: sex, ROS, gas and
inflammation. F1000Res. 2016;5.
13. Chatham JC, Young ME. Regulation of myocardial metabolism by the cardiomyocyte circa-
dian clock. J Mol Cell Cardiol. 2013;55:139–46.
14. Young ME. The circadian clock within the heart: potential influence on myocardial gene
expression, metabolism, and function. Am J Physiol Heart Circ Physiol. 2006;290:H1–16.
172 4  Second Level: The Brainstem

15. Smolensky MH, Hermida RC, Portaluppi F. Circadian mechanisms of 24-hour blood pressure
regulation and patterning. Sleep Med Rev. 2017;33:4–16.
16. Chouchou F, Desseilles M. Heart rate variability: a tool to explore the sleeping brain? Front
Neurosci. 2014;8:402.
17. Portaluppi F, Tiseo R, Smolensky MH, Hermida RC, Ayala DE, Fabbian F. Circadian rhythms
and cardiovascular health. Sleep Med Rev. 2012;16:151–66.
18. Alibhai FJ, Tsimakouridze EV, Reitz CJ, Pyle WG, Martino TA. Consequences of circadian
and sleep disturbances for the cardiovascular system. Can J Cardiol. 2015;31:860–72.
19. Gilbey MP. Chapter 30 – Entrainment of sympathetic rhythms. In: Biaggioni I, Burnstock G,
Low PA, Paton JFR, editors. Primer on the autonomic nervous system. 3rd ed. San Diego:
Academic; 2012. p. 147–9.
20. Guyenet PG. Regulation of breathing and autonomic outflows by chemoreceptors. Compr
Physiol. 2014;4:1511–62.
21. Costa KM, Accorsi-Mendonca D, Moraes DJ, Machado BH. Evolution and physiology of
neural oxygen sensing. Front Physiol. 2014;5:302.
22. Prabhakar NR. Chapter 68 – Oxygen sensing. In: Biaggioni I, Burnstock G, Low PA, Paton JFR,
editors. Primer on the autonomic nervous system. 3rd ed. San Diego: Academic; 2012. p. 331–3.
23. Barnes PJ. Chapter 41 – Autonomic control of the lower airways. In: Biaggioni I, Burnstock
G, Low PA, Paton JFR, editors. Primer on the autonomic nervous system. 3rd ed. San Diego:
Academic; 2012. p. 201–4.
24. Mazzone SB, Canning BJ. Chapter 18 – Autonomic neural control of the airways. In: Buijs RM,
Swaab D, editors. Handbook of clinical neurology. Autonomic nervous system. New York:
Elsevier; 2013. p. 215–28.
25. Mansukhani MP, Kara T, Caples SM, Somers VK. Chemoreflexes, sleep apnea, and sympa-
thetic dysregulation. Curr Hypertens Rep. 2014;16:476.
26. Harper RM, Bandler R, Spriggs D, Alger JR. Lateralized and widespread brain activation
during transient blood pressure elevation revealed by magnetic resonance imaging. J Comp
Neurol. 2000;417:195–204.
27. Alexander MP, Gillingham S, Schweizer T, Stuss DT. Cognitive impairments due to focal
cerebellar injuries in adults. Cortex. 2012;48:980–90.
28. Xu F, Frazier DT. Role of the cerebellar deep nuclei in respiratory modulation. Cerebellum.
2002;1:35–40.
29. Strata P. The emotional cerebellum. Cerebellum. 2015;14:570–7.
30. Zhu JN, Yung WH, Kwok-Chong CB, Chan YS, Wang JJ. The cerebellar-hypothalamic cir-
cuits: potential pathways underlying cerebellar involvement in somatic-visceral integration.
Brain Res Rev. 2006;52:93–106.
31. Li B, Zhu JN, Wang JJ. Histaminergic afferent system in the cerebellum: structure and func-
tion. Cerebellum Ataxias. 2014;1:5.
32. Li B, Zhuang QX, Gao HR, Wang JJ, Zhu JN. Medial cerebellar nucleus projects to feeding-­
related neurons in the ventromedial hypothalamic nucleus in rats. Brain Struct Funct.
2017;222:957–71.
33. Zhu JN, Wang JJ. The cerebellum in feeding control: possible function and mechanism. Cell
Mol Neurobiol. 2008;28:469–78.
34. Ciriello J, Caverson MM. Hypothalamic orexin-A (hypocretin-1) neuronal projections to the
vestibular complex and cerebellum in the rat. Brain Res. 2014, 1579:20–34.
35. Lu JH, Wang XQ, Huang Y, Qiu YH, Peng YP. GABAergic neurons in cerebellar interposed
nucleus modulate cellular and humoral immunity via hypothalamic and sympathetic pathways.
J Neuroimmunol. 2015;283:30–8.
36. Hanoun M, Maryanovich M, Arnal-Estape A, Frenette PS. Neural regulation of hematopoiesis,
inflammation, and cancer. Neuron. 2015;86:360–73.
37. Kox M, Pickkers P. Modulation of the innate immune response through the vagus nerve.
Nephron. 2015;131:79–84.
38. Sankowski R, Mader S, Valdes-Ferrer SI. Systemic inflammation and the brain: novel roles
of genetic, molecular, and environmental cues as drivers of neurodegeneration. Front Cell
Neurosci. 2015;9:28.
References 173

39. Fernandez R, Nardocci G, Navarro C, Reyes EP, Acuna-Castillo C, Cortes PP. Neural reflex
regulation of systemic inflammation: potential new targets for sepsis therapy. Front Physiol.
2014;5:489.
40. Boeckxstaens G. Chapter 11 – The clinical importance of the anti-inflammatory vagovagal
reflex. In: Buijs RM, Swaab D, editors. Handbook of clinical neurology. Autonomic nervous
system. New York: Elsevier; 2013. p. 119–34.
41. Parekh PJ, Nayi VR, Johnson DA, Vinik AI. The role of gut microflora and the cholinergic
anti-inflammatory neuroendocrine system in diabetes mellitus. Front Endocrinol (Lausanne).
2016;7:55.
42. Leposavic G, Pilipovic I, Perisic M. Cellular and nerve fibre catecholaminergic thymic net-
work: steroid hormone dependent activity. Physiol Res. 2011;60(Suppl 1):S71–82.
43. Cardinali DP, Esquifino AI. Neuroimmunoendocrinology of the cervical autonomic nervous
system. Biomed Rev. 1998;9:47–59.
44. Kawashima K, Fujii T, Moriwaki Y, Misawa H. Critical roles of acetylcholine and the mus-
carinic and nicotinic acetylcholine receptors in the regulation of immune function. Life Sci.
2012;91:1027–32.
45. Nobis CC, Labrecque N, Cermakian N. Circadian control of antigen-specific T cell responses.
ChronoPhysiol Ther. 2016;6:65–74.
46. Besedovsky L, Dimitrov S, Born J, Lange T. Nocturnal sleep uniformly reduces numbers of
different T-cell subsets in the blood of healthy men. Am J Physiol Regul Integr Comp Physiol.
2016;311:R637–42.
47. Cardinali DP, Esquifino AI. Sleep and the immune system. Curr Immunol Rev. 2012;8:50–62.
48. Cardinali DP. Ma Vie en Noir. Fifty years with melatonin and the stone of madness. Switzerland:
Springer; 2016.
49. Geiger SS, Fagundes CT, Siegel RM. Chrono-immunology: progress and challenges in under-
standing links between the circadian and immune systems. Immunology. 2015;146:349–58.
50. Bollinger T, Bollinger A, Oster H, Solbach W. Sleep, immunity, and circadian clocks: a mech-
anistic model. Gerontology. 2010;56:574–80.
51. Irwin MR. Why sleep is important for health: a psychoneuroimmunology perspective. Annu
Rev Psychol. 2015;66:143–72.
52. Dimitrov S, Besedovsky L, Born J, Lange T. Differential acute effects of sleep on sponta-
neous and stimulated production of tumor necrosis factor in men. Brain Behav Immun.
2015;47:201–10.
53. Born J, Lange T, Hansen K, Molle M, Fehm HL. Effects of sleep and circadian rhythm on
human circulating immune cells. J Immunol. 1997;158:4454–64.
54. Faraut B, Nakib S, Drogou C, Elbaz M, Sauvet F, De Bandt JP, Leger D. Napping reverses the
salivary interleukin-6 and urinary norepinephrine changes induced by sleep restriction. J Clin
Endocrinol Metab. 2015;100:E416–26.
55. Kibble JD, Halsey CR. Gastrointestinal physiology. In: Medical physiology: the big picture.
New York: McGraw-Hill Education; 2015.
56. Hasler WL, Owyang C. Approach to the patient with gastrointestinal disease. In: Kasper D,
Fauci A, Hauser S, Longo D, Jameson JL, Loscalzo J, editors. Harrison’s principles of internal
medicine. 19th ed. New York: McGraw-Hill Education; 2015.
57. Mills JC, Stappenbeck TS. Gastrointestinal disease. In: Hammer GD, McPhee SJ, editors.
Pathophysiology of disease: an introduction to clinical medicine. 7th ed. New York: McGraw-­
Hill Education; 2013.
58. Barrett KE. Gastrointestinal physiology. 2nd ed. New York: McGraw-Hill Education; 2014.
59. Kleitman N. Basic rest-activity cycle--22 years later. Sleep. 1982;5:311–7.
60. Chen CQ, Fichna J, Bashashati M, Li YY, Storr M. Distribution, function and physiological
role of melatonin in the lower gut. World J Gastroenterol. 2011;17:3888–98.
61. Bubenik GA. Thirty four years since the discovery of gastrointestinal melatonin. J Physiol
Pharmacol. 2008;59(Suppl 2):33–51.
62. Hoogerwerf WA. Role of clock genes in gastrointestinal motility. Am J Physiol Gastrointest
Liver Physiol. 2010;299:G549–55.
Third Level: The Hypothalamus
5

Abstract
The hypothalamus is a phylogenetically ancient part of the CNS, and its structure
has remained relatively constant in terrestrial vertebrates throughout evolution.
The hypothalamus participates as the third level of the ANS hierarchy in the
coordination of specific behaviors (defense behavior, thermoregulation, feeding,
sexual and maternal behavior). This Chapter discusses the hypothalamic mecha-
nisms that control hormone secretion and how they vary during a 24-h cycle. The
nature of the defense behavior, feeding behavior and temperature regulation is
described.

Keywords
Allostasis • Body temperature regulation • Brown adipose tissue • Constitutive
and reactive secretion of hormones • Defense behavior • Feeding behavior
• Hydroelectrolytic balance • Hypothalamic nuclei and areas • Hypophysiotropic
area • Sexual and maternal behavior

Objectives
After studying this chapter, you should be able to:
• Name the areas and principal nuclei of the hypothalamus and describe how
the hypothalamus participates as the third level of the ANS hierarchy in the
coordination of specific behaviors
• Describe the neural mechanisms that control the constitutive and reactive
secretion of hypophysiotropic, adenohypophyseal, pineal, thyroid, and
parathyroid hormones, and how they vary in the three body configurations
(wakefulness, NREM sleep, REM sleep) during a 24-h cycle.
• Describe the nature of the defense behavior as a paradigm of reactive
homeostasis and differentiate homeostasis from allostasis.

© Springer International Publishing AG 2018 175


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1_5
176 5  Third Level: The Hypothalamus

• Identify the principal physiological components of the feeding behavior


and the location and function of the participating neurotransmitters and
hormones. Describe how they vary in the three body configurations (wake-
fulness, NREM sleep, REM sleep) during a 24-h cycle.
• Describe the neural and hormonal mechanisms that control 24-h rhythms
in plasma osmolality and the intravascular volume by modifying water and
electrolyte intake behavior.
• Understand how the SON–PVN–neurohypophyseal system, the renin–
angiotensin system, and the secretion of ANP vary in the three body con-
figurations (wakefulness, NREM sleep, REM sleep) during a 24-h cycle.
• Identify the homeostatic components of body temperature regulation and
the location and function of the participating neural and hormonal mecha-
nisms of shivering and nonshivering thermogenesis in the three body con-
figurations (wakefulness, NREM sleep, REM sleep) during a 24-h cycle.
• Name the hypothalamic basis of sexual and maternal behavior

 ypothalamic Behaviors Comprise Coordinated Mechanisms,


H
Including Autonomic, Neuroendocrine, and Motivational
Components

The hypothalamus is a phylogenetically ancient part of the CNS, and its structure
has remained relatively constant in terrestrial vertebrates throughout evolution,
unlike other brain regions, such as the neocortex or limbic system, which show
marked interspecies differences.
The hypothalamus is the main governing center for homeostatic functions. A decer-
ebrate animal, in which the brainstem is cut between the upper and lower colliculi,
keeps the regulation of its internal environment intact, as the cut leaves the hypothala-
mus intact. In contrast, an animal with hypothalamic lesions requires extreme care to
survive, as its homeostatic functions are greatly diminished or suppressed [1].
The hypothalamus is organized to perform autonomic, endocrine, and somatic func-
tions. To do this, the region connects with the various components of the autonomic
motor hierarchy, and with the somatosensory, motor, endocrine, and immune systems.
The principal hypothalamic nuclei are depicted in Figs. 5.1 and 5.2. In humans,
the hypothalamus weighs about 5–8 g. Its anatomical limits are rather diffuse. As a
ventral part of the diencephalon, the hypothalamus borders the third ventricle, below
the thalamus. Caudally, it limits with the mesencephalon and rostrally, with the lam-
ina terminalis, anterior commissure, and optic chiasm. Lateral to the hypothalamus
the optic tract, the internal capsule, and the subthalamic structures are found.
From a functional point of view, the hypothalamus should be considered part of a
continuous neuronal component that extends from the midbrain to the basal regions
of the telencephalon (limbic cortex). Together, this portion of the brain is anatomi-
cally linked to the old olfactory, limbic system by the medial forebrain bundle.
Hypothalamic Behaviors 177

Posterior hypothalamic nucleus Dorsal hypothalamic area

Dorsomedial nucleus Paraventricular nucleus


Ventromedial nucleus Anterior hypothalamic
area
Premammillary nucleus
Preoptic area

Medial mammillary nucleus Supraoptic nucleus

Lateral mammillary nucleus


Suprachiasmatic
nucleus

Mammillary body Arcuate nucleus

Primary capillary Optic chiasma


plexus
Median eminence
Superior hypophyseal
artery

Portal vessel
Posterior pituitary

Anterior pituitary

Fig. 5.1  Principal hypothalamic nuclei. Modified with permission from Cardinali [1]

The hypothalamus comprises three zones (Fig. 5.2):

• The periventricular area is a thin layer of nerve tissue adjacent to the third
ventricle.
• In the medial zone, there are several nuclei. The hypophysiotropic area is the
portion of the medial zone participating in adenohypophyseal regulation. In the
supraoptic nucleus (SON) and paraventricular nucleus (PVN), the perikarya of
the fibers forming the hypothalamic–neurohypophyseal tract and secreting argi-
nine vasopressin (AVP) and oxytocin are located.
• The lateral hypothalamic area (LHA) contains no distinguishable nuclei, but
instead Golgi type II neurons, which surround the medial forebrain bundle. This
bundle is continued rostrally with the basolateral structures of the limbic system
and caudally with the rostral structures of the brainstem. The medial forebrain
bundle is one of the largest CNS fiber bundles, which interconnects forebrain
structures with brainstem structures, forming an extensive network profile.

The LHA is an area of greatest interconnectivity of the hypothalamus, allowing


the modulation of various functions, from cognitive to autonomic. The LHA, also
called interstitial nucleus of the medial forebrain bundle, is one of the regions of the
CNS with maximal biochemical and cellular heterogeneity (more than 35 types of
neurons described) [2].
The afferent and efferent connections of the hypothalamus indicate that it is
an important center of integration for autonomic, somatic, and neuroendocrine
178 5  Third Level: The Hypothalamus

Anterior commissure

PVN
Mammillary body
MPOA DMN

AHA
VMN
SCN
SON
Optic chiasm
ARC Pituitary
ME gland

Thrid ventricle Third ventricle

LHA LHA DMN


PVN
SCN VMN
SON
ARC

Optic
tract
Optic ME
chiasm

Fig. 5.2  Sagittal view of the hypothalamus and at the level of the indicated cuts (optic chiasm and
median eminence, ME). MPOA medial preoptic area, PVN paraventricular nucleus, DMN dorso-
medial nucleus, AHA anterior hypothalamic area, VMN ventromedial nucleus, SCN suprachias-
matic nucleus, SON supraoptic nucleus, ARC arcuate nucleus, LHA lateral hypothalamic area

functions. In general, LHA is reciprocally connected with the upper portion of the
brainstem and upper limbic structures. It also receives somatic, inter- and extero-
ceptive impulses through the thalamus, cerebellum, and limbic system.
The medial hypothalamus has abundant reciprocal connections to the lateral
hypothalamus, but receives few projections from other brain areas. Its function is
mainly neuroendocrine. It contains receptors for humoral signals from the internal
environment (glucose and other metabolites, temperature, osmolality, various hor-
mones), its efferences being neuroendocrine (hypophysiotropic and neurohypophy-
seal peptides).
From a functional point of view, the hypothalamus is the level of the autonomic
hierarchy that provides the complex program of the various homeostatic reactions,
with their autonomic, neuroendocrine, and behavioral components. To achieve this,
the hypothalamus uses the various elementary, segmental (medullary level) or system
(brainstem level) programs, contained at lower levels of the autonomic hierarchy.
Hypothalamic Behaviors 179

Third ventricle Mammillary


Optic bodies
chiasm 1 4

Primary capillary
internal
plexus
carotid
artery
Superior
hypophyseal
artery
5
Inferior
hypophyseal
Long
2 artery
portal
vessels

Short
portal
3 vessels
3

to cavernous sinus

Fig. 5.3  Neuroendocrine connections of the medial hypothalamus. Hypothalamic hypophysiotro-


pic hormones ① reach the adenohypophysis through the pituitary portal system ② ③. The hypotha-
lamic–neurohypophyseal system is neurally connected to the neurohypophysis ④ ⑤

There are four hypothalamic functions: (a) neuroendocrine function; (b) regula-
tion of the ANS; (c) hypothalamic behavior regulation; (d) control of biological
rhythms [1].
The neuroendocrine function is mediated by the SON/PVN neurohypophyseal
systems and by the hypophysiotropic area (Fig. 5.3).
Regarding autonomic regulation, the electrical stimulation of almost all hypotha-
lamic regions produces complex ANS responses (cardiovascular, respiratory, diges-
tive, piloerection, etc.). These results are the basis for the hypothesis that the
complex motor programs of the autonomic responses are contained in the hypo-
thalamus, which are conveyed through the sympathetic and parasympathetic centers
of the brainstem and spinal cord [3].
Let us take the cardiovascular function as an example. We discussed in Chap. 4 the
cardiovascular servocontrol system (BP, cardiac output, vasomotor) that resides in the
brainstem and acts through the autonomic segmental reactions in the spinal cord. The
efferent components of this servocontrol system are the vagus nerve and sympathetic
descending pathways to the intermediolateral columns of the spinal cord.
180 5  Third Level: The Hypothalamus

The afferences come from baro- and chemoreceptors, and atrial and ventricular car-
diac mechanoreceptors [4]. This level of regulation of the brainstem, self-­sufficient
for exclusively vascular reflexes, depends on complex stereotyped responses involv-
ing several systems under hypothalamic control. This control is exerted both by
neural connections between the hypothalamus and the brainstem centers (in particu-
lar, the nucleus of the solitary tract, NTS) and by descending hypothalamic projec-
tions of the dorsal longitudinal fasciculus type [5]. The cardiovascular system is
under hypothalamic control in all responses involving greater complexity than sim-
ple brainstem servomechanisms (e.g., cardiovascular changes during thermoregula-
tion, food intake, defensive behavior) [6, 7].
The involvement of the hypothalamus in the regulation of various behaviors is
revealed by the variety of responses triggered by the electrical stimulation of hypo-
thalamic areas. Hypothalamic behaviors comprise the coordinated manifestation of
neurovegetative, neuroendocrine, somatic, and motivational mechanisms. The main
behaviors coordinated by the hypothalamus are: (a) defense behavior; (b) nutritive
or appetitive behavior; (c) thermoregulatory behavior; (d) sexual and maternal
behavior. The hypothalamus is also the site of integration of environmental and
endogenous signals that determine the different biological rhythms (Chap. 2) [1].

24-h Rhythms in Neuroendocrine Function

The reactive and constitutive aspects of homeostasis (Chap. 1) are clearly mani-
fested by studies on hormonal secretion. A variety of stimuli trigger the reactive (or
facultative) secretion of a hormone, whereas periodic changes are the manifesta-
tion of the constitutive secretion (predictive homeostasis) linked to biological
rhythms [8].
A broad spectrum of periodic processes characterizes the endocrine system,
including infradian rhythms, whose length ranges from months to years (like the
menstrual cycle or seasonal rhythms in reproduction), circadian 24-h rhythms, and
secretions that occur with intervals of hours (ultradian rhythms). In many cases,
ultrafast or high-frequency variations, with periods of 5–15 min, are superimposed
onto the hourly variations, caused by pulsatile variations or episodic secretions of
pituitary hormones.
The 24-h temporal pattern of adenohypophyseal hormone release is controlled
by the circadian system (C process) and the sleep homeostat (S process) as dis-
cussed in Chap. 2 (Fig. 5.4). The daily rhythmicity of the pituitary hormones occurs
mainly by the circadian modulation of the amplitude of the secretory pulses,
whereas the sleep/wake homeostat exerts its influence mainly on the frequency of
pulse secretion.
To discern the relative roles of the circadian pacemaker and the effects of sleep
on daily hormonal variation, strategies are based on the notion that the circadian
rhythmicity needs several days to adapt to a sudden change in the sleep/wake cycle.
Studies conducted in individuals studied under normal sleep–wake cycles cannot
establish whether the rhythm is circadian or is driven by external factors such as
24-h Rhythms in Neuroendocrine Function 181

Sleep
(S process)

Ultradian Pulse
Generator
(KNDy neurons)

(frequency)

)
de
itu
pl
m
(a
GH
PRL
ACTH
TSH
FSH
Circadian LH
Clock
(C process)

Fig. 5.4  The 24-h temporal pattern of adenohypophyseal hormone release is controlled by the
circadian system (C process) and the sleep homeostat (S process). The daily rhythmicity of the
pituitary hormones occurs mainly by the circadian modulation of the amplitude of the secretory
pulses, whereas the sleep/wake homeostat exerts its influence mainly on the frequency of pulse
secretion. KNDy neurons expressing kisspeptin, neurokinin B, and dynorphin. The figure was pre-
pared in part using image vectors from Servier Medical Art (www.servier.com), licensed under the
Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

sleep–wake. A “gold standard” to disclose between the two is the constant routine
protocol. Participants remain awake in bed for up to 50 h in a semirecumbent pos-
ture under dim light conditions and are fed hourly isocaloric meals.
A less precise but more practicable experimental design to discern the type of
control of the constitutive secretion of a hormone is blood sampling (every 20 min)
during normal sleep (day 1), night deprivation (day 2), and recovery of lost sleep
from 11 am on at day 3 (Fig. 5.5) [9, 10]. By means of this strategy, it was verified
that cortisol shows its maximum level at the expected time (the last part of the night
and early morning) regardless of whether sleep was allowed or not (Fig. 5.5, upper
panel), whereas growth hormone (GH) is secreted whenever N3 NREM sleep is
detected, i.e., in the first part of the night of day 1 or during the recovery nap on day
2 (Fig. 5.5, lower panel).
For hormones controlled by the hypothalamic-pituitary axis, the pulsatile varia-
tion of their plasma levels is caused by intermittent discharges of an ultradian pulse
generator (Fig. 5.4). Gonadotropin pulses were the result of gonadotropin-releasing
182 5  Third Level: The Hypothalamus

DAY 1 DAY 2
30

25
Cortisol (µg/dL plasma)

20

15

10

5
Sleep Sleep Diurnal
deprivation sleep
30

25
GH (µg/dL plasma)

20

15

10

5
18 22 02 06 10 14 18 22 02 06 10 14 18
Time of day (h)

Fig. 5.5  Schematic representation of a daily hormonal rhythm controlled by the circadian clock
or C process (cortisol) or by the S process (growth hormone, GH). The subject was deprived of
sleep for one night and could recover the lost sleep from 11 am the next day. Cortisol is secreted in
the phase established by the circadian clock preceding wakefulness, in the presence or absence of
sleep. GH is secreted whenever N3 sleep occurs. Data from Van Cauter and Refetoff and Copinschi
et al. [9, 10]. Ultradian variations in hormone secretion and values <5 μg/dL in both hormones are
not represented

hormone (GnRH) pulses originating from synchronous discharges of GnRH-­


producing neurons, controlled by a pacemaker located in the anteroventral periven-
tricular nucleus (AVPV) and the arcuate nucleus (ARC; Fig. 5.6) [11]. In primates
with hypothalamic lesions, normal levels of gonadotropins are only restored by pul-
satile, but not continuous, administration of GnRH. Rather, the continuous adminis-
tration of GnRH inhibits gonadotropin release. These findings have been applied to
the treatment of a wide variety of disorders of the pituitary–gonadal axis, using the
pulsating administration of GnRH to correct deficient production of endogenous
GnRH, and the continuous administration of GnRH analogs to inhibit pituitary
gonadotropin synthesis [12].
The concept of the hypothalamic GnRH pulse generator emerged in the 1980s,
but remained undefined for more than 20 years. The discovery of a hypogonado-
tropic hypogonadism associated with loss of function of GPR54, the receptor for a
24-h Rhythms in Neuroendocrine Function 183

AVPV

KNDy
neurons AVP SCN
AVP
DMH
GnIH

VIP

ARC
KNDy neurons
GnRH
MPOA

Fig. 5.6 Hypothalamic nuclei involved in the regulation of gonadotropin release. The


gonadotropin-­releasing hormone (GnRH) neurons of the MPOA project to the ME. Modulatory
inputs upstream of the GnRH neurons include neurons that express kisspeptin (Kiss1; stimulatory
of GnRH release), neurokinin B (stimulatory of GnRH release), and dynorphin (inhibitory of
GnRH release; KNDy neurons). They are in the anteroventral periventricular nucleus (AVPV) and
ARC. A major inhibitory signal derives from gonadotropin-inhibiting hormone (GnIH) releasing
neurons of the dorsomedial hypothalamus (DMH). The SCN influences multiple sites involved in
the control of gonadotropin release via monosynaptic projections from the SCN shell to the posi-
tive (KNDy and GnRH) and the negative signal of GnRH release (GnIH). The figure was prepared
in part using image vectors from Servier Medical Art (www.servier.com), licensed under the
Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

54-amino acid peptide called kisspeptin (Kiss1), which is a very potent GnRH
secretagogue, changed the field dramatically. Kiss1 is critical for the onset of
puberty, the regulation of sex steroid-mediated feedback, and the control of adult
fertility. Further studies indicated that other two peptides, neurokinin B (stimulatory
of GnRH release) and dynorphin (inhibitory of GnRH release) are co-­expressed in
a subset of neurons in the AVPV and ARC (Fig. 5.6). The acronym, KNDy, was
coined to describe these neurons. Current knowledge holds that the hypothalamic
timing mechanism is initiated in the KNDy neuronal network of the nucleus by the
reciprocating interplay of stimulatory kisspeptin/neurokinin B signals and inhibi-
tory dynorphin one (Fig. 5.6) [13]. Kisspeptin neuron activity oscillates on a circa-
dian basis, these neurons expressing clock genes that regulate their rhythmic
activities [14].
The daily rhythmicity of the pituitary hormones occurs mainly by the circadian
modulation of the secretory pulse amplitude, whereas the sleep/wake homeostat
184 5  Third Level: The Hypothalamus

DAY 1 DAY 2
300
Plasma PRL (% of mean)

250

200

150

100
Sleep Sleep Diurnal
deprivation sleep
6

5
TSH (µU/mL plasma)

2
18 22 02 06 10 14 18 22 02 06 10 14 18

Time of day (h)

Fig. 5.7  Schematic representation of the daily rhythm of prolactin (PRL; upper panel) controlled by
the S process, and that of thyroid-stimulating hormone (TSH), under the double control of C and S
processes (lower panel). The volunteer was deprived of sleep for one night and could recover the lost
sleep from 11 am the next day. Ultradian variations in hormone secretion or values <100% of mean
plasma PRL or <2 μU of TSH/mL plasma were not represented. Data from Copinschi et al. [10]

exerts its influence mainly on the frequency of secretion of these pulses (Fig. 5.4).
As stated, sleep dependence on rhythm is indicated because the effects of sleep
occur independently of the time of day when the waking to sleep transition occurs.
The persistence of rhythms in the absence of periods of sleep revealed the depen-
dence of the circadian pacemaker on the generation of these hormonal rhythms.
Constitutive GH (Fig. 5.5) and prolactin (PRL; Fig. 5.7) secretion is controlled
by the sleep/wake homeostat. In normal adult subjects, the daily profiles of GH in
plasma are characterized by stable low levels interrupted abruptly by peaks of secre-
tion [10]. Thus, GH release is pulsatile throughout the day, although in adults the
most frequent secretory pulses (approximately 75% of pulses) occur in the early
hours of sleep associated with slow-wave sleep (N3), with very low secretory activ-
ity during REM sleep.
The episodic secretion of GH depends on the specific rhythms of growth
hormone-­ releasing hormone (GHRH) and on the hypothalamic GH inhibitory
24-h Rhythms in Neuroendocrine Function 185

hormone, somatostatin, whose releases are 180 degrees out of phase. Variations in
the secretion of somatostatin explain different sensitivities to GHRH during sleep
[15]. Thus, stage N3 sleep is associated with low levels of hypothalamic somatosta-
tin and REM sleep with high levels. Somatostatin secretion fixes the time (frequency
and duration) of the release of GH, whereas the secretion of GHRH determines its
magnitude. Humans are the only species that shows a close relationship between
GH secretion and sleep. It has been suggested that the association between sleep
and GH secretion in humans is the result of the consolidation of the sleep–wake
cycle [16, 17].
The total amount of GH secreted daily is closely related to the individual’s age.
In the prepubertal phase, GH pulses occur predominantly between 22:00 and 04:00
h, whereas at puberty the GH pulses are larger in number and amplitude, distribut-
ing throughout the 24-h period. Prepubertal children have daily mean values of GH
concentration like those post-puberty and in adult life, whereas at the end of puberty,
the total amount of GH secreted daily reaches its maximum value. As age increases,
nocturnal GH pulses decrease in frequency and amplitude, with no appreciable
changes in plasma GH concentration during the day [10].
In normal young adult males, most of the daily GH secretion (>70%) occurs
shortly after sleep onset during NREM N3 sleep. In young women, there are also
pulses during the day that are more frequent and of greater amplitude than the men,
related to the estrogens. From the age of 50–60 years, there is no constitutive release
of GH, which is responsible for the loss of lean muscle mass (about 2 kg per decade);
this coincides with the disappearance of N3 sleep. The major primary alteration that
accounts for GH hyposecretion in old age is the increased secretion of hypothalamic
somatostatin. There is also a decrease in the effectiveness of GHRH at releasing
GH. Women generally secrete more GH than men of the same age and in women the
effect of aging on GH secretion correlates with the decrease in circulating estradiol
concentration [17].
Under normal conditions, the daily profile of PRL plasma levels follows a pat-
tern with minimal morning concentrations, a gradual increase in the afternoon, and
a higher nocturnal elevation that begins after the onset of sleep and culminates after
around half of it [18]. This is outlined in Fig. 5.7, top panel. In addition, over 24 h,
there are episodic pulses of PRL that can range from 7 to 24 throughout the day.
Modulation of the amplitude of the secretory pulses gives rise to the daily pattern of
hormonal secretion.
The constitutive secretion of PRL is strongly related to sleep homeostasis, as
daytime sleep or naps are associated with an increase in the release of the hormone.
Similarly, studies after reversal of the sleep–wake cycle have shown that sleep onset
is associated with an increase in PRL secretion, whereas sleep deprivation prevents
a nocturnal increase in PRL. However, in changes in the sleep–wake cycle and in
sleep deprivation, it has been shown that the temporary organization of PRL release
also presents some inherent circadian rhythmicity shown by a slight hormonal ele-
vation corresponding to the moment of omitted sleep. PRL secretion plays a poten-
tial role in circadian REM sleep and N3 sleep control, which increases in patients
with hyperprolactinemia and in lactating women [17].
186 5  Third Level: The Hypothalamus

The circadian clock controls the constitutive secretion of cortisol and melatonin.
The periodicity of 24 h in the secretion of the pituitary–adrenal axis is considered a
paradigm of circadian rhythmicity in humans (Fig. 5.4, upper panel). It begins at the
age of 6 months and persists in adults until old age. The daily profiles of plasma
cortisol are parallel and show a 2–3 h delay to those of adrenocorticotropic hormone
(ACTH). This pattern shows a maximum at the beginning of the morning (04:00–
08:00 h), a decrease throughout the day, followed by a low concentration during the
night period (giving rise to a quiescent period from 24:00 at 02:00 h) and an abrupt
rise at the end of the sleep period [19].
During the 24 h, about 15 pulses of cortisol and ACTH occur, which indicates
that the daily variation of cortisol levels reflects a modulation of the amplitude of
the pulses rather than the frequency. That is, the 24-h rhythm of cortisol is mainly
dependent on the circadian pattern of ACTH release, which is amplified by a daily
variation in the adrenal response to it. This depends on the autonomic innervation of
the adrenal, a case like that described for thyroid innervation in Chap. 4. In turn, the
rate of ACTH release is the result of periodic changes in the release of corticotropin-­
releasing hormone (CRH) [19].
The rhythm of secretion of the hypophyseal–adrenal axis persists in elderly sub-
jects, although in general, the elevation in cortisol levels occurs earlier, with reduced
latency to REM sleep. In addition, an increase in free plasma cortisol levels has
been observed in older people, attributed to a decrease in the concentration of trans-
port proteins and/or a decrease in the binding capacity of these proteins [18].
The circadian pacemaker directly controls the daily rhythm of cortisol, as dem-
onstrated by the rapid phase shift observed after exposure to bright light at certain
circadian moments. This rhythmicity of cortisol, which is of an endogenous nature,
persists during sleep deprivation, although a reduction of 10–20% of the amplitude
is observed. Nighttime sleep exerts an inhibitory effect on the secretion of cortisol,
superimposed on its endogenous circadian rhythmicity [20].
In old age, free cortisol levels increase by 20–50% compared with those in young
people, and typically, the minimum (nadir) of the rhythm of cortisol in an individual
over 70 years old is higher than that of a young adult. There is also a phase advance-
ment of the rhythm of cortisol with age [18]. Studies in animals and in human clin-
ics have indicated important neurodegenerative effects of cortisol, especially at
hippocampal levels, which are more pronounced in the nadir of the rhythm than in
its crest. We will discuss them further in this Chapter. Therefore, even modest eleva-
tions of evening cortisol in the elderly facilitate the development of disturbances
associated with excess glucocorticoids, such as memory deficit and insulin resis-
tance associated with old age [18].
The plasma concentration of melatonin has peak nocturnal values ranging from
0.4 to 0.8 pmol/mL and a minimum during the day (0.02 pmol/mL) in humans. In
saliva, its values are 0.6–0.8 pmol/mL at the time of the nocturnal maximum
(Fig. 5.8). In man, peak melatonin secretion is not related to the sleep phase. In the
same way as cortisol, the daily rhythm of melatonin secretion remains in conditions
of sleep deprivation (Fig. 5.8). Owing to its marked intraindividual reproducibility,
the rise in melatonin in the afternoon in dim light (dim light melatonin onset, DLMO)
is the most accurate marker of τ, the circadian rhythm period (Fig. 5.8) [21].
24-h Rhythms in Neuroendocrine Function 187

DAY 1 DAY 2

Sleep Sleep deprivation


0.06
Pmoles of melatonin / mL saliva

0.04

0.03

0.02
DLMO

0.01
14 18 22 02 06 10 14 18 22 02 06 10 14
Time of day (h)

Fig. 5.8  Saliva melatonin levels in a volunteer subjected to a night of sleep deprivation. Saliva
samples were taken every 20 min. Melatonin was assayed by radioimmunoassay. DMLO dim light
melatonin onset. Values of melatonin <0.01 pmol/mL of saliva are not depicted. Cardinali et al.,
unpublished data

The hypothalamic–pituitary–thyroid axis presents a complicated temporal struc-


ture with rhythmic variations of multiple frequencies at all levels of the system,
from the hypothalamic neurons to the cells of the peripheral effector tissues. The
range of frequencies consists from rapid neural discharges to circadian and circan-
nual rhythms, superposing on these rhythmic variations according to age. Rhythmic
and nonrhythmic variations interact, modulate, and are modulated by variations of
other neuroendocrine, metabolic, and immune functions. In normal subjects, thy-
roid-stimulating hormone (TSH) is secreted from the pituitary in a pulsatile form,
demonstrating that the pituitary secretion of TSH responds better to an intermittent
continuous stimulation of thyrotropin-releasing hormone (TRH). The average fre-
quency is nine pulses in 24 h, increasing in amplitude and frequency at night, which
leads to a nocturnal rise in hormone levels [22].
Under normal conditions, TSH levels are low throughout the day, experiencing a
sudden increase at 20:00 h, reaching the maximum levels at night, this maximum
being circadian pacemaker-controlled (Fig. 5.7, lower panel) [23]. High hormone
levels remain throughout the night, corresponding to an increase of approximately
175% of daytime levels, to decrease rapidly in the morning, reaching a nadir in the
early morning hours.
The existence of circadian rhythms of thyroid hormones is debatable. A 12-h
hemicircadian rhythm with morning and evening minima may occur [24]. It has
188 5  Third Level: The Hypothalamus

been suggested that these variations of low amplitude might be related to daily
variations in plasma proteins dependent on posture. As we discussed in Chap. 4,
autonomic innervation plays a major role in the modulation of the thyroid response
to TSH and may be responsible for this dissociation between TSH and thyroid hor-
mone rhythms.
To investigate the role played by sleep in the daily pattern of TSH, sleep depriva-
tion or sleep–wake cycle inversion indicate, contrary to prediction, an inhibitory
influence of sleep [23]. The inhibitory influence of sleep on TSH secretion is evi-
dent during sleep deprivation, more than doubling the increase in nocturnal values
(Fig. 5.7).
The daily variation of PTH is characterized by an increase in the mean nocturnal
levels, with a temporal interrelation between blood concentrations of PTH and cal-
cium. Plasma PTH levels are normally high during the night and early morning and
lowest at approximately 10:00 h [25]. Concerning 1,25-dihydroxycholecalciferol,
results are conflicting on the existence of circadian variations. Small daily fluctua-
tions in calcitonin concentration have been reported in humans with increased cal-
citonin level during the afternoon. Food intake influences serum calcitonin level in
healthy young subjects.
Bone resorption is closely linked to sleep quality. Bone cells exhibit 24-h cycles
with nocturnal expression of genes that regulate osteoblast function, bone mineral-
ization and ossification, and bone resorption markers. Interruption in melatonin-­
mediated signaling is responsible for the increased risk of hip and wrist fracture and
low bone density in shift workers (Fig. 5.9) [21].

darkness darkness

melatonin

bone
resorption

darkness darkness

more bone
resorption

less melatonin
(LAN, aging)

Fig. 5.9  Relationship between melatonin secretion and bone resorption in a 24-h cycle. As the
image shows, both bone resorption (dotted line) and melatonin (solid line) show a daily rhythm,
with peaks occurring during dark hours. Suppression of nocturnal melatonin levels, either by expo-
sure to light at night (LAN) or in aging, increases bone resorption. The restoration of nocturnal
melatonin maximum protects the bone loss
24-h Rhythms in Neuroendocrine Function 189

By means of studies in volunteers in the sleep laboratory to which several neuro-


peptides were administered in pulsatile form, the link between neuropeptides and
the different stages of sleep determined by polysomnography (PSG) could be veri-
fied [26]. During the first half of the night and coincident with slow-wave sleep, the
mechanisms associated with GH secretion predominate, whereas in the second half
of the night there is a coincidence in the percentage increase in REM sleep with the
secretion of the hypothalamic–pituitary–adrenal axis hormones. Based on this, the
homology of GH secretion with the S process and activation of the ACTH/cortisol
axis with the C process has been proposed (Fig. 5.10). Neuropeptide perfusion stud-
ies indicated that: (a) GHRH promotes GH release and slow-wave sleep; (b) CRH
promotes the activation of the pituitary–adrenal axis and REM sleep; (c) galanin

20 CRH
GHRH CRH GHRH

15
Cortisol (µg/dL)

10

0
18 22 02 06 10 14 18 22 02 06 10 14 18 22

GHRH CRH CRH GHRH


20

15
GH (µg/dL)

10

0
18 22 02 06 10 14 18 22 02 06 10 14 18 22

Fig. 5.10  Neuropeptide perfusion studies in healthy volunteers during a PSG show that GHRH
perfusion promotes GH release and NREM sleep, whereas infusion of corticotropin-releasing hor-
mone (CRH) promotes pituitary–adrenal axis activation and REM sleep. A reciprocal interaction
of GHRH and CRH as a key to sleep regulation was proposed. GHRH predominates during the first
half of the night, resulting in slow sleep, GH secretion, and minimal secretion of adrenocortico-
tropic hormone (ACTH) and cortisol, whereas the second half of the night is dominated by CRH
(ACTH and cortisol secretion). These periods are superimposed on the results of an experiment in
which a healthy volunteer was deprived of sleep for one night and could recover the lost sleep from
11 am the next day. Modified with permission from Cardinali [1]
190 5  Third Level: The Hypothalamus

increases slow-wave sleep in the absence of changes in GH and cortisol; (d) soma-
tostatin inhibits GH and slow-wave sleep and promotes REM sleep; (e) NPY is an
endogenous CRH antagonist and contributes to fixing the time of onset of sleep; (f)
ghrelin increases GH, NREM sleep, and cortisol [26].
It should be noted that changes in EEG of sleep following administration of
GHRH and CRH are not due to GH or cortisol secretion, as slow sleep decreases
after GH administration, whereas slow-wave sleep and GH increase after injecting
cortisol. Therefore, the results are best explained by the inhibitory feedback on
GHRH and CRH respectively.
These studies have been of interest in linking normal aging with depression [26].
In both situations, there is deterioration of slow-wave sleep, shortened REM sleep
and increases in REM sleep, in addition to changes in the continuity of sleep (pro-
longed sleep latency, frequent nocturnal awakenings, early morning awakening).
Endocrine changes are similar in both situations: there is an elevation of ACTH and
cortisol and suppression of GH. This is because the GHRH/CRH ratio changes in
favor of CRH during an episode of depression because of CRH hypersecretion,
whereas in aging the GHRH/CRH ratio changes in favor of CRH because of the
reduction of GHRH activity [26].
The rhythmic aspects of the activity of the hypothalamic–pituitary–gonadal axis,
the neurohypophyseal system, and the secretion of hormones associated with energy
homeostasis and blood volume control are discussed below with their specific
behaviors.

Defense Behavior as a Paradigm of Reactive Homeostasis

Nothing in physiology illustrates the concept of reactive homeostasis better than


defense behavior. Hans Selye named it “general adaptation syndrome” to define the
set of changes arising as the body’s response to a wide variety of noxious stimuli
(stressors) [27]. Stressors are those stimuli whose perception by the nervous system
does not match the neural representation of past experiences, and to which a change
in coping strategy (e.g., a particular behavior) is not successful (Fig. 5.11).
In general, the stimuli stressors can be classified into four groups: (a) physical/
chemical stressors (heat, cold, intense radiation, noise, vibration, toxic substances,
etc.); (b) psychological stressors (emotional and behavioral changes, such as anxi-
ety, fear, frustration); (c) social stressors (hostile environment, disruption of rela-
tions); (d) those that alter ANS homeostasis (exercise, orthostatic, body tilt,
hypoglycemia, bleeding, etc.).
An animal that is faced with a threatening situation must decide, among different
strategies, what is apparently the best choice for the conservation of its life and, in
a broad sense, of its species. For example, when sympathetic activation occurs,
cardiac output and systemic blood flow increase and, although vasodilation occurs
to increase the blood supply to active skeletal muscles for posture and intended
movements, vasoconstriction in the skin can reduce blood loss in a region that is
susceptible to damage [28]. A decreased time for blood clotting avoids a large
Defense Behavior as a Paradigm of Reactive Homeostasis 191

Social context, Physical & Psychol-


Social status Stimulus ogical challenges
Behavior: Interpretation and
reaction to the challenge
Learning, Individual
Genetic predisposition,
Processor psychosocial history
Developmental status,
Gender

Risk free Risk

Known source Unknown source


No stress
Stress
Response repertoire

Low cost
Intermediate cost High cost

Helplessness Hypervigilance
Aggressive,
(anxiety) (anxiety)
risky behavior

BIOLOGICAL RESPONSES

Fig. 5.11  Behavioral and biological aspects of the defense behavior. The sequence of phenomena
following exposure to a novel situation is described. Strategies are of two types, high or low energy
level. Their objective is adaptation. If adaptation does not occur, the stress reaction appears, with
stimulation of the pituitary–adrenal axis

hemorrhage, and higher sudoresis makes the skin moistened and the animal difficult
to catch. At the same time, there is enhanced breathing movements and airflow vol-
ume for gaseous exchanges and pH regulation. Liver glycogenolysis is stimulated,
along with the mobilization of fatty acids from adipose tissue [29]. Muscular
strength and muscle glycolysis are also improved and, for additional energy produc-
tion, a higher oxygen–hemoglobin dissociation and oxygen delivery to activated
cells also occur. Wave frequencies in the EEG increase, which reflects diffuse neu-
ronal activation, and mydriasis serves to provide more visual information and to
decide the best behavioral strategy to execute, such as skipping out of a dangerous
place. By the action of α-adrenoceptor receptors, the secretion of insulin is inhibited
(it would be useless to restore energy reserves at this moment), whereas
β-adrenoceptors stimulate glucagon secretion. Cortisol, in addition to other stress-
related hormones, can indirectly affect the consolidation of memories. Cells of the
immunological system, lymphoid tissues, and cytokines are also involved in this
integrated response (Fig. 5.12).
Thus, coping with stressors includes two strategies (Figs. 5.11 and 5.13). An
active coping strategy (fight-or-flight) is evoked if the stress is predictable,
192 5  Third Level: The Hypothalamus

Fig. 5.12  Via ACTH


release and activation of
adrenal sympathetic
nerves, the defense Hypothalamus
behavior is developed.
Modified with permission
from Cardinali [1]
ACTH Pituitary gland

Adrenal gland

Glucorticoids Epinephrine Norepinephrine


Catabolism
Tachycardia, tachypnea
Blood pressure increase
Redistribution of blood flow
Immunosupression

controllable, or escapable. A passive coping strategy (immobility or decreased


responsiveness to the environment) is evoked if the stress is inescapable [1].
The active strategy is associated with sympathoexcitation (hypertension, tachy-
cardia, thermogenesis), whereas the passive strategy is associated with sympathoin-
hibition and/or parasympathetic activation (hypotension, bradycardia). The passive
strategy also helps to facilitate recovery and healing.
The active strategy is called the “fight-or-flight” response from a behavioral
point of view or the “defense response” from an autonomic point of view. The pas-
sive strategy is sometimes called “paradoxical fear” or “playing dead.” Parts of the
neural substrates that mediate active versus passive emotional coping have been
identified within the brainstem [30, 31].
If strategies to new situations (high- or low-energy consumption) are not suc-
cessful, the stress reaction is triggered (Figs. 5.10 and 5.12). The responses corre-
spond to different functional configurations of the limbic system (amygdala
dominance in the fight situation; predominance of septo-hippocampal components
in defeat; Fig. 5.13).
During stress, the physiological processes that do not pose a short-term benefit
are inhibited, such as inflammation, digestion, reproduction, and growth. When the
intensity or duration of stressors exceed certain limits, pathological changes such as
hypertension, gastric ulcers or neurological abnormalities ensue.
The defense capability or adaptation of an organism to stress depends on the
magnitude and duration of the stressor, and on the sensitivity to the stressor. There
are differences in the ways in which individuals face situations of stress. These dif-
ferences are related to genetics, environmental influences, previous experience in
such situations, training, social support, and physical and mental health.
Defense Behavior as a Paradigm of Reactive Homeostasis 193

Stressor

Control STRATEGIES Loss of


Genetic Control
epigenetic

ACTIVITY
(FIGHT OR FLIGHT) PASIVITY

Amygdala Hippocampus,
alert, attempt to overcome septum behavioral
the situation inhibition

Defeat
Defense reaction
Loss of territoriality
Aggresiveness, motility
Sexual and maternal behaviors inhibited
Visceral fat

Sympatho-adrenal system
Corticoadrenal system

Control Fight for Control Loss of Control


NE ↑ E↑ ACTH ↑
Cortisol ↔ PRL ↑ Cortisol ↑
Gonadotropins ↑ β Endorphin ↑ Catecholamines ↔
Testosterone ↑ Renin ↑ Gonadotropins
Oxytocin ↑ FFA ↑ Testosterone ↓
Glycogenolysis ↑ Pepsin ↑

Fig. 5.13  The different neuroendocrine profiles of defense behavior

An important difference is the programming of ANS connections during early


extrauterine life (Fig. 5.14) [32]. The conditions of reaction to stress have clear
individual nuances, arising from fixed initial experiences in a learning phase, during
the early stages of life. This explains why the same stressful situation, which is not
harmful to many individuals, leads some to myocardial infarction, others to peptic
ulcer or ulcerative colitis, and others to hyperthyroidism. That is, the responses of
the ANS depend on the previous history and individual experience.
Stress responses involve a complex neurobehavioral cascade, which is elicited
when the organism is confronted with a potentially harmful stimulus. As this stress
cascade consists of a range of neural and endocrine pathways, stress can be concep-
tualized as a communication process on the descending branch of the brain–body
axis. Therefore, interoception and stress are associated via the bi-directional
194 5  Third Level: The Hypothalamus

Fetus, Adult Individual pattern of:


Newborn Behavior
Cognition
CNS CNS
Memory
Limbic Limbic Learning
development function Emotion
Early
environment

Development Function Individual pattern of:


of HA axis of HA axis corticoid effects:
- cardiovascular
- immune
- metabolism

Fig. 5.14  There are differences in the ways in which individuals face stressful situations. These
differences are related to genetics, environmental influences, previous experience in such situa-
tions, training, social support, and physical and mental health. An important difference is the pro-
gramming of ANS connections during early extrauterine life. Modified with permission from
Cardinali [1]

transmission of information on the brain–body axis. The excessive and/or enduring


activation (e.g., by acute or chronic stress) of neural circuits, which are responsible
for successful communication on the brain–body axis, induces malfunction and
dysregulation of these information processes. Therefore, interoceptive signal pro-
cessing may be altered, resulting in physical symptoms contributing to the develop-
ment and/or maintenance of body-related mental disorders, which are associated
with stress. A positive feedback model involving stress (early life or chronic stress,
and major adverse events), the dysregulation of physiological stress axes, altered
perception of bodily sensations, and the generation of physical symptoms, which
may in turn facilitate stress, have been proposed [33, 34] (Fig. 5.14).
In the event of a challenge to homeostasis (stress), recovery may occur with dis-
appearance of the stressor. For example, in a high salt diet, homeostatic mechanisms
promote the elimination of excess sodium to restore equilibrium. However, even in
situations when the stressor is not eliminated, it is possible to restore and maintain
homeostasis (Fig. 5.15). To achieve this, animals must have to preserve the stability
of their variables through changes of state related to challenging circumstances. In
this sense, the concept of “allostasis” (allo, meaning different, and stasis, meaning
constancy, i.e., “achieving stability through change”) was introduced to take into
account the regulatory systems that develop variable set points of control, showing
individual differences in expression according to the capacity of the animal to cope
with new situations [30]. This is associated with anticipatory behavioral and physi-
ological responses, and is vulnerable to physiological overload and the breakdown
of regulatory capacities.
The allostatic mechanisms can maintain homeostasis in the presence of a
stressor (Figs. 5.15 and 5.16). As important as the mobilization of functional and
Defense Behavior as a Paradigm of Reactive Homeostasis 195

HOMEOSTASIS

Homeostatic mechanisms

Stressor

Stressor Adaptation Allostatic load


to stress
STRESS

HOMEOSTASIS ALLOSTASIS

Fig. 5.15  In the event of a challenge to homeostasis (stress), recovery may occur with disappear-
ance of the stressor (left). However, even in situations when the stressor is not eliminated, it is
possible to restore and maintain homeostasis (right). The concept of “allostasis” was introduced to
consider the regulatory systems that develop variable set points of control, showing individual dif-
ferences in expression according to the capacity of the animal to cope with new situations. The
figure was prepared in part using image vectors from Servier Medical Art (www.servier.com),
licensed under the Creative Commons Attribution 3.0 Unported License (http://creativecommons.
org/license/by/3.0/)

behavioral processes to maintain the homeostatic balance in response to the


stressor, is the appropriate demobilization of these processes when terminating the
stressor stimulus. Thus, both the mobilization and the inadequate demobilization
of allostatic mechanisms (Fig. 5.16) can cause adaptation diseases, such as hyper-
tension, obesity, diabetes, stroke, autoimmune diseases, inflammatory disorders,
and gastric ulceration [30].
The coping responses during stress can be defined as cognitive and behavioral
responses to managing the stress. The primary objectives of the success of coping
responses are: to eliminate or mitigate the harmful environmental conditions and to
enhance the prospect of recovery; to tolerate and adjust the body to negative events;
to maintain a positive self-image; to maintain emotional balance; and to preserve
social relations [35].
Sleep implies a recovery process from previous wakefulness. Not only the dura-
tion of the waking period affects sleep architecture and sleep EEG, the quality of
wakefulness is also highly important. Studies in rats have shown that social defeat
stress, in which experimental animals are attacked and defeated by a dominant
196 5  Third Level: The Hypothalamus

Normal

Physiological response
Stressor

Activity Recovery

Allostatic load Time

Frequent stress Loss of adaptation


Physiological response Physiological response

Physiological response Physiological response


Repetitive response Normal adaptation

Time Time
Prolonged response Inadequate response

Absence of recovery

Time Time

Fig. 5.16  Allostatic load. The upper panel illustrates the normal allostatic response, which begins
with a stressor-induced response, maintained for a suitable period, and then reversed. The remain-
ing panels illustrate four conditions that give rise to allostatic load: (1) repetitive appearance of
multiple stressors (e.g., repeated elevation of BP leads to arteriosclerosis); (2) lack of adaptation
(e.g., those individuals who have to speak in public and do not reduce cortisolemia with repeti-
tion); (3) Prolonged responses due to a delayed reversal (e.g., catecholamine and cortisol secretion
due to stress returns to basal levels more slowly in the elderly), (4) an inadequate, low response
resulting in the compensatory hyperactivity of other mediators (e.g., autoimmune disorders with
inadequate secretion of corticosteroids and a high concentration of cytokines, which are usually
counter-regulated by corticosteroids). Modified with permission from Cardinali [1]

conspecific, is followed by an acute increase in NREM sleep on EEG and the sup-
pression of REM sleep. This occurred in both winners and losers, indicating that in
rodents a social conflict with an unpredictable outcome has quantitatively and
qualitatively similar acute effects on subsequent sleep, regardless of the results of
the conflict [36].
There is evidence that acute or chronic exposure to a stressor can start or cause
recurrence of psychiatric disorders such as depression, bipolar disorder, post-­
traumatic stress disorder, anxiety, and schizophrenia. In situations of melancholic
depression, anorexia nervosa, panicky anxiety, obsessive–compulsive disorder,
chronic active alcoholism, excessive exercise, abstinence from alcohol and narcotics,
Defense Behavior as a Paradigm of Reactive Homeostasis 197

Cortisol - Stress
CRH

+ + + Paraventricular
Adrenal cortex nucleus

+
ACTH Behavioral
ACh activation
- Pituitary gland

Medulla
+
+ -
+
NE

NE NE NE
Tachycardia,
increased BP

increased vascular
resistance
Adrenal medulla
Epinephrine release Gastroparesis

Fig. 5.17  The CRH neurons of the paraventricular nucleus of the hypothalamus command the
different manifestations of defense behavior. Modified with permission from Cardinali [1]

malnutrition, sexual abuse, the chronic activation of the hypothalamic–pituitary–


adrenal axis occur. On the other hand, in situations such as seasonal or atypical
depression, the postpartum period, the period after smoking cessation, fibromyalgia
syndrome or chronic fatigue, CRH secretion is diminished and symptoms occur such
as increased appetite and weight gain, somnolence, and fatigue [37, 38].
The parvocellular neurons of the hypothalamic PVN that synthesize and release
CRH and/or AVP are the final common pathway for the regulation of ACTH secre-
tion (Fig. 5.17) [30]. The relative proportion of active CRH/AVP neurons increases
significantly in stress [39]. Among other brain areas, these neurons project the
median eminence, to noradrenergic neurons of the brainstem, and to the ARC. In the
latter, they activate pro-opiomelanocortin -containing neurons, which, in turn,
reciprocally innervate CRH/AVP PVN neurons. Dendritic release of neuropeptides
has been described as a novel interpopulation signaling modality in the PVN [40].
A high density of neuronal cell bodies containing CRH is also found in the cen-
tral amygdaloid nucleus and the bed nucleus of the stria terminalis. CRH and its
receptors are found in extra-hypothalamic structures such as the limbic system,
basal forebrain, PBN, and paragigantocellular nuclei of the medulla oblongata, LC,
and other groups of noradrenergic neurons of the pons and medulla (sympathetic
noradrenergic system) and spinal cord [39].
There are reciprocal neural connections between CRH neurons in the PVN and
noradrenergic neurons in the brainstem. Cholinergic, serotonergic, and glutamatergic
198 5  Third Level: The Hypothalamus

neurons have stimulatory action in those circuits. The brainstem catecholaminergic


pathways to the PVN CRH neurons can be activated in bleeding, hypotension, respi-
ratory distress, and immune challenges. The LC is one of the regions of the brain
that is most responsive to stress, especially to hemorrhage. It also exerts its effects
on the hypothalamic–pituitary–adrenal axis through central limbic structures and it
innervates the PVN directly.
Glucocorticoids have both direct and indirect inhibitory actions on PVN neu-
rons. The hippocampus, which shows large numbers of binding sites for glucocorti-
coids and mineralocorticoids, is one of the structures that inhibit PVN activity [19].
Hippocampal damage increases mRNA expression for CRH and AVP in the PVN
and hippocampal stimulation decreases the activity of the hypothalamic–pituitary–
adrenal axis in rats and humans. The lateral septal area, the mPOA, and prefrontal
cortex are also structures that have inhibitory actions on the PVN.
The role of the medial prefrontal cortex as a coordinator of behavioral and physi-
ological stress responses across multiple temporal and contextual domains has been
proposed [41]. Glucocorticoids act as one of the primary messengers in the realloca-
tion of energetic resources, having profound effects locally within the medial pre-
frontal cortex, and shaping how the brain region acts within a network of brain
structures to modulate responses to stress.
Several neurotransmitters are released in the brain during stress, but CRH is
the main coordinator of the psychological and behavioral changes found (Fig. 5.17)
[31]. Experimental studies have shown that exogenous administration of CRH
stimulates the release of ACTH by the pituitary, can change the EEG, and induces
psychological and behavioral changes such as those observed in stress, for exam-
ple, reduced feeding behavior. Hormones of the pituitary–adrenal axis do not
block behavioral effects, but they can be reversed by the neutralization of CRH,
indicating that CRH can modulate behaviors independently of the pituitary–adre-
nal axis (Fig. 5.18).
Therefore CRH produces activation of ANS via direct central action, rather than
through the activity of the hypothalamic–pituitary–adrenal axis. CRH is part of a
family of peptides that includes urocortin and urotensin. These peptides act via two
receptor types: CRH-R1 and CRH-R2. The CRH-R1 receptors are most abundant in
the brain and in peripheral tissues and have a higher affinity for CRH and urocortin.
The CRH-R2 receptors are less abundant in the nervous system, are more significant
in peripheral tissues and exhibit higher affinity for urocortin, urocortin II, and uro-
cortin III than for CRH [31].
The predominant receptor in the activation of the hypothalamic–pituitary–adre-
nal axis is CRH-R1, whereas CRH-R2 plays a prominent role in the energy expen-
diture processes involved in homeostatic responses. Animals genetically deficient in
CRH-R1 show a failed stress response, whereas those deficient in CRH-R2 are
hypersensitive to stress and exhibit increased anxiety. It is possible that the acute
stress response phase is linked to rCRH-1, whereas the final recovery stage includes
rCRH-2 effects. Selective antagonists of rCRH-1 inhibit the anxiogenic action of
CRH and have been proposed to be a new type of anxiolytics. Hypersecretion of
CRH is as a phenotypic element of emotional disease vulnerability (Fig. 5.18).
Defense Behavior as a Paradigm of Reactive Homeostasis 199

Genetic Early vital


vulnerability experiences

Development

Trauma CRH antagonists


Life events Vulnerable Antidepressants
Daily stress Phenotype: ↑ CRH Psychotherapy

Biological changes: Behavioral changes:


HPA & ANS activation depression, anxiety

Fig. 5.18  Partly for genetic and epigenetic reasons, a vulnerable phenotype of CRH hyper-­
responsiveness to stress leads to emotional disorders. The use of pharmacological agents or psy-
chotherapy counteracts the consequences of this phenotype. HPA pituitary-adrenal axis. Modified
with permission from Cardinali [1]

Significance has also been given to the early effects of glucocorticoid hypersecre-
tion, with lasting changes in these systems in the life of the individual [31].
The hierarchical relationship between the hypothalamus and the limbic system
that controls it as the last level of autonomic motor hierarchy is clear when the
defense behavior is tested. An animal bearing a hypothalamic deafferentation that
disconnects the limbic system responds to the appearance in the visual field of any
object with the reaction of “false rage.” This indicates that the hypothalamus con-
tains the “program” of the defense reaction, which acquires emotional meaning and
purpose under the control of the limbic system. Maintaining high levels of cortisol
in stress leads to verifiable chronic damage of hippocampal neurons with significant
cognitive deficits (Fig. 5.19) [30, 42].
Stress interferes with the hypothalamic circuitry regulators of appetite and sati-
ety. CRH is a potent anorexic substance whereas NPY is the most potent known
orexinergic substance. NPY neurons stimulate CRH neurons, constituting a feed-
back loop involved in the control of food intake. Simultaneously, NPY inhibits the
LC, and stimulates food intake [43]. Glucocorticoids induce gluconeogenesis and
insulin resistance, by inhibiting lipolytic enzymes and stimulating enzymes that
facilitate the deposition of fat, thus antagonizing the actions of GH and gonadal
hormones on lipolysis and muscle and bone anabolism. Chronic stress can cause
increased visceral adiposity, the suppression of osteoblastic activity, and the reduc-
tion of lean mass (muscle and bone). Monkeys subjected to chronic social stress
show the classical symptoms of metabolic syndrome, such as high deposition of
200 5  Third Level: The Hypothalamus

Pyramidal CA1
neuron

CA3
Granule cell

Enthorinal cortex

Dentate Pyramidal
gyrus neuron Presynaptic
kainate receptor

5-HT STRESS

From enthorinal Glucocorticoids Glucocorticoids


cortex

NMDA (+)
INCREASED NEUROGENESIS
- NMDA blockage Stem cell
- Estrogens
- Behavioral stimulation MR + GR MR + GR
GABA
- Exercise BZP
- Learning NMDA (+)
-Adrenalectomy DECREASED NEUROGENESIS
Interneurons
- NMDA activation
- Stress
- Glucocorticoids
DENTATE GYRUS CA3

Fig. 5.19  Glucocorticoids and hippocampal neurogenesis. Several factors are listed that decrease
or increase neurogenesis. MR, GR receptors for mineralo- and glucocorticoids. Modified with
permission from Cardinali [1]

visceral fat, higher incidence of coronary atherosclerosis, hyperglycemia, and insu-


lin resistance. Hypertension, adrenal hypertrophy, hypersensitivity to adrenal
ACTH, hypogonadism and high cholesterol and lipid levels in the blood are also
found. The correlation between chronic stress and metabolic syndrome is also
observed in humans [44]. The phenotype of central or visceral obesity and a decrease
in lean body mass are seen in patients with Cushing’s syndrome, and in patients
with melancholic depression and chronic anxiety, all situations of hypercorti-
solemia. Visceral adiposity establishes a vicious cycle of an increased need for insu-
lin, hyperglycemia, and hypercholesterolemia.
Chronic stress affects therapeutic efficacy in cancer. For example, epidemiologi-
cal studies revealed strong correlations between long-term survival/cancer progres-
sion and β-adrenoceptor blocker use in patients [45, 46]. The contributions of stress
to immunosuppression in the tumor microenvironment and the implications of these
findings for the efficacy of immunotherapies have been discussed [47]. Effects of
norepinephrine (NE) on immune cells, such as those discussed in Fig. 4.25, are
presumably involved.
Studies in rodent tumor models have identified adrenergic receptor expres-
sion on various cancer cells including mammary carcinogen-induced tumors,
melanoma, and pituitary tumors. In humans, the Ewing sarcoma, neuroblastoma,
Defense Behavior as a Paradigm of Reactive Homeostasis 201

rhabdomyosarcoma, lymphoma, melanoma, and pancreatic, lung, breast, and pros-


tate cancer cells all displayed detectable levels of adrenoceptors. Several studies
demonstrate that the activation of adrenergic receptors promotes tumor progres-
sion. For example, chronic activation of G protein-coupled receptors, such as the
α1B-adrenergic receptor, can induce malignant transformation in normal cell lines,
promote DNA damage, and enhance tumor formation. In addition to a role in cell
survival, adrenergic receptor signaling has also been widely studied for mediating
metastasis. The regulation of metastasis by adrenergic receptors occurs at multiple
levels and involves not only cancer cells, but also cells in the tumor microenviron-
ment and in the metastatic niche. [47, 48]. Sympathetic activation modulates gene
expression programs that promote the metastasis of solid tumors by stimulating
macrophage infiltration, inflammation, angiogenesis, epithelial–mesenchymal tran-
sition, and tumor invasion, and by inhibiting cellular immune responses and pro-
grammed cell death. Hematological cancers are modulated by ANS regulation of
stem cell biology and hematopoietic differentiation programs [49].
In the author’s laboratory, two breast cancer tumor lines were used to assess the
reactivity of sympathetically denervated murine skin [50]. M3 tumors had a rela-
tively high capacity for local growth and a low capacity for metastasis, whereas
MM3-LN tumors grew locally at a slower rate, but metastasize very early to the
lung. After local implantation in the ear, the growth of M3 and MM3-LN tumors
was significantly slowed in the previously sympathetically denervated skin terri-
tory; however, their metastatic capacity remained unaffected by SCGx [50].
Stress related to inflammatory and infectious processes can alter the reproductive
function by the inhibitory action of substances released by the immune system cells
on the hypothalamic–pituitary–gonadal axis [51]. CRH suppresses GnRH release,
whereas glucocorticoids act on the pituitary and gonads causing inhibition of the
secretion of luteinizing hormone (LH), follicle-stimulating hormone (FSH), and
gonadal steroids. In stressed women, a decreased frequency of LH pulses occurs
together with increased cortisol secretion, which are assigned respectively to the
decreased activity of the GnRH pulse generator and increased CRH activity. The
central administration of cytokines reduces plasma LH levels, probably by increas-
ing opioids and prostaglandins, which inhibit the secretion of GnRH. Additionally,
cytokines can inhibit the activity of enzymes necessary for gonadal steroidogenesis.
The interaction of immune activity with the stress system is illustrated by the fact
that vaccination may not be completely effective at establishing the immune defense
to immunization that occurs during stress. Antigens are stressful stimuli (noncogni-
tive), and the antigenic challenge features a stressful situation (Fig. 5.20). Infection
stressor is a stimulus that induces the production of cytokines by the immune sys-
tem and glial cells. Inflammatory and infectious processes stimulate visceral vagal
afferents that activate the CRH stress system via the LC [52]. The hyperactivation
of the hypothalamic–pituitary–adrenal axis induces the hypersecretion of glucocor-
ticoids and immunosuppression, which are among the main consequences of stress
(Fig.  5.20). The hypothalamic–pituitary–adrenal axis produces glucocorticoids,
inhibiting the immune system. This, in turn, produces cytokines that are stimulatory
202 5  Third Level: The Hypothalamus

Stress Limbic system


(hippocampus, amygdala)

PVN

CRH
ADH
Hypothalamus

Inflammation
(IL1, TNFα)

Anterior pituitary
Corticotropic
cell

ACTH

Adrenal cortex

Cortisol

Fig. 5.20  Antigens are stressful stimuli (noncognitive), and the antigenic challenge features a
stressful situation

for CRH release. Thus, this constitutes a mechanism for feedback between the neu-
roendocrine and immune components in stress. The prevalence of the neuroendo-
crine component causes immunosuppression, whereas the prevalence of the immune
component causes autoimmune disease [19, 53].
The above-mentioned mechanisms also include the peripheral portion of the
ANS. From studies in the SCG territory, it can be concluded that during the aug-
mented NE release from peripheral sympathetic nerves innervating the hypotha-
lamic–hypophyseal unit, the modulatory role of peripheral nerve terminals is
differentially exerted on the release of hypophysiotropic hormones, inhibiting the
release of all pituitary hormones, except for ACTH (Fig. 5.21) [54]. In animals with
chronic SCGx, both ACTH and corticosterone rhythms were suppressed. The
mechanism through which peripheral sympathetic neurons are capable of modify-
ing the function of the median eminence remains undefined. As SCG efferences
play an important role in regulating the cerebral and choroidal blood flow, the out-
comes of SCGx on the median eminence may represent a particularly sensitive
vasomotor effect of the peripheral noradrenergic terminals. Otherwise, they may
involve a more complex interrelationship, such as a direct modulatory effect on
hypophysiotropic hormone release [54].
24-h Rhythms in Food Intake, Energy Storage, and Metabolism 203

LIMBIC SYSTEM
ASSOCIATION
(frontal and
CORTEX
mesencephalic poles)

HYPOTHALAMUS Autonomic descendent pathways

Hypophysiotropic
SON, PVN
area
SCG Cervical
Median
sympathetic
eminence ↓ CRH, AVP, SS
NE MCG, ICG ganglia
↑GnRH, GRH, TRH

Thoracic and lumbar


ANTERIOR POSTERIOR sympathetic ganglia
HYPOPHYSIS HYPOPHYSIS
NE
Vagus nerve

Supressed release of TSH, GH, AVP, LH, FSH and PRL (estrous cycle)
Stimulated release of ACTH and (stress)

Fig. 5.21  Diagram that summarizes the changes in the hypothalamic–pituitary unit after cervical
sympathetic activation. SON supraoptic nucleus, PVN paraventricular nucleus, SCG superior cer-
vical ganglion, MCG medial cervical ganglion, ICG inferior cervical ganglion. Reproduced with
permission from Cardinali [21]

24-h Rhythms in Food Intake, Energy Storage, and Metabolism

The first observations about daily changes related to nutrition and feeding were
those of Sanctorius in the seventeenth century, who conducted the first autorhyth-
mometry study by building a large scale in which he lived for months, recording his
weight and the ingested food, and collecting his feces and urine. Among other
observations, he detected a daily rhythm in body weight.
The complexity of the circadian system, described in Chap. 2, is exemplified
when the rhythmicity of food and nutrition is studied (Fig. 5.22). The 24-h rhythm
in feeding activity is controlled by the SCN. On the one hand, feeding time has
synchronizing effects on the circadian system and food restriction can entrain cer-
tain rhythms through a food-related pacemaker. Finally, food, besides being a syn-
chronizer and a variable controlled by the clock, food is also a masking stimulus
that modifies many rhythms related to digestion, absorption, and metabolism of
nutrients.
Feeding behavior is the first element to consider in the nutritional process of
organisms. Most of the studies have focused on explaining how the homeostatic
regulation (of the quantity and quality of food) takes place, with the hypothalamus
204 5  Third Level: The Hypothalamus

SCN Circadian
Regulation

Third ventricle
LHA DMN

VMN
Food Intake
ARC

Optic
tract
ME

Insulin, Leptin Homeostatic Hypothalamic nuclei


CCK, other Regulation

Fig. 5.22  Rhythmicity of feeding and nutrition. On the one hand, the feeding time has synchro-
nizing effects; on the other, the food restriction can synchronize certain rhythms through a food-­
controlled pacemaker. Food activity is in turn controlled by the SCN. In addition to being a
synchronizer and a variable controlled by the clock, food is a masking agent to directly modify
numerous rhythms related to the digestion, absorption, and metabolism of nutrients. The figure
was prepared in part using image vectors from Servier Medical Art (www.servier.com), licensed
under the Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/
license/by/3.0/)

as the main neural structure involved in close coordination with the release of
peripheral hormones such as leptin, CCK, and insulin [55]. Fewer studies have been
devoted to the temporal aspects of such regulation.
Eating is a consummatory behavior that involves the participation of various
organs, tissues, and systems. The control of eating behavior involves at least two
organizational levels: (a) a homeostatic one, in which afferent information of nutri-
tional status/energy reaches the CNS derived from adipose tissue, plasma, stomach,
the small and large intestines, the pancreas, and the liver; (b) a hedonistic one, in
which the CNS evaluates both the peripheral afferents and those from other brain
areas, including the visual, olfactory, gustatory, oral and lingual, general sensitivity,
and visceral sensitivity, pondering the nutritional value, palatability, and how pleas-
ant the food is (Fig. 5.23) [55].
Regulating appetite and adiposity involves short-term signals originating from
meals (quantity and taste) that determine the start and end of a meal; the level of
energy reserve does not trigger them. Other, long-term signals are related to adipos-
ity and intake coupled with caloric expenditure (Figs. 5.24 and 5.25). The regulation
of each meal is based on the ingested volume, as the post-absorptive signals have
little influence on the intake that produces them. Instead, the post-absorptive signals
24-h Rhythms in Food Intake, Energy Storage, and Metabolism 205

Circadian modulation

Hedonic control Homeostatic control Peripheral signals


HYPOTHALAMUS Ghrelin
ARC Leptin
VTA,
N. Accumbens AgRP, NPY POMC,CART
Insulin
GLP-1
MESOLIMBIC NPY αMSH CCK
DOPAMINERGIC PYY-3,36
SYSTEM PP
Orexin, MCH CRH, Oxytocin Glucose
FFA
LHA PVN
Dorsal striatum

Raphe L.C.
HUNGER SACIETY

Eating behavior

Fig. 5.23  Food behavior is regulated by homeostatic and hedonic mechanisms and by peripheral sig-
nals. VTA ventral tegmental area, LC locus coeruleus, AgRP peptide related to the agouti gene, NPY
neuropeptide Y, MCH melanin concentrating hormone, POMC proopiomelanocortin, CART transcript
regulated by cocaine-amphetamine, GLP-1 glucagon-like peptide, PP pancreatic polypeptide

Cerebral cortex

NPY/NE MCH/Orexin
PVN
NPY-AgRP
POMC-CART
Oxytocin

sPVz DMN
CCK ANS
POMC-CART PBN
NPY-AgRP

SCN MCH/
CCK POMC-
VMN CART Orexin

LHA
Arcuate NPY/NE
NPY-AgRP
Orexin POMC-CART

Oxytocin/ADH Pituitary gland


CRH/TRH

Fig. 5.24  Circuits involved in the control of food intake. The pituitary gland and the CNS regions
with their respective neuropeptides and/or hormones are represented. SPVz subparaventricular
zone. For other abbreviations see the text
206 5  Third Level: The Hypothalamus

Dorsomedial Paraventricular n.
hypothalamus
Lateral
Ventromedial hypothalamic area
hypothalamus

Arcuate nucleus

Anterior pituitary

Appetite
– +

Fasting Hypothalamus
PVN

LHA LHA

– GHRH Somatostatin
ARC
+ –
– +
Fat Adeno-
hypophysis
+ – +
Leptin

PYY3,36 GH

GIT Ghrelin
+ Growth

Fig. 5.25  Peripheral signals in appetite control. Leptin and PYY 3.36 inhibit it; ghrelin increases
it. Modified with permission from Cardinali [1]

do influence the start of the next feeding period. Thus, satiation (or psychosensory
satiety) is determined by the total volume of food ingested; it also includes specific
satiation for different nutrients (carbohydrates, fats, and proteins). Satiety (or meta-
bolic satiety) includes the suppression of the sensation of hunger and food intake
and the duration of this phenomenon (interprandial period) [56].
There are three groups of afferent signals that modulate intake: (a) signals origi-
nating in the sensory systems (vision, smell, taste); (b) afferent signals originating
from the gastrointestinal tract; (c) post-intake signals caused by nutrients or metab-
olism. Long-term intake is adjusted by metabolic indicators related to the degree of
filling of fat reserves (e.g., leptin or visceral afferents originating in the adipose
tissue). Homeostatic and hedonic controls are closely interrelated and often act in
unison at the unconscious level to achieve biologically adaptive responses [57].
24-h Rhythms in Food Intake, Energy Storage, and Metabolism 207

In relation to the signals involved in satiety, they are elicited by dilation and activity
of the digestive tract and by various nutrients and the gastrointestinal hormones
produced (Fig. 5.25). They include the anorexic leptin signal produced by adipose
tissue and the peptide PYY 3-36 synthesized by the intestinal wall. Orexinergic
signals such as ghrelin are also produced by the gastrointestinal tract; ghrelin stimu-
lates appetite in addition to increasing GH and PRL release (Fig. 5.25) [58].
Classical studies have shown that experimental animals subjected to the electri-
cal stimulation of LHA develop a typical food-seeking behavior (Fig. 5.26) [1, 59].
The reaction includes somatic motor activity, salivation, increased intestinal motil-
ity and blood flow, and decreased muscle blood flow. Conversely, stimulation of the
ventromedial hypothalamus (VMH) produces satiety and catabolic-type behavior.
Stimulation of LHA produces hunger, increased parasympathetic activity, and
from a metabolic point of view, glycogen synthesis, inhibition of gluconeogenesis,
hypoglycemia, insulin release, and lipogenesis. VMH stimulation produces satiety,
increased sympathetic activity, and from a metabolic point of view, glycogenolysis,
gluconeogenesis, hyperglycemia, glucagon secretion, and lipolysis (Fig. 5.26).

Central control Controlled


signal systems
(From the limbic
system)
– +
Sympathetic
activity
+ Para-

sympathetic
activity
H + Insulin

Y secretion
P Ventromedial
O nucleus – +
Glucagon
T
secretion
H
Σ A + –
L Lateral Glycogen
A hypothalamic synthesis
M area – +
U
Glycemia
S
+ –
Lipid
deposits

+ Eating

behavior

Signal error
Feedback signal
Σ Short term
From glucoreceptors Sensors Long term
and receptors for
Central
other metabolites Controlled
receptor
Σ variable
Peripheral Appetite
receptor
Reference signal

Fig. 5.26  Analysis of hypothalamic control of appetite according to the systems control theory.
Modified with permission from Cardinali [1]
208 5  Third Level: The Hypothalamus

By contrast, lesions of each of these areas have opposite effects. VMH lesions
produce hyperinsulinemia, hyperphagia, overeating, and weight gain. Reaching a
certain level of obesity, the animal properly regulates intake and maintains its
weight. In this model, vagotomy blocks hyperphagia, obesity, and hyperinsulinemia.
LHA lesions produce opposite effects, with decreased food intake and weight.
There is a close correlation of anhedonia with reduced dopaminergic neurotrans-
mission in the nucleus accumbens (Chap. 6) [1].
The amygdala, nucleus accumbens, and orbitofrontal cortex, as parts of the lim-
bic system, integrate the homeostatic information with motivational cognitive
aspects (aversive, hedonic) of eating (Fig. 5.23). Different limbic association areas
are stimulated synchronously with the LHA and VMH during food intake. Specific
lesion experiments indicate that the limbic areas provide purpose to the conduct
regulated by the hypothalamus.
Figures 5.23, 5.24, and 5.27 summarize the orexigenic and anorexigenic circuits
found in the hypothalamus. They include [2, 56]:

Activation of the hypophyseal adrenal axis ANS Wakefulness,


Food intake

PVN
Sleep

Lat. H
VLPO
Thermoregulation

dSPZ
MPO vSPZ

DMH
SCN
VMH

Nutrients ARC
Leptin
Ghrelin

Fig. 5.27  The SCN projects on the subparaventricular zones, both the ventral (vSPZ) and the
dorsal (dSPZ), and on the DMH. vSPZ neurons receive information for the organization of the
daily sleep/wake rhythm and project it to the DMH, which integrates information with other
sources (e.g., leptin and hhrelin in the ARC). The DMH is the source of projections that regulate
circadian sleep/wake rhythms, pituitary–adrenal axis activity, and descending ANS pathways, in
addition to vigilance and food intake (at the level of orexinergic neurons and MCH of the lateral
hypothalamus. Integration stations allow the adaptation of circadian rhythms to environmental
stimuli such as food availability, and cognitive and emotional influences of the limbic system.
Modified with permission from Cardinali [1]
24-h Rhythms in Food Intake, Energy Storage, and Metabolism 209

• A neural network of orexigenic activity that includes neurons having NPY,


GABA, galanin, orexin, opioids, agouti-related peptide (AgRP) or melanin-­
concentrating hormone (MCH) as a transmitter and that translate and release
appetite stimulus signals. The orexinergic/MCH neurons also participate in the
alertness processes (Chap. 2).
• Different anorexigenic pathways, including neurons using CRH, glucagon-like
peptide-1 (GLP-1), α-melanocyte stimulating hormone (α-MSH) or cocaine- and
amphetamine-regulated transcript (CART), which are responsible for the extinc-
tion of appetite by interrupting the action of the orexinergic network.
• Nuclei of the medial hypothalamus (dorsomedial hypothalamus, DMH, and the
VMH) that tonically restrict orexigenic signals in the interprandial interval. They
are activated by leptin, a hormonal signal that decreases orexigenic activity.
Ghrelin exerts opposite effects to leptin.
• A circadian regulation, coordinated by the SCN (Figs. 5.23, 5.24, and 5.27).

The integration of metabolic information requires multiple specialized areas of


the CNS: (a) brainstem (NTS, dorsal vagal complex), integrating peripheral neural
information with limbic structures (nucleus accumbens, amygdala, orbitofrontal
cortex); (b) mesencephalon and thalamus, receiving sensory information from dif-
ferent levels of the gastrointestinal tract; (c) anterior brain, providing aversion and
gratification generated by food [56].
Glycemia is one of the stimulatory signals important in regulating short-term
appetitive behavior. The nervous system has two types of receptors for glucose. The
peripheral receptors are located on the tongue, portal vein, duodenum, intestine and
pancreas, and generate changes in the activity of their respective visceral afferents
[60]. Central glucose receptors are located in the hypothalamus [61].
The first relay point of peripheral information on glycemia is the brainstem at the
level of the NTS. The integrated information at this level follows two directions: (a)
to return caudally via a reflex including neurons of the dorsal motor nucleus of the
vagus to affect the viscera; (b) in a rostral direction, the information reaches the
hypothalamus, where it undergoes a second process of integration, adding to the
information provided by central glucose receptors, which monitor the concentration
of blood and CSF glucose.
The ARC and NTS are the only brain areas with neurons expressing pro-­
opiomelanocortin (POMC), a precursor of several neuropeptides with signaling capac-
ity in food intake control (Figs. 5.24, 5.28, and 5.29). The most important neuropeptide
in feeding control is the POMC product α-MSH, which is associated with the reduction
of food intake [62]. The NTS and dorsal nucleus of the vagus also have the highest
concentrations of the α-MSH receptor, melanocortin receptor type 4 (MCR4). The
injection of an agonist of MCR4 (i.e., α-MSH) at the fourth ventricle or directly into the
NTS is effective in reducing food intake and body weight, whereas administration of
an MC4R antagonist (e.g., AgRP) increases food intake and body weight [62].
The ARC is in a dorsal position immediately adjacent to the median eminence
(Fig. 5.2). A group of neurons located in the medial portion of the ARC expresses
210 5  Third Level: The Hypothalamus

Pre - proopiomelanocortin (POMC)

Signal peptide

POMC (265 aa)

Pro-ACTH b-LPH (93 aa)

N-POMC ACTH (39 aa) g-LPH b-E (31 aa)

LPH = lipotropin
E = endorphin N-POMC g-MSH BP a-MSH (13 aa) b–MSH a–E

BP = binding peptide

Fig. 5.28  Processing of POMC

NPY, which is a major stimulator of feeding behavior [63]. The injection of NPY in
the lateral ventricles or in the area around the fornix evokes a substantial increase in
food consumption. ARC neurons that express NPY project to the PVN and the
LHA. Another group of ARC neurons synthesizes POMC and mainly project to the
PVN, the DMH, and to brainstem neurons controlling the sympathetic activity via
the intermediolateral column of the spinal cord [63].
Leptin conveys information on the nutritional status of the individual and reaches
the neurons of the ARC through the bloodstream, specifically via fenestrated ves-
sels of the median eminence. Leptin, ghrelin, and insulin modulate NPY expression
in ARC neurons. There is strong evidence that many of the actions of leptin are
mediated by stimulation of the melanocortin system, decreasing NPY. In the hypo-
thalamus, NPY is synthesized by neurons of the ARC and secreted from their termi-
nals in the PVN and lateral hypothalamus [64].
Administration of leptin to fasted rats (in which circulating leptin levels are
decreased) decreases the augmented synthesis of NPY found in ARC. Ghrelin, on
the other hand, acts in an opposite way to leptin, by increasing the expression of
NPY mRNA and ultimately inducing an increased consumption of food. Insulin
acts synergistically with leptin to decrease the augmented expression of NPY in
fasted animals. Another peptide, PYY 3-36, which is released into the bloodstream
by the GIT cells proportionally to the amount of food ingested, acts as another feed-
back signal inhibiting the release of NPY [64].
Neuropeptide Y neurons express AgRP. This agouti protein is a paracrine
secretion molecule acting as an antagonist of melanocortin receptors [65]. The
MC-4 receptors have been largely related to the control of feeding behavior,
24-h Rhythms in Food Intake, Energy Storage, and Metabolism 211

Median mPOA
PON MC4R
Ventrolateral
POA TRH CRH
MC4R
sPVz
PVN NMD
Vagus
SCN
POMC
MC4R NTS
Circadian
DMH Orexin NPY/
system
a/MCH AgRP
LHA Vagal and
sympathetic
afferents
VMH
Saciety
signals

POMC/ NPY/
CART AgRP
ARC

Adiposity signals,
Nutrients

Fig. 5.29  Anatomical circuits involved in energy homeostasis. The circadian apparatus (in red),
the centers of control of the intake (in blue) and links with the ANS (in gray) are represented sche-
matically. Overlays are indicated in mixed red/blue. The master clock of the circadian network is
the SCN. The key structure of the energy network that integrates adiposity (leptin and insulin) and
signals related to nutrients (glucose, fatty acids) from the periphery is the ARC. Both circadian and
energetic networks are indirectly connected through the DMH and the LHA, which are both sensi-
tive to nutritional and circadian information. The nucleus of the solitary tract (NTS) receives sig-
nals of satiation from the peripheral organs via vagal and sympathetic afferents and is innervated
by the PVN and the LHA; thus, it also receives signals from the circadian and energetic networks
of the hypothalamus

and mutations in their genes cause obesity in rodents and humans (Figs. 5.23,
5.24, and 5.27). As already mentioned, the MC4R preferential agonist ligand is
α-MSH. Intracerebroventricular injections of α-MSH decrease food intake in nor-
mal animals. In fasted animals, POMC mRNA is decreased, which can be restored
by the administration of exogenous leptin or insulin.
The PVN and LHA exhibit a significant amount of MC4R, and a relatively dense
innervation of fibers containing α-MSH. Fibers containing AgRP and NPY also
innervate these regions. Thus, the PVN and the LHA contain two distinct popula-
tions of nerve terminals releasing POMC or AgRP, both responsive to circulating
leptin, whose neurotransmitters would act on the same receptors (MC-4, MC-3) in
diametrically opposed ways, acting as agonists (α-MSH) or antagonists (AgRP) on
the same neural system (Figs. 5.23, 5.24, and 5.27) [62].
Another participating neuropeptide in PVN and LHA is CART, which is pro-
duced by the same neurons that express POMC. Its co-location in POMC neurons
indicates a conjoint action in the same areas on the MC-4 receptors. Studies using
neural mapping techniques have shown that ARC neurons that express CART and
212 5  Third Level: The Hypothalamus

POMC also project to the intermediolateral column of the upper thoracic spinal
cord and participate in the regulation of thermogenesis. Thus, the ARC neurons can
act on the maintenance of body weight by regulating both food intake and energy
expenditure [63].
A large concentration of leptin receptors is found in the dorsomedial portion of
the VMH, where many of the neurons located therein are sensitive to glucose. The
VMH projects heavily to the sPVNz, which in turn, receives dense projection from
the SCN. Many of the VMH neurons that project to the sPVNz are responsive to
circulating leptin, thus providing an anatomical substrate by which leptin can con-
trol the circadian variation of feeding (Fig. 5.27) [2].
The VMN participates in the control of endocrine and autonomic systems, indi-
rectly modulating the information starting from neurons located in sPVNz and tar-
geting the DMH, which is essentially an integrating region that ultimately projects
to the PVN, modulating endocrine and visceral responses [40]. Many studies sug-
gest the involvement of the DMH in controlling the intake of water and food, and
body weight. Stimulation of DMH results in changes in the activity of the pancreatic
nerves, and DMH lesions induce hyperglycemia, indicating that the DMH regulates
insulin secretion via projections to the autonomic centers. In addition, several stud-
ies suggest DMH involvement in the control of the cardiovascular system, stress and
anxiety, and locomotion. Because of this complexity, it has been proposed that the
DMH is one of the main components of a hypothalamic pattern generator for the
visceromotor system [66].
A population of neurons from the LHA that contain MCH projects extensively to
various regions of the CNS of mammals (Figs. 5.23, 5.24 and 5.27) [2]. The MCH
receptor (MCH-1R) is widely distributed in the CNS, with particularly dense
expression in the cerebral cortex (including the orbitofrontal, pre-limbic and senso-
rimotor regions, and the rhinencephalon), the nucleus accumbens, hippocampal for-
mation, the NTS, and the LC. Several studies have demonstrated the occurrence of
an increase in food intake following injection of MCH. Furthermore, the increase in
MCH expression leads to obesity and insulin resistance. Another population of neu-
rons in the LHA express orexin and project to regions of the brainstem and spinal
cord, such as the LC and the dorsomedial nucleus of the vagus nerve (Fig. 2.16) [2].
When orexin is injected into the cerebral ventricles, it causes increased food intake,
whereas orexin receptor antagonists decrease the intake. Mice whose orexin gene
was deleted exhibit narcolepsy and hypophagia. Thus, this peptide signals to the
neural systems that play a significant role in feeding behavior and the sleep–wake
cycle, possibly by coordinating a set of responses to complex behaviors and com-
plementary autonomic responses. Its link to sleep was discussed in Chap. 2.
In neuroimaging studies, hungry individuals show greater activity in the prefron-
tal cortex and decreased activity in the hypothalamus, thalamus, insular cortex, the
cingulate gyrus, orbitofrontal cortex, the basal ganglia, temporal cortex, and cere-
bellum. In obese and satiated individuals, activation of the prefrontal cortex was
greater than that found in normal individuals, and there is a greater reduction in the
activity of the limbic and paralimbic cortex compared with normal individuals of
both sexes [67, 68].
24-h Rhythms in Food Intake, Energy Storage, and Metabolism 213

There are separate mechanisms for controlling the balance of different nutrients
[55]. For the carbohydrate balance, the system operates to increase the intake of
carbohydrates when their peripheral utilization decreases. GABA, NPY, and NE in
the hypothalamic VMH, and glucocorticoids specifically increase carbohydrate
intake, with a circadian rhythm of maximum effect at the beginning of the activity
phase, coinciding with the peak plasma cortisol. 5-HT, CCK, insulin, and leptin
inhibit the effect.
For fat balance, the neuropeptide galanin, opioids, and mineralocorticoid
hormones are involved. These substances enhance specific fat intake by acting
on VMH. NPY and leptin in particular mediate in this mechanism. The effect
has a characteristic daily rhythm, with the maximum toward the end of the phase
of activity.
In humans, an appetite for fat is verified rising toward the afternoon. Dopamine
(DA) antagonizes the effect of galanin, opioids or aldosterone on fat intake. This is the
basis of the anorexic action (appetite suppressant) of amphetamine that acts by releas-
ing endogenous DA. Dopamine antagonists (neuroleptics) increase the appetite for fat.
For protein balance, GHRH and opioid peptides, leptin, and NPY are involved.
The circadian rhythm of protein preference is similar to that of fat, with the maxi-
mum occurring late afternoon.
Estrogens play a role in stimulating the appetite for carbohydrates. In women, a
greater preference for carbohydrate diets is detected. In men a greater preference for
predominantly protein diets occurs. The appetite for fat increases during puberty, in
both boys and girls, coinciding with increased hypothalamic galanin detected in
experimental animals. High levels of estrogen in obese women further increase the
synthesis of galanin in the hypothalamus, with increasing preference for fats. Thus,
there is a significant correlation between the content of hypothalamic galanin and
body weight in different species [69, 70].
Although there has been progress in determining the factors that modify the spe-
cific appetites for different nutrients, there is yet no simple physiological scheme
that accounts for the long-term control of body weight. Clearly, lipid deposits play
an important role and identification of leptin was a milestone in this regard.
Adipocytes produce leptin and it is now known that leptin serves as a signal for
adequate energy intake (Fig. 5.30). Circulating leptin is a signal indicating whether
deposits of sufficient energy exist to initiate a process of great energy demand such
as puberty. Leptin promotes inflammation, which provides a pathophysiological
link between obesity and insulin resistance, atherosclerosis, and autoimmune pro-
cesses. These processes are inflammatory diseases characterized by increased pro-
inflammatory cytokines such as IL-6, which is an example of the link between
energy homeostasis and immune function [71].
It is noteworthy that there is a viscerotopic representation of lipid deposits in
central areas [72]. The existence of a permanent regional neural mechanism that
links the load sensor of deposits to the activity of the higher centers is evident. For
example, an early overload of these deposits (in obese children) produces definitive
changes in the “set point” of the total mass contained, with the result of obesity in
adulthood.
214 5  Third Level: The Hypothalamus

Leptin resistance

Hypothalamus Mesolimbic
POA VMH dopaminergic
- GnRH ARC - BNDF system
PVN - POMC/CART LHA VTA SN
+ CRH + NPY/AgRP + Orexin
- TRH + MCH
¯ Leptin  Leptin
Energy Energy
deficit NTS
Pituitary gland excess
- GH
- TSH
- FSH/LH + ACTH

Neuroendocrine function Energy homeostasis Hedonic processing


¯ Sex hormones  Appetite Altered eating behavior
¯ Thyroid hormones ¯ Energy expense
¯ IGF1
¯ Cortisol

Fig. 5.30  Leptin action on the brain during states of excess energy and energy deficiency. During
excess energy, access of the leptin to the hypothalamus and other areas of the brain is impaired
(leptin resistance). In states of energy deficiency and therefore of leptin deficiency the neuropep-
tides that are normally inhibited by leptin are elevated (+) and neuropeptides stimulated by leptin
are suppressed (−). Changes in the concentrations of these neuropeptides lead to alterations in
neuroendocrine function and energy homeostasis. Alterations in leptin levels may also affect the
hedonic aspects of eating behavior

Circadian rhythms are described for most of the factors involved in food intake
[73]. Although in humans the rhythms of food intake are conditioned by cultural
factors, the relative stability of dietary habits between cultures supports the exis-
tence of strong biological determinants. As mentioned, nutrients that are a source of
quick energy such as carbohydrates are usually selected at the beginning of the
activity period, whereas fats and proteins are preferred before the beginning of the
period of rest. The rhythms of appetite, digestion, absorption, and activity of key
enzymes of metabolism are largely responsible for this (Fig. 4.32).
Meal times influence factors as varied as weight gain, glucose, glucose tolerance,
triglycerides, cardiovascular risk, etc. These effects are especially apparent in night-­
shift workers, who fail to produce a full synchronization of their metabolic rhythms
at nighttime (Chap. 8).
Unlike animals used as a model for studies of feeding behavior, humans do not
necessarily eat according to biological impulses, but also for cultural, religious, and
hedonic reasons. However, there are situations in which these cultural conditions
can be minimized. Newborns cry to demand food every 90–120 min, after 2 or
3 months they demand food only four or five times a day, and after 6 months the
feeding frequency begins to resemble that of the adult. The three main meals of an
adult are maintained throughout life, even in those situations in which there is tem-
porary external isolation.
24-h Rhythms in Food Intake, Energy Storage, and Metabolism 215

There are rhythms in neuroendocrine regulatory factors and in caloric and mac-
ronutrient intake [73]. During nocturnal sleep, levels of glucose and insulin secre-
tion increased significantly to return to baseline in the morning. During sleep
deprivation, glucose levels and insulin secretion were negligible whereas daytime
sleep was associated with rises in glycemia and insulin secretion (Fig. 5.31). The
diurnal variation in insulin secretion was inversely related to the cortisol rhythm;
however, sleep-associated rises in glucose correlated with the amount of concomi-
tant GH secreted [74].
Similarly, leptin is regulated by the sleep/wake homeostat and has a negative cor-
relation with ACTH and cortisol being high during NREM sleep. The stomach
secretes ghrelin, it has orexic activity, and food intake suppresses its secretion. The
mean values of the day are higher than those in the evening and overnight levels are
higher in the first half than those of the second half [58].
Increased leptin is one of the signals responsible for appetite suppression during
sleep. Coincident with this regulation of secretion of leptin, insulin is released

DAY 1 DAY 2
140
Glycemia (% of mean)

130

120

110

100

Sleep Sleep Diurnal


deprivation sleep
180
Insulin (% of mean)

160

140

120

100
18 22 02 06 10 14 18 22 02 06 10 14 18
Time of day (h)

Fig. 5.31  Insulin is a hormone whose constitutive secretion is mainly controlled by the sleep–
wake homeostasis. The volunteer was deprived of sleep for one night and could recover the lost
sleep from 11 am the next day. Ultradian variations in hormone secretion or glycemia or values
<100% of mean plasma incurred sample reanalysis or glycemia are not represented. Data from Van
Cauter et al. [74]
216 5  Third Level: The Hypothalamus

during slow-wave sleep. Alterations in insulin and leptin secretion are present in
shift workers, who tend to suffer a chronic sleep disturbance. Thus, sleep loss is
associated with insulin resistance, lower levels of leptin and higher levels of ghrelin,
leading to a serious risk of obesity [75].
In humans and rats, with age, a decline in melatonin levels occurs, whereas the
levels of visceral fat, insulin, and leptin increase [21]. These changes are often asso-
ciated with adverse metabolic consequences: glucose intolerance, insulin resistance,
diabetes, dyslipidemia, and hypertension. Treatment with melatonin reverses age-­
associated changes in retroperitoneal and epididymal fat, and plasma concentrations
of insulin and leptin to those found in young individuals, without significantly
affecting food intake. These findings, together with the ability of pinealectomy to
increase leptin levels, suggest that melatonin might exert an inhibitory effect on the
release of this hormone [21].
Damage of the hypothalamic SCN causes immediate loss of the circadian
rhythm of food intake, without alteration of the regulation of the total amount
of food ingested over 24 h. The homeostatic and temporal regulation reside in
different parts of the hypothalamus, but do overlap (Fig. 5.29). In addition to
influencing the internal temporal structure of the body, food, when it is restricted
in quantity and provided at a limited time, can act as synchronizer of structures
that act as a pacemaker on numerous circadian rhythms (Fig. 5.32) [76]. In this

Zeitgeber Oscillator Clock output

Light
Food intake
CNS
SCN
Other
Sleep/wake cycle

Food Autonomic and


endocrine
pathways

Hormones
Gastrointestinal
tract and other
peripheral tissues
Metabolic pathways

Fig. 5.32  Food as a circadian synchronizer. The figure was prepared in part using image vectors
from Servier Medical Art (www.servier.com), licensed under the Creative Commons Attribution
3.0 Unported License (http://creativecommons.org/license/by/3.0/)
24-h Rhythms in Food Intake, Energy Storage, and Metabolism 217

Fig. 5.33  Food-entrained oscillators. Under normal physiological conditions, synchronization of


behavioral rhythms, such as feeding, is controlled by the SCN. The importance of food as zeitgeber
is seen under conditions of food restriction, in particular hypoenergetic. The figure was prepared
in part using image vectors from Servier Medical Art (www.servier.com), licensed under the
Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

situation, a dissociation between the rhythms controlled by light, through the


SCN, and rhythms synchronized by food (food entrained oscillators, FEO) may
appear (Fig. 5.33). When an animal receives a single meal with a calorie content
of less than its daily needs, it progressively entrains rhythms to the feeding sched-
ule, showing increased levels of certain variables (motor activity, gastrointestinal
motility, activity of digestive enzymes, plasma cortisol) 1 or 2 h before the sched-
uled food availability. This anticipatory activity is controlled by an independent
pacemaker to SCN, because it remains in animals with SCN lesions (Fig. 5.33)
[76]. When food is abundant, the pacemaker entrained by light is directing all the
circadian rhythms; however, when food is scarce and available only at specific
times, a second pacemaker (FEO) takes over certain rhythms favoring the use of
food, but without neglecting other rhythmic processes that continue to be con-
trolled by the pacemaker synchronized by light (SCN). The FEO is very potent in
the liver and less effective in other peripheral organs such as the lung [76].
In Chap. 2, we discussed the molecular bases of the circadian clock as a network
of transcription–translation feedback loops. Rhythmic clock output is achieved
through E-box elements in controlled clock genes, which affect numerous cellular
processes (Fig. 5.34), including lipids and glucose metabolism (liver, muscle), adi-
pogenesis (white and brown fat), insulin sensitivity, and lipogenesis. In turn, various
218 5  Third Level: The Hypothalamus

Hexosamine path
PARP GSK3b O-GlcN-Ac
NADH/NADPH
NIGHT
(SIRTUIN 1) Proteosomal
CRY PER CRY PER degradation
NIGHT
BMAL 1

BMAL 1
NPAS 2
CLOCK

EVENING
CKIe/d
Per 1,2,3 PER PER
Lipid & glucose
E-box Target genes Cry 1,2 CRY metabolism
(liver, muscle)
MORNING Rev-erba REV-ERBa
AMPK
Rora RORa
PPARa Adipogenesis
CCGs (WAT & BAT)
Nucleus
PPARg
PGC-1a FAS
(+) Insulin
PPRE RORE Bmal1 BMAL1
sensitivity
(+) (-) (+)
PPAR REV-ERBa RORa
Lipogenesis

PPAR
Cytoplasm

Fig. 5.34 The circadian clock is a feedback network of transcription–translation loops


(Chap. 2). The rhythmic output of the clock emerges through E-box elements in clock-controlled
genes (CCGs), which affect numerous cellular processes, including lipid and glucose metabo-
lism (liver, muscle), adipogenesis (white adipose tissue, WAT, brown adipose tissue, BAT), insulin
sensitivity, and lipogenesis. Several metabolic signals provide feedback to the cellular circa-
dian system: AMPK, sirtuin 1, poly ADP-ribose polymerase (PARP), peroxisome proliferator-
activated receptors (PPARs), glycogen synthase kinase 3β (GSK3β), hexosamine/O-GlcN-Ac
O-β-DN-acetylglucosamine)

metabolic signals provide feedback to the cellular circadian system [73, 77]. They
include:
• AMPK: AMP-activated protein kinase. metabolic indicator of cell energy charge
• NAD(P)H: NAD(P)+: a metabolic indicator of a redox state (sirtuin 1)
• Poly ADP-ribose polymerase (PARP): secondary sensor cell energy charge and
redox state
• Peroxisome proliferator-activated receptors (PPARs), e.g., PGC-1α (peroxisome
proliferator-activated receptor γ co-activator 1-α): lipid metabolism sensors
• Metabolism of glycogen (glycogen synthase kinase 3β, GSK3)
• Via hexosamine/O-GlcN-Ac (O-β-D-N-acetylglucosamine): glucose level
signaling
Sirtuin 1 is an example of a metabolic sensor. It is an NAD+-dependent deacety-
lase (redox state sensor that measures the ratio NAD+/NADH) and a regulator of
clock machinery that binds to and inhibits the activity of CLOCK/BMAL1. Sirtuin
1 plays a key role in the regulation of gluconeogenesis, fat metabolism, insulin
secretion, and apoptosis. Inactivation of sirtuin 1 by oxidative stress leads to the
24-h Rhythms in Food Intake, Energy Storage, and Metabolism 219

abnormal transcription of proapoptotic and proinflammatory genes, oxidative stress,


inflammation, and premature cell senescence [20].
Different experiments in mice genetically modified in clock genes indicate their
involvement in feeding behavior. For example, feeding Clock-mutant mice with a
diet rich in calories doubles the accumulation of body mass and energy relative to
control. Mice lacking Per3 have a high body mass and body composition and
impaired adipogenesis. Mice with knockout of Bmal1 in the liver have altered
hepatic glucose metabolism, whereas mice with knockout for pancreatic Bmal1
show hyperglycemia, impaired glucose tolerance, and impaired insulin secretion.
Also, mice bearing knockout of Bmal1 in white adipose tissue has obesity by reduc-
ing regulatory signs to central regions of appetite [78].
As discussed in Chap. 2, the SCN can act on metabolism via both hormonal and
autonomic neural pathways. In the case of the adrenal gland, glucocorticoid and
blood sugar increase in the early phase of activity is the result of the activity of SCN
on CRH-producing hypothalamic neurons, which stimulate the secretion of ACTH
by the pituitary and this in turn activates the secretion of glucocorticoids by the
adrenal cortex. In addition, the adrenal gland is subjected to a circadian rhythm of
ACTH sensitivity, with maximum sensitivity during the activity phase. By neural
retrograde labeling techniques, it was found that the adrenal is connected to the
SCN via the ANS, through a multisynaptic path that regulates the sensitivity of the
gland (Fig. 5.12) [79]. Using similar procedures, neural pathways between the SCN
and the heart, pancreas, liver, thyroid, and pineal have been described (Chap. 4).
Therefore, through these communication mechanisms, the SCN may activate or
silence individual tissues depending on their function at different times of the day.
It has been shown that both the sympathetic and the parasympathetic autonomic
nervous system divisions may discriminate between different compartments of adi-
pose tissue, such as subcutaneous and intraabdominal fat. Compartmentalization of
motor neurons of the autonomic nervous system is the basis for the selective effects
of the sympathetic–parasympathetic balance in the different compartments of the
body. This anatomical segregation of autonomic neurons provides a physiological
basis for selective changes in the sympatho-parasympathetic balance in various
body compartments and at different times of the day [80].
From a circadian point of view, the body is divided into two functional autono-
mous compartments: (a) a thoraco-muscular compartment; (b) a visceral compart-
ment (Fig. 5.35). In the period of wakefulness, locomotor activity requires glucose
and free fatty acids. As a homeostatic reaction, the brain facilitates the release of
energy from the storage organs such as the liver and adipose tissue. If this activity is
repeated on a regular daily basis, the ANS is programmed to facilitate operation in
the form of an anticipatory daily rhythm for energy requirements. By contrast, dur-
ing the period of NREM sleep, the ANS switches into a state of anabolic control,
recovery, with energy accumulation in the organs of deposit and lower peripheral
glucose utilization [80].
Clinical and pre-clinical studies have established close associations between
insulin resistance and sympathetic activation and thus suggest coinciding mecha-
nisms in their development. Acute peripheral and central increases in insulin levels
220 5  Third Level: The Hypothalamus

UNBALANCED ANS
Metabolic
syndrome

Thorax
Blood pressure
Sympathetic

Parasympathetic

CV system

Abdomen
Insulin Abdominal fat
Sympathetic

Parasympathetic

Adipose
Pancreas
tissue

Muscle
Sympathetic Glucose uptake

Muscle

Fig. 5.35  Autonomic imbalance of the abdominal and thoraco-muscular territories in the meta-
bolic syndrome. The result is hypertension, increased insulin resistance, and abdominal obesity.
Modified with permission from Cardinali [1]

can elevate sympathetic activity through central insulin receptor action in key brain
regions regulating autonomic function. Conversely, central manipulation of the
sympathetic nervous system can affect insulin sensitivity peripherally through a
variety of mechanisms including modulation of the renin–angiotensin–aldosterone
system and pancreatic autonomic nerves. Enhanced insulin resistance and elevated
sympathoexcitation likely function in a feed-forward loop in which hyperinsu-
linemia enhances central sympathetic output, which in turn further promotes hyper-
insulinemia. In this manner, insulin and over-activation of the sympathetic nervous
system interact to adversely affect insulin metabolic signaling and contribute to the
development of metabolic syndrome and its associated complications [81].
What is possible to verify in the metabolic syndrome is a regional imbalance in
favor of the parasympathetic in visceral territory (abdominal fat), and in favor of the
sympathetic in the thoraco-muscular area (increased BP and insulin resistance;
Fig. 5.35). The analysis of anatomical circuits involved in energy homeostasis illus-
trated in Fig. 5.29 provides the bases for the intertwined regulation of circadian and
food intake mechanisms. Both networks, circadian and energy, are indirectly con-
nected at the DMH and are sensitive to time and circadian nutritional information
(Fig. 5.29). The NTS receives satiety signals from peripheral organs via vagal and
sympathetic afferents and is innervated by PVN and LHA, which also receive sig-
nals from the circadian hypothalamic networks and energy [82].
24-h Rhythms in Plasma Osmolality and the Intravascular Volume 221

Feeding time has demonstrable effects on weight gain [55]. In subjects who ate
for a week one meal a day of approximately 2000 kcal. composed of 50% carbohy-
drates, 15% protein, and 35% fat, a weight loss of about 1 or 2 kg was observed if
the food was taken in the morning. In contrast, none was observed if the food was
ingested at 17:30 h. Overweight people often have increased hunger at night com-
pared with those of normal weight; they also consume large proportions of their
caloric intake in the evening. The food consumed at night induces an increase in low
density lipoproteins (LDL) and decrease in high density liproteins (HDL), along
with increased insulin resistance and higher levels of nocturnal glucose. In general,
human late chronotypes (a) sleep 1 h less per night; (b) consumes more calories at
dinner; (c) consumes more calories after 20:00 h; (d) have diets of lower quality
(junk food, sugary drinks, fewer vegetables) [83].

 4-h Rhythms in Plasma Osmolality and the Intravascular


2
Volume

Both plasma osmolality and intravascular volume are the variables controlled by
water and electrolyte intake behavior. The hypothalamus regulates this behavior
through the supraoptic–PVN–neurohypophyseal system and AVP secretion, and via
several other hormonal systems, including activation of the renin–angiotensin sys-
tem and secretion of ANP [84].
The constancy of the internal medium composition is maintained primarily by
controlling the intake and renal excretion of salt and water. The dehydration and the
consequent need for water occur when there is a loss of water and/or an increase in
effective solute, especially sodium. The role of sodium is indicated by the fact that
a hypertonic solution of NaCl is more effective at raising water intake than equimo-
lar solutions of other non-ionic solutes [85].
A large body of evidence demonstrates that the 24-h changes in blood volume
and BP regulation largely depend on the interactions existing among the sympa-
thetic nervous system, the renin–angiotensin system, and renal sodium excretion
(Fig. 5.36) [86]. Twenty-four-hour rhythms in renal blood flow, glomerular filtra-
tion rate, urine volume, and urinary sodium, potassium, and chloride, with after-
noon to early evening peak time in diurnally active persons, are well-known and
persist independently of sleep, indicating their circadian nature [87]. Renal blood
flow, vascular resistance, and glomerular filtration rate decline at night, although
the decrease in urine flow, particularly in the non-elderly, is much more pro-
nounced than expected. This indicates a 24-h periodicity of tubular reabsorption
with nighttime peaks mediated by circadian rhythms of intrarenal angiotensin II
and AVP [86].
Thirst is a feeling that motivates water consumption, and is triggered by cellular
dehydration. A tiny increase in plasma osmolality of 1–2% can start neuroendocrine
responses and the search for water. Changes in blood volume and extracellular fluid
(ECF) pressure are also an important stimulus for the intake of water, but to a lesser
extent, i.e., a reduction of about 10% of blood volume or BP is needed to induce the
intake of water (Fig. 5.37) [88].
222 5  Third Level: The Hypothalamus

EFFECTOR
AFFERENT
RESPONSES
SIGNALS
Behavior:
Plasma osmolarity
Na+ and water intake
[Na+]

Blood volume AVP secretion


ANS activation
Renin-angiotensin activation
ANP secretion
Cardiovascular adaptation

Excretion :
Renal excretion of water
and electrolytes
Awake

Fig. 5.36  Schematic representation of the mechanisms controlling water and salt intake. This
control changes as a function of the three body configurations in a 24-h period. The figure was
prepared in part using image vectors from Servier Medical Art (www.servier.com), licensed under
the Creative Commons Attribution 3.0 Unported License (http://creativecommons.org/license/
by/3.0/)

Osmoreceptors are highly specialized neurons, able to translate changes in


external osmotic pressure into electrical signals. Current evidence supports the
hypothesis that cells sensitive to changes in CSF osmolality or sodium concentra-
tion are located in the circumventricular organs in the anterior ventral region of the
third ventricle (AV3V) and the area postrema (Fig. 5.37) [89]. Lesions of AV3V
cause adipsia.
In addition to its location in the CNS, the sodium receptors are also present on
afferent nerve terminals adjacent to the liver, kidney, and intestinal vessels. The
increase in sodium concertation in the portal vein stimulates hepatic receptors lead-
ing via vagal afferents to the activation of neurons in the NTS; these, in turn, stimu-
late neural structures, which induce increased natriuresis and the inhibition of
intestinal absorption of sodium [84].
When water is provided to an animal kept for 24–48 h with water restriction, in
about 3–10 min its thirst is quenched, as if the body could predict the exact amount
of water needed for the correction of osmolality simply by measuring the volume of
water that passed through the mouth and the stomach (Fig. 5.37). In fact, stimuli
generated in the mouth, throat or stomach are converted into afferent impulses to
CNS structures involved in the corresponding integrative response to thirst [90].
24-h Rhythms in Plasma Osmolality and the Intravascular Volume 223

Plasma Dry mouth and Water intake


Volemia osmolarity throat mettering (GIT)

Baroreceptors Osmoreceptors

+ +
+ -
Angiotensin Thirst
+

Fig. 5.37  Physiological components of the thirst response. Modified with permission from
Cardinali [1]

SFO
d MnPOn

SFO AC
SCO BD
ME OVLT vMnPOn
PIN
AV3V
NH OVLT
AP
OC

Fig. 5.38  Left: circumventricular organs. AP area postrema, NH neurohypophysis, OVLT orga-
num vasculosus of the lamina terminalis, SCO subcommissural organ, SFO subfornical organ, PIN
pineal gland. Right: the lamina terminalis is a forebrain structure that contains the SFO, the median
preoptic nucleus (MnPOn), and the OVLT. The AV3V includes the ventral part of MnPOn (vMn-
POn) and the OVLT. The AV3V region, the SFO, and the AP in the fourth ventricle contain neu-
rons that are sensitive to changes in osmolality. dMnPOn dorsal median preoptic nucleus, BD
diagonal band of Broca, AC anterior commissure, OC optic chiasm. The figure was prepared in
part using image vectors from Servier Medical Art (www.servier.com), licensed under the Creative
Commons Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)

The lamina terminalis is a forebrain structure containing the subfornical organ


(SFO), the MnPO, and the OVLT (Fig. 5.38). The AV3V region includes the ventral
part of the MnPO and the OVLT. AV3V, SFO, and the area postrema in the fourth
ventricle contain neurons that are sensitive to changes in osmolality. These cells
have direct projections to PVN and to the dorsal raphe and LC nuclei. These con-
nections are important for conveying information involved in hypothalamic behav-
ior to restore the balance of body fluids [85].
224 5  Third Level: The Hypothalamus

These areas also have connections with the kidneys. By using injections of a
neurotropic virus that causes retrograde infections of rat kidneys, a chain of neurons
involved in homeostatic regulation was identified, including the OVLT, MnPO,
SFO, bed nucleus of the stria terminalis, periventricular anteroventral nucleus,
SON, primary motor cortex, and the visceral area of the insular cortex [85].
The kidneys play a central role in cardiovascular homeostasis as they ensure a
balance between the fluid taken in and that lost and excreted during everyday activi-
ties. This ensures the stability of extracellular fluid volume and maintenance of
normal BP. Renal fluid handling is controlled via neural and humoral influences,
with the former determining a rapid dynamic response to the changing intake of
sodium, whereas the latter cause a slower longer-term modulation of sodium and
water handling [6, 91].
Activity in the renal sympathetic nerves arises from an integration of information
from the high and low pressure cardiovascular baroreceptors, the somatosensory
and visceral systems, and the higher cortical centers. Each sensory system provides
varying input to the autonomic centers of the hypothalamic and medullary areas of
the brain at a level appropriate to the activity being performed [92]. Renal innerva-
tion considered entirely of sympathetic origin participates in the homeostatic regu-
lation of volume and osmolality of organic liquids, exercising control over three
important aspects of renal function: (a) renal blood flow; (b) tubular reabsorption of
electrolytes; (c) renin secretion [91]. The renal nerves regulate the function of blood
vessels, tubules, and juxtaglomerular granule cells.
Activation of renal nerves release renin to the renal blood flow and decrease
urinary sodium excretion (Fig. 5.39). The basal discharge rate of renal sympathetic
nerves is 0.5–2 Hz, causing a continuous release of NE [93]. Renal sympathetic
nerves release renin (threshold, 0.5 Hz; β1-adrenoceptor-mediated), increase tubular
transport (threshold, 1 Hz; α1-adrenoceptor-mediated), and constrict the renal vas-
culature (threshold, 2.5 Hz; α1-adrenoceptor-mediated). Autonomic control of the
kidney contributes to blood volume restoration following a positive or negative per-
turbation in volume status and helps balance the work load between the two kid-
neys. Renal efferent sympathetic nerve activity impairs sodium excretion and shifts
the renal pressure–natriuresis curve to the right such that higher long-term levels of
BP are required to maintain sodium excretion in balance with sodium intake [6].
In humans, the AVP involved in the water–salt balance is released from the neu-
rohypophysis as several peaks overnight, characterized by an increase of 100–300%
of the levels of the hormone in plasma. These peaks are of short duration, consistent
with the short half-life of plasma AVP. No relationship was found between these
episodes of AVP release and plasma sodium level, which remains constant through-
out the night. Furthermore, no relationship was found between sleep stages and the
levels of AVP. The activity and posture are all factors that can influence the daily
AVP variations. However, in healthy subjects with constant recumbent position and
fluid intake, a daily pattern of AVP remained, albeit with lower amplitude, proving
the endogenous character of the rhythmicity (Fig. 5.40) [87].
One possible explanation for the nocturnal rise in AVP could be the decrease in BP
pressure at these times. In this sense, there is a circadian pattern of circulating blood
24-h Rhythms in Plasma Osmolality and the Intravascular Volume 225

Decreased filltered β2-globulin


sodium load (angiotensinogen)

Decreased Juxta-
perfusion pressure glomerular Renin
apparatus
+ 10 AA Angiotensin I

β-adrenergic
+ Converting enzyme
terminals
Angiotensin II
Atrial natriuretic peptide (ANP)
8 AA Angiotensin II

CH2OH
+
O C O
HO CH

Vasocontriction

+
CNS
O Aldosterone ANP
Somatostatin
ACTH Dopamine Thirst
MSH ADH secretion
hypokalemia

Fig. 5.39  The renin–angiotensin–aldosterone system with the different neural and endocrine
interactions

volume, high during the day (12:00–18:00 h), which drops significantly (6%) at night
and increases again in the morning (06:00 h). This circadian variation of blood vol-
ume could add to the BP changes to cause the increased release of AVP at night.
The amplitude of the AVP circadian rhythm declines with age. This in part
explains the shift in the peak time of urine production and volume from afternoon/
early evening in young persons to 00:00–0002.00 h in the elderly, commonly asso-
ciated with nocturia. Another important cause of nocturia in the elderly is abnormal
sleep-time BP decline, manifesting as nondipping 24-h patterning, and/or sleep-­
time hypertension [87].
The urine volume and electrolyte secretion is usually lower during sleep. REM
sleep is associated with a decreased flow of urine and increased osmolarity. The
activity of plasma renin and angiotensin levels and aldosterone are elevated during
sleep (Figs. 5.40 and 5.41) [87]. ANP is elevated in untreated sleep apnea (by atrial
impact of changes of negative intrathoracic pressure), resulting in nocturia and
natriuresis.
These effects are mainly related to sleep rather than circadian factors. Thus,
sleep deprivation inhibits the nocturnal increase in renin and aldosterone (Fig. 5.40).
Renin increases during NREM sleep and decreases during REM sleep. In the transi-
tion from REM sleep to NREM sleep, renin levels rise [87].
226 5  Third Level: The Hypothalamus

DAY 1 DAY 2
2.0 100

ANP (pg/mL plasma)


AVP (pg/mLplasma)

80

1.5 60

40

1.0 20

Sleep Sleep
deprivation
5
Angiotensin (pg/mL plasma)

1
18 22 02 06 10 14 18 22 02 06 10 14 18
Time of day (h)

Fig. 5.40  Arginine vasopressin (AVP) and atrial natriuretic peptide (ANP) are hormones whose
constitutive secretion is controlled by the C process (upper panel). Twenty-four hour rhythm in
angiotensin production is dependent on the S process (lower panel). Ultradian variations in hor-
mone secretion or values <1 pg/mL plasma (AVP, angiotensin) or <20 pg/mL plasma (ANP) are
not represented. Data from Kamperis et al. [87]

The circadian rhythm of renal Na+, K+, and H2O management is driven to a large
extent by the circadian rhythm of aldosterone [86]. Nyctohemeral fluctuation of
ANP and AVP, which both peak early during sleep, also modulates the 24-h rhythm
of urinary Na excretion. The relationship between BP and natriuresis is controlled
during the daytime by upright posture and activity; thus, it is mainly during the
nighttime when Na+ sensitivity, which is present to varying extents among all per-
sons, most strongly exerts corrective effects. Thus, in acute and chronic situations
when Na+ intake is excessive or its daytime elimination compromised, the innate
pressure natriuresis mechanism adjusts BP to an elevated level during nocturnal rest
as a compensatory response, giving rise to abnormally elevated sleep-time BP, i.e.,
a nondipping 24-h pattern and/or nocturnal hypertension. This in turn promotes
blood volume reduction through enhanced overnight natriuresis and diuresis [86].
Atrial natriuretic peptide belongs to a family of peptides that includes three other
members, brain natriuretic peptide (BNP), C-type natriuretic peptide (CNP), and
urodilatin, all encoded by different and independent genes. These peptides are pres-
ent in mammals as potent protective agents against volume overload. ANP and BNP
are produced mainly in the heart, but they can be found, together with their receptor
24-h Rhythms in Plasma Osmolality and the Intravascular Volume 227

DAY 1 DAY 2
50
Aldosterone (ng/dL plasma)

40

30

20

10
Sleep Sleep
deprivation
50
Renin (ng/mL plasma/h)

40

30

20

10
18 22 02 06 10 14 18 22 02 06 10 14 18
Time of day (h)

Fig. 5.41  The constitutive secretion of renin and aldosterone is dependent on the S process.
Ultradian variations in hormone secretion or values <10 ng/dL plasma (aldosterone) or <10 ng/mL
plasma (renin) are not represented. Data from Kamperis et al. [87]

type A, in hypothalamic and brainstem areas involved in the regulation of body fluid
volume and BP [94].
Expansion of blood volume acts directly on the heart by stretching cardiomyo-
cytes, increasing the release of ANP, and leading to an effective reduction of circu-
lating blood volume. In addition, there is strong evidence that hypothalamic ANP is
also released after the expansion of blood volume because of the increased activity
of afferent pathways stimulated by baroreceptors.
ANP produced on a large scale in the heart reduces the strength of the heart con-
traction and the frequency of heart beats, causing relaxation in the vessels, decreased
peripheral resistance, diuresis, natriuresis, and inhibition of salt and water intake
[94]. Sleep deprivation markedly increased the diuresis and led to excess renal
sodium excretion. Renal water handling and AVP and ANP levels remained unaltered
during sleep deprivation, but the circadian rhythm of the hormones of the renin–
angiotensin–aldosterone system is significantly affected (Figs. 5.40 and 5.41).
Hemodynamic changes were characterized by the attenuation of nocturnal BP dip-
ping and an increase in creatinine clearance. Acute deprivation of sleep induces
228 5  Third Level: The Hypothalamus

natriuresis and osmotic diuresis, leading to excess urine production. The amount of
urine produced during these sleepless nights by far exceeds bladder reservoir ability,
thus leading to nocturia. Enuresis in children and nocturia in the elderly is in many
cases the result of excess nocturnal urine production.
The nocturnal levels of plasma renin, angiotensin II, and aldosterone are sup-
pressed during sleep deprivation, directly leading to reduced sodium reabsorption in
renal tubuli (Figs. 5.40 and 5.41) [87]. Suppression of the rhythm of renin–angioten-
sin activity may be the result of a direct effect of sleep deprivation on the sensitivity
of the renin–angiotensin system or be mediated through sympathetic–parasympa-
thetic system imbalance.
Nocturnal levels of ANP are unaltered during sleep deprivation (Fig. 5.40), despite
the attenuation of nocturnal BP dipping, indicating that they are not involved in sleep
deprivation-induced natriuresis [87, 95]. In clinical settings related to disturbed sleep
due to sleep apnea, ANP plays a pivotal role in the natriuresis and polyuria observed.
Under basal conditions, there is a time-dependent sleep–wake cycle relationship
between the renin–angiotensin system and the adrenocortical system for the release of
aldosterone. Increasing the amplitude of the pulses of aldosterone observed upon awak-
ening is attributed to increased activity of the corticotropic–adrenal axis reflected by the
large increase in cortisol at that time. Most of the aldosterone release pulses occurring at
the end of sleep, are synchronous with cortisol. The permanence of high levels of cortisol
during the period 07:00–15:00 h, could explain the high values of aldosterone at this time,
regardless of whether the subject is asleep or awake. By contrast, aldosterone pulses dur-
ing sleep periods are primarily related to major fluctuations in plasma renin [84].
DA is synthesized in the renal proximal tubule cells from filtered L-Dopa and
secreted into the renal proximal tubule, where it binds to D1 receptors inhibiting Na+
reabsorption [96]. Renal DA acts as a paracrine substance that opposes the actions
of angiotensin to increase Na+ reabsorption via AT1 receptors. DA affects renin
release from renal juxtaglomerular cells via the D1-like receptor family. Renal DA
serves as one of several paracrine mediators in renal Na+ excretion.
A reliable humoral biological marker of REM and NREM cycles is renin secre-
tion. Their oscillations are strongly attached to the REM and NREM sleep cycles.
NREM sleep coincides with an increase in plasma renin, whereas it decreases dur-
ing REM sleep. Therefore, nights deprived of sleep are characterized by natriuresis,
osmotic diuresis, and a dramatic increase in urine output. BP dipping is attenuated,
and the renin–angiotensin system is clearly suppressed. It is of importance to evalu-
ate sleep architecture and its disturbances in clinical settings with nocturnal poly-
uria and natriuresis, such as enuresis in children and nocturia in the elderly.

24-h Rhythms in Body Temperature Control

Humans maintain core temperature within a few tenths of a degree of 37 °C over a
very wide range of environmental exposures and activity levels. During hyperther-
mia, heat dissipation occurs primarily via sympathetically mediated sweating and
cutaneous vasodilation. During cold exposure, sympathetic cutaneous vasoconstric-
tion helps to decrease heat dissipation; shivering increases heat generation when
cooling is more extreme [97].
24-h Rhythms in Body Temperature Control 229

Centrally, thermoregulation is controlled at the preoptic/anterior hypothalamus,


which acts as a thermostat by integrating central (brain) temperature with thermal
information from peripheral afferents. Changes in body temperature thus elicit
reflex responses for heat dissipation, heat conservation, and/or heat generation, as
appropriate. These minimize changes in body temperature in classical negative
feedback fashion and help to prevent the attainment of potentially dangerous body
temperatures at either extreme [98]. Physiological thermoregulatory responses to
heat in humans are much more efficient than responses to cold. Therefore, physio-
logical thermoregulation is primary during hyperthermia.
Body temperature varies with the anatomical regions (e.g., axillary: 36.7 °C;
scrotal: 32 °C; rectal: 37.2 °C) and has 24-h rhythm directly linked to the activity of
the SCN, with a minimum in the middle of the night and a maximum toward the end
of the day. The amplitude of this variation is about 0.6 °C (Fig. 5.42) [1].

t (C)
Skin temperature


Suprachiasmatic 0,6 C
Nucleus

Hair

+
Blood temperature Set point Cytokines
~ 37 C Bacterial pyrogens

Preoptic DMH
Hypothalamus Area

Shivering
TRH ANS Motor system thermogenesis
Changes in
Non-shivering
posture
TSH thermogenesis

Sweating Motor
T3-T4 behavior
Vasodilation

Metabolic Vasoconstriction
rate

Piloerection

Fig. 5.42  Components of the temperature controlling mechanism


230 5  Third Level: The Hypothalamus

In two of the three physiological configurations in which human life elapses,


i.e., wakefulness and NREM sleep, body temperature is regulated precisely. By
contrast, in REM sleep, a transient disconnection of the autonomic regulatory
mechanisms at the supraspinal level occurs, leading to a state of poikilothermia,
that is, the central temperature tends to adapt to the ambient temperature. As an
individual 75 years of age spends about 6 years in REM sleep, it can be said that
we are not warm-blooded animals for an considerable part of our life. The pre-
cise mechanisms that we discuss are operational only during NREM sleep and
wakefulness.
Temperature is a critical variable in health and disease. The constraint of human
body core temperature within a degree or two of 37 °C, which is the optimal tem-
perature for normal cellular function, has three causes. The first is the stable cli-
mate, which maintains temperatures across most of the surface of planet Earth
within a range compatible with human life. The second cause is the ANS, which
reacts robustly to thermal challenges by orchestrating a complex array of neural
responses below the level of conscious awareness. The autonomic responses to cold
stress include cutaneous vasoconstriction to retain bodily heat along with metabolic
and shivering thermogenesis [99]. The autonomic responses to heat stress include
cutaneous vasodilatation, which liberates heat by radiant and convective heat loss,
and sweating, which liberates heat by evaporation. The third and least predictable
cause is human behavior, which responds to thermal sensory input by seeking
warmth or coolness (Figs. 5.42 and 5.43).

Controlled systems
Heat
production
Central control
signal
Vasomotor
system

Sweating

Controller
Σ Behavior
Hypothalamus

Signal error
Σ
Sensors cold, warm
Feedback signal
Central
From thermoreceptors thermoreceptor
Controlled
Σ variable

Peripheral Body temperature


thermoreceptor

Reference signal

Fig. 5.43  Analysis of hypothalamic temperature control according to the systems control theory.
Modified with permission from Cardinali [1]
24-h Rhythms in Body Temperature Control 231

Body temperature is controlled by central and peripheral adjustment and heat


dissipation [97]. The analysis of body temperature regulation as a control theory is
shown in Fig. 5.43. Maintaining core temperature is in the first instance given by
vasomotor changes and only when this range is exceeded do mechanisms of heat
production or dissipation start.
Heat production (thermogenesis) is under neural control. There are two types of
thermogenesis, shivering and nonshivering. Shivering thermogenesis is induced by
the caudal hypothalamus via projections to the brainstem nuclei of the somatic
motor system. Nonshivering thermogenesis is controlled by the ANS via sympa-
thetic β-adrenoceptor innervation of lipid deposits and involves heat production by
a particular form of adipose tissue, the brown adipose tissue (BAT). The thermo-
genic potential of BAT is due to the presence of uncoupling protein 1, a protein
uniquely found in the inner membrane of the brown adipocytes’ mitochondria. The
protein uncouples substrate oxidation from electron transport. In humans, BAT is
found in subscapular, cervical, perispinal, mediastinal, periaortic, pericardial, and
periadrenal regions [100, 101].
The sympathetic system is also responsible for vasomotor changes, which
include changes in perfusion of the limbs and trunk, and sweat secretion by cholin-
ergic terminals in the corresponding territory. Behavioral changes (looking for shel-
ter, etc.) also contribute as an effector control system. Only in newborns are TRH
release and subsequent activation of the pituitary–thyroid axis significant [99].
In addition to the cold-activated pathways (e.g., the thermoregulatory pathway),
several hypothalamic nuclei have been linked to the control of BAT thermogenesis
to allow diet-induced thermogenesis (e.g., the energy homeostasis regulatory path-
way; Fig. 5.44). These nuclei include the ARC, preoptic area (POA), DMH, para-
ventricular hypothalamus, LHA, and VMH [97, 102].
Regarding temperature receptors involved in thermoregulatory behavior a dual
mechanism exists, i.e., peripheral and central thermoreceptors. There are two types
of peripheral thermoreceptors: (a) cold, with optimal activation from 10° to 3 °C; (b)
heat, with activation above 45 °C. They are free terminations of sensory fibers such
as nociceptive A δ and C fibers for cold and C fibers for heat. Temperature-sensing
receptors belong to the superfamily of transient receptor potential (TRP) channels,
which are located in the nerve endings of sensory cells throughout the skin [99]. The
interoceptive aspects of behavioral thermoregulation have been emphasized, includ-
ing the primary importance of skin temperature, the concept of thermal discomfort,
and the important contribution of orbitofrontal, insular, somatosensory, and amyg-
dala cortical regions deployed to anticipate and avoid thermal stress.
Central thermoreceptors are located at various levels of the CNS, including the
spinal cord, reticular formation, and the POA itself. Most thermoreceptors in the
spinal cord are heat-sensitive neurons (50–70% of neurons), whereas only 2% of
them are sensitive to cold. These neurons are nonspecific, and respond to changes in
osmolarity, BP, glucose, and sexual steroids (Fig. 5.44).
The signals from the peripheral receptors enter the CNS through the dorsal root
of the spinal cord and send their information to the upper levels of thermal integra-
tion by the contralateral paleospinothalamic tract (Fig. 5.44). Projections of
232 5  Third Level: The Hypothalamus

rRPa

P
POA CVC

PGE2
P
BAT
WS
MPA CVC DMH

MnPO WS P
BAT shiver

WS
shiver
warm
LPB

cold

Fig. 5.44  Thermosensory signals driving thermoregulatory responses are transmitted from the
lateral parabrachial nucleus (LPB) to the preoptic area (POA), which contains the microcircuitry
through which cutaneous and core thermal signals are integrated to regulate the balance of POA
outputs that are excitatory (dashed green) and inhibitory (dashed red) to thermogenesis-promoting
neurons in the DMH. Other excitatory median preoptic area (MnPOA) neurons (dashed brown)
either project to the cutaneous vasoconstriction (CVC) sympathetic premotor neurons in the rostral
raphe pallidus (rRPa) or to DMH. Within the POA, GABAergic interneurons (red) in the MnPO
subnucleus receive glutamatergic inputs from skin-cooling-activated neurons in the LPB and
inhibit each of the distinct populations of warm-sensitive (W-S) neurons in the medial preoptic
area (MPA) that control CVC, BAT, and shivering. Prostaglandin E2 binds to receptors to inhibit
the activity of each of the classes of W-S neurons in the POA. Redrawn from Morrison [97]. The
figure was prepared in part using image vectors from Servier Medical Art (www.servier.com),
licensed under the Creative Commons Attribution 3.0 Unported License (http://creativecommons.
org/license/by/3.0/)

peripheral thermoreceptors from different areas of the skin converge on the PBN at
the brainstem from where it projects to the POA of the anterior hypothalamus via
the medial forebrain bundle and periventricular striatum (Fig. 5.44).
There is a hierarchy of structures involved in thermoregulation, extending
from the POA to the brainstem and spinal cord [97]. Tetraplegic patients who
have suffered transection at the cervical level of the spinal cord can maintain
their body temperature at around 37 °C, although present thermoregulatory
instability when subjected to rapid changes in ambient temperature. In situa-
tions where the POA and the anterior hypothalamus are lesioned, fever in
response to pyrogens is abolished.
Hypothalamic heating produces panting in anesthetized cats, and abolishes
thermogenesis and produces peripheral vasodilation in dogs. All these mechanisms
lead to a reduction in body temperature. The POA region is an important site for
thermoregulation, being situated on top of the hierarchy for neuronal regulation of
body temperature [97].
Figure 5.44 summarizes the neuronal pathways involved in thermoregulation.
Retrograde transneuronal viral tracing has been of paramount importance in
24-h Rhythms in Body Temperature Control 233

delineating the brain regions and the neuronal circuits connected to BAT and
white adipose tissue (WAT). POA output is excitatory or inhibitory to thermogen-
esis-promoting neurons in the DMH and to cutaneous vasoconstriction sympa-
thetic premotor neurons in the rostral raphe pallidus (rRPa). As already mentioned,
the DMH plays a wide range of metabolic and behavioral roles, including body
weight regulation [66].
The BAT and WAT send feedback information to the CNS via sensory nerves
that connect adipocytes via dorsal root ganglia with the brain [103–105]. Incoming
(afferent) sympathetic nerves can be distinguished from outgoing (efferent) sensory
nerves with multisynaptic anterograde (Herpes virus) and retrograde (pseudorabies
virus) viral tracers that are injected into the BAT or WAT. Many CNS sites showed
both sympathetic and sensory connectivity; thus, an extensive feedback system for
incoming and outgoing signals is likely.
Thermogenesis of BAT occurs as part of the basic rest–activity cycle (BRAC)
and contributes to increases in body and brain temperature [105]. With ad libitum
food, eating begins 15 min after the onset of BAT thermogenesis in rats. The initia-
tion of eating is centrally programmed, and is a component of the BRAC. This
increase in brain temperature that precedes eating may facilitate the cognitive pro-
cessing that occurs during the search for food, when the rat engages with the exter-
nal environment. Rather than being triggered by changes in levels of body fuels or
other meal-associated factors, in sedentary laboratory rats with ad libitum access to
food, meal initiation normally occurs as part of the centrally programmed ultradian
BRAC. BRAC-associated BAT temperature increases occur in a thermoneutral
environment; thus, they are not preceded by falls in body or brain temperature as a
homeostatic thermoregulatory response [105].
Rats with hypothyroidism, despite having the normal diet-induced thermogene-
sis, are unable to survive in cold environments owing to a reduced capacity to pro-
duce heat without trembling. In addition, they also fail to increase thermogenesis in
BAT in response to the infusion of NE. As discussed in Chap. 4, the SCG innerva-
tion of the thyroid gland plays a substantial role in the normal adaptive response to
cold stress.
Tremor consisting of involuntary rhythmic movements of the skeletal muscles is
seen in response to exposure to cold, without any change in body position. As there
is no mechanical work, almost all the energy is released as heat [106]. Almost all
body muscles participate in this response, except the middle ear, facial, perineal,
and extraocular muscles. Premotor neurons in the rRPa play a substantial role in the
efferent control of shivering thermogenesis. These, in turn, project to motoneurons
in the ventral horn. The spinal cord seems to contain the basic mechanisms for these
movements to occur, because cooling causes tremor in spinal animals. The ANS
exerts a fine control of shivering, e.g., NE increases the sensitivity of skeletal mus-
cle fiber to acetylcholine (ACh).
Concerning the cutaneous blood flow, the increase or reduction of skin blood
flow, especially in the hands, feet, lips, ears, and nose, can facilitate or hinder the
loss of heat to the environment respectively. Exposure to cold causes cutaneous
vasoconstriction via NE effects on α1-adrenoceptors in arterioles and arteriovenous
234 5  Third Level: The Hypothalamus

anastomoses, which supply the venous plexus of the skin. On the other hand, during
exposure to heat, vasodilation in these regions is largely the consequence of vaso-
constrictor activity removal [99].
The importance of evaporative heat loss becomes larger as the environmental
heat load increases, and is the only means of heat loss when the ambient tempera-
ture rises above body temperature. Cooling by evaporation can be achieved in dif-
ferent ways, depending on the species: sweating in humans and cattle; panting in the
dog, the sheep, and lizard; or salivation and licking in rats and kangaroo. The sweat
glands respond to heat stress by sympathetic cholinergic stimulation (secretion is
blocked by atropine), although they also have detectable adrenoceptors. The most
potent stimulus for inducing sweat is the increase in body temperature; however,
there is a modulation of the average skin temperature of the whole body and the
location of the response.
Pyrogen production during bacterial infection, e.g., cytokines such as IL-1,
affects the CNS in areas outside the blood–brain barrier, with increased “set-point”
temperature. The absence of estrogen in perimenopausal women produces the typi-
cal hot flashes, because the thermostable zone is narrowed in the absence of estro-
gen (Fig. 5.45).

Increased heat Peripheral Increased


production vasomotor heat loss
tone
Intensity of effect

Central temperature (C)

Endogenous Antipyretics:
pyrogens: IL-1, GC,alpha MSH,
IL-6, PGE2 AVP

Fig. 5.45  Pyrogens move the equilibrium point of the system to the right, with increasing central
temperature. Endogenous antipyretics (AVP; glucocorticoids, GC; α-MSH) or pharmacological
agents re-establish the equilibrium point. The thermostable zone (light blue) is the central tempera-
ture range in which changes in central temperature occur without neurovegetative impact.
Perimenopausal lack of estrogen reduces this thermostable zone (dark blue); thus, minimal
changes in central temperature produce ANS symptoms (“hot flashes”). Modified with permission
from Cardinali [1]
Sexual and Maternal Behavior 235

Sexual and Maternal Behavior

In the hypothalamus, there are neural command groups for sexual and maternal
behaviors. For example, the same group of GnRH neurons that regulates the release
of pituitary gonadotropins, and thus the central event of the sexual cycle, project to
the limbic system, which regulates sexual behavior. Hence, the same group of com-
mand neurons is regulating the various components, endocrine (pituitary and
gonadal hormone release), regional (erection, orgasm), and motivational (libido) of
sexual behavior. A comparable case is that of oxytocin, which acts as a hormone in
lactation (milk ejection reflex) and as a transmitter in the limbic pathways to induce
maternal behavior.
Sexual function requires interaction of multiple levels and areas of the central
and peripheral somatic and ANS. The medial preoptic area (mPOA), the ventral
medullary reticular formation, the nucleus paragigantocellularis in the ventral
medulla, the periaqueductal gray, the mesencephalic ventral tegmental area, neu-
rons in the central tegmental mesencephalic region, and the medial amygdala are
the main central areas involved in arousal and sexual responses [107]. Central neu-
rotransmitters include DA (enhances sexual desire, arousal), NE (gates sensory
input from the genitalia and maintains sexual arousal), ACh (mediates lubrication
and vaginal engorgement), His, 5-HT (diminishes excitatory effects, interferes with
arousal and orgasm), PRL (causes sexual satiety and post-orgasmic relief), and oxy-
tocin (promotes sexual receptivity and bonding). Peripheral sympathetic and para-
sympathetic pathways are similar in men and women.
Hormones influence female sexual function. Progesterone furthers partner recep-
tivity, and estrogens enhance desire, arousal, sensory thresholds, and genital arterial
blood flow. Testosterone helps to initiate sexual activity [107].
Gonadotropic axis rhythms cover a wide range of frequencies, whose interaction
provides a temporary program coordinated for the reproductive axis at each stage of
maturation. The secretion of gonadotropins occurs in erratic pulses in the absence
of hypothalamic regulation (Fig. 5.4). Thus, the pulse frequency of GnRH regulates
the expression of GnRH receptors in the pituitary [12].
Neurons secreting GnRH are located within the mPOA and serve as the final
output pathway regulating the LH and FSH surge [108]. GnRH neurons send axons
to the median eminence of the hypothalamus where they release GnRH into the
pituitary portal system, thereby triggering LH secretion and ovulation (Fig. 5.6).
Females, but not males, can produce an LH surge even though there is no sex differ-
ence in the GnRH neurons themselves. The difference in ability to generate a GnRH/
LH surge is believed to be upstream of the GnRH neurons and is the result of orga-
nizational processes shaped by gonadal steroid exposure during neonatal develop-
ment. The circadian timing system (by acting on sexually differentiated neurons of
the AVPV) regulates the dynamics of the neural circuits leading to the rhythmic
generation of the GnRH/LH surge and ultimately ovulation (Fig. 5.6).
Studies in patients with defective hypothalamic GnRH secretion confirmed the
close correlation between the pulsatile LH response and exogenous GnRH pulses.
The synthesis and secretion of LH and FSH are differentially regulated by the
236 5  Third Level: The Hypothalamus

frequency of the pulses of GnRH. Thus, a low frequency of GnRH pulses preferen-


tially stimulates FSH, whereas more frequent GnRH pulses are needed for the opti-
mal stimulation of LH [108].
The AVPV) sits upstream of GnRH neurons and is characterized by multiple sex
differences (Fig. 5.6). Lesions of the AVPV prevent spontaneous and steroid-­
induced preovulatory surges of LH. Females have more AVPV neurons than males,
and a denser projection from the AVPV) to the GnRH neurons. One sex difference
that has been unequivocally implicated in the generation of the LH surge is the
expression of the kisspeptin gene in AVPV neurons [14, 109]. Females have up to
25 times more AVPV kisspeptin-expressing) neurons than males. Moreover, both
AVPV KiSS1 and C-FOS expression levels increase during the preovulatory LH
surge. Treatment with estradiol produces a surge of LH in females but not males,
and central infusion of kisspeptin receptor (Kiss1 R) antagonist into the mPOA
blocks the estradiol-induced LH surge and estrous rhythmicity [14, 109].
The sex difference in AVPV kisspeptin) results from exposure to hormones at criti-
cal times during development, suggesting that hormones might have organizational
effects on the AVPV in early development. Females treated postnatally with testoster-
one or estradiol have fewer kisspeptin neurons than untreated females and they are
unable to produce the estradiol-induced LH surge in adulthood. Further, males who
undergo castration shortly after birth have greater numbers of AVPV )kisspeptin neu-
rons than intact males, and they can produce an estradiol-induced surge of LH [14].
A subset of neurons within the DMH express gonadotropin-inhibiting hormone
(GnIH), a neuropeptide that has been shown to act as an inhibitor of the reproduc-
tive axis (Fig. 5.6) [110]. These neurons project directly to the GnRH neurons of the
mPOA. In all mammals studied to date, GnIH rapidly suppresses LH release. The
SCN sends projections to a large portion of GnIH cells, suggesting the circadian
regulation of this population of neurons [14].
The effects of the circadian sleep schedule and secretion of LH and FSH are seen
more clearly during puberty than in adulthood [111]. Before puberty, gonadotropins are
secreted in pulses of low amplitude throughout the day and night. When children of both
sexes approach puberty, the amplitude of night pulses gradually increases, and a daily
hormonal rhythm becomes evident, with high levels of LH and FSH overnight [111].
Deep sleep, rather than REM or lighter sleep stages, provides a critical stimulus
for LH secretion in puberty. The causal relationship between deep sleep and GnRH
pulse generation is supported by neuroanatomical evidence for a direct synaptic
connection between VLPO and GnRH neurons. The absence of GnRH axonal fibers
in the VLPO is compatible with the hypothesis that sleep stimulates GnRH secre-
tion, rather than the vice versa.
Going from puberty to adulthood, the amplitude of the pulses in daylight hours
increases, thus eliminating the daily rhythm of LH and FSH. In girls at the age of 16
years, a characteristic LH pattern of adult women is established at the beginning of
the follicular phase [8].
In studies during puberty in males, in which a change of 12 h in the sleep–wake
cycle was imposed, it was shown that elevated night levels of LH, partly represent a
stimulating effect of sleep itself. However, after reversing the hours of sleep,
Sexual and Maternal Behavior 237

increased LH release takes place not only during the daytime sleep, but also the dur-
ing the hours when sleep should take place under normal conditions, indicating the
influence of circadian rhythmicity in the temporal pattern of LH release. The role of
the circadian control in the nocturnal LH surge is also evident during sleep depriva-
tion, maintaining that elevation, but on a smaller scale (Fig. 5.46) [9, 112].
In women, daily LH pulsatility changes are subject to complex modulation by
the menstrual cycle. At the beginning of the follicular phase, LH pulses are large
and infrequent, the night period being associated with a reduction in the frequency
of the pulses; thus, the average levels of LH decrease during sleep. In the middle of
the follicular phase, the pulse amplitude is decreased and the effects of sleep on
modulating the frequency of the pulses is less noticeable. At the end of the follicular
phase pulse amplitude increases, but sleep modulation is not present until the begin-
ning of the luteal phase, resulting night pulsatility becoming slower [113, 114]. It is
not known whether night variation in LH pulsatility in women is dependent on the
sleep–wake cycle and/or the circadian pacemaker cycle.

DAY 1 DAY 2
25

20

15

10
LH (mU/mL plasma)

5
Sleep Sleep Diurnal
deprivation sleep
25

20 Puberal males
Adult males
15 Prepuberal males

10

5
18 22 02 06 10 14 18 22 04 08 12 16
Time of day (h)

Fig. 5.46  Daily variations of plasma LH concentration in normal prepubertal, pubertal, and adult
male subjects. Only puberal males showed nocturnal maxima that were partially impaired by sleep
deprivation with a shift to diurnal sleep on the second day. The ultradian variations and values
<5 mU/mL are not represented. Data from Van Cauter and Refetoff and Brabant et al. [9, 112]
238 5  Third Level: The Hypothalamus

An interaction between the menstrual cycle and circadian rhythmicity also


appears to be involved in the temporal pattern of pre-ovulatory LH release, which
occurs most often at the end of sleep or early in the morning. Toward the meno-
pause, levels of gonadotropins increase their pulsatility, giving rise to high values
without a consistent daily variation. In postmenopausal women, significant gonado-
tropin pulses, possibly reflecting pituitary secretion, are associated with large pulses
of LH and FSH [115].
During puberty in children, night testosterone elevation coincides with increased
serum levels of gonadotropins. In girls, puberty coincides with a circadian variation,
with higher estradiol levels during the day than at night. Testosterone secretion is
also pulsatile in adults, with 17–18 pulses detected per day, approximately coincid-
ing with LH. A daily rhythm is also observed, with an amplitude of 25% of the
mean value. It presents a night increase, starting shortly after midnight, reaching the
maximum value at the beginning of the morning (08:00 h), and minimum values
between 19:00 and 21:00 h (Fig. 5.47) [112].
As the secretory daily patterns of LH in adults are highly variable and do not
show a clear nocturnal rise, in contrast to the stable testosterone circadian rhythm,
it seems clear that other factors are involved. Ovarian innervation, which has a clear
regulatory function, can play an important role in this phenomenon (Chap. 4).
In adults, the modulatory role of sleep and circadian rhythmicity in gonadotropic
function is unclear. The nocturnal increase seems to be independent of sleep, a fact
that supports the hypothesis of an intrinsic circadian rhythmicity of pituitary–
gonadal activity. In adults, a positive correlation between the number of REM epi-
sodes and night average levels of testosterone occurs. With age, the frequencies of
the LH pulses and testosterone decrease, canceling out most circadian rhythmicity.
The morning plasma concentration of testosterone in elderly men correlated with
the amount of nighttime sleep measured by PSG [116].
There is evidence for the relationship between the reproductive cycle and the
secretion of oxytocin. In animal experiments, it was found that estrogens increase

1000
ng of testosterone/mL plasma

750

500

250
Fig. 5.47 Plasma
testosterone levels in
normal adult males. 0
Redrawn from Brabant 00:00 06:00 12:00 18:00 24:00
et al. [112] Time of day
References 239

the release of oxytocin, whereas progesterone inhibits it. Studies in women indicate
that oxytocin levels vary throughout the menstrual cycle, characterized by a peak
associated with ovulation. Although in the middle of the cycle, levels of oxytocin in
the corpora lutea are far superior to those of the peripheral circulation, the data sug-
gest that during the follicular phase, this hormone comes mainly from the pituitary
gland and not from the ovaries. Moreover, in animal studies, it was found that the
response to gonadal steroids creates a yearly cycle of oxytocin, which matches the
corresponding seasonal periodicity of the estrous cycle [117].
In oxytocin-releasing neurons, there is evidence for the relationship between
synchronous electrical activity and the pulsatile release of the hormone [118]. There
is a positive correlation between the frequency of the action potential of hypotha-
lamic neurons and the amount of oxytocin released from their axon terminals
located in the posterior lobe of the pituitary. The pulsatile release of oxytocin, with
rapid release and short half-life, often renders plasma levels of this hormone unde-
tectable, even during stimulated secretion.
The transformation from a nonparental to a maternal state involves several dramatic
and wide-ranging alterations, including changes in the CNS, behavior, and physiology.
An interplay between the neuroendocrine system, including estradiol, progesterone,
and PRL, and CNS neuromodulators, including oxytocin, DA, and AVP, helps to
orchestrate multiple maternal functions [117]. Although hormonal changes occurring
throughout pregnancy and at the time of parturition have been demonstrated to prime
the maternal brain and trigger the onset of mother–infant interactions, extended experi-
ence with neonates can induce similar behavioral interactions [119].
Oxytocin is known to facilitate maternal behavior in many species. For example,
knockout and pharmacological studies in mice suggested that oxytocin might facili-
tate maternal behavior, whereas a reduction in oxytocin function promoted infanti-
cidal behavior. However, those approaches often produced global increases or
decreases in oxytocin function, affecting multiple brain sites, and likely multiple
oxytocin functions [118, 120].
Prolactin is known for its role in promoting maternal behavior in mammals and
in promoting parental care in males and females of biparental bird species [121].
PRL regulates the onset of maternal behavior but not the maintenance, which is
controlled by pup exposure. The role of PRL in mammalian paternal behavior is
subtler and possibly more species-specific than for some other hormones. It has
been suggested that PRL might be involved in the transition from a nonpaternal to a
paternal state in male mammals [121].

References
1. Cardinali DP. Neurociencia Aplicada. Sus Fundamentos. Buenos Aires: Editorial Médica
Panamericana; 2007.
2. Stuber GD, Wise RA. Lateral hypothalamic circuits for feeding and reward. Nat Neurosci.
2016;19:198–205.
3. Takahashi C, Hinson HE, Baguley IJ. Autonomic dysfunction syndromes after acute brain
injury. Handb Clin Neurol. 2015;128:539–51.
240 5  Third Level: The Hypothalamus

4. Wehrwein EA, Joyner MJ. Chapter 8 – Regulation of blood pressure by the arterial barore-
flex and autonomic nervous system. In: Buijs RM, Swaab D, editors. Handbook of Clinical
Neurology. Autonomic Nervous System. New York: Elsevier; 2013. p. 89–102.
5. James C, Macefield VG, Henderson LA. Real-time imaging of cortical and subcortical control
of muscle sympathetic nerve activity in awake human subjects. Neuroimage. 2013;70:59–65.
6. Zheng H, Patel KP. Integration of renal sensory afferents at the level of the paraventricular
nucleus dictating sympathetic outflow. Auton Neurosci. 2016;204:57–64.
7. Shenton FC, Pyner S. Vagal afferents, sympathetic efferents and the role of the PVN in heart
failure. Auton Neurosci. 2016;199:38–47.
8. Cardinali DP. In: Pandi-Perumal SR, editor. Neuroendocrine correlates of sleep/wakefulness.
New York: Springer; 2006.
9. Van Cauter E, Refetoff S. Multifactorial control of the 24-hour secretory profiles of pituitary
hormones. J Endocrinol Invest. 1985;8:381–91.
10. Copinschi G, Spiegel K, Leproult R, Van CE. Pathophysiology of human circadian rhythms.
Novartis Found Symp. 2000;227:143–57.
11. Herbison AE. Control of puberty onset and fertility by gonadotropin-releasing hormone neu-
rons. Nat Rev Endocrinol. 2016;12:452–66.
12. Maggi R, Cariboni AM, Marelli MM, Moretti RM, Andre V, Marzagalli M, Limonta P. GnRH
and GnRH receptors in the pathophysiology of the human female reproductive system. Hum
Reprod Update. 2016;22.
13. Clarke SA, Dhillo WS. Kisspeptin across the human lifespan:evidence from animal studies
and beyond. J Endocrinol. 2016;229:R83–98.
14. Putteeraj M, Soga T, Ubuka T, Parhar IS. A “timed” kiss Is essential for reproduction: lessons
from mammalian studies. Front Endocrinol (Lausanne). 2016;7:121.
15. Steyn FJ, Tolle V, Chen C, Epelbaum J. Neuroendocrine regulation of growth hormone secre-
tion. Compr Physiol. 2016;6:687–735.
16. Copinschi G, Van OA, L’Hermite-Baleriaux M, Szyper M, Caufriez A, Bosson D, L’Hermite
M, Robyn C, Turek FW, Van CE. Effects of the short-acting benzodiazepine triazolam,
taken at bedtime, on circadian and sleep-related hormonal profiles in normal men. Sleep.
1990;13:232–44.
17. Spiegel K, Leproult R, Van CE. [Impact of sleep debt on physiological rhythms]. Rev Neurol
(Paris). 2003;159:6S11–20.
18. Copinschi G, Caufriez A. Sleep and hormonal changes in aging. Endocrinol Metab Clin North
Am. 2013;42:371–89.
19. Nicolaides NC, Charmandari E, Chrousos GP, Kino T. Circadian endocrine rhythms: the hypo-
thalamic-pituitary-adrenal axis and its actions. Ann N Y Acad Sci. 2014;1318:71–80.
20. Tsang AH, Astiz M, Friedrichs M, Oster H. Endocrine regulation of circadian physiology.
J Endocrinol. 2016;230:R1–R11.
21. Cardinali DP. Ma Vie en Noir. Fifty years with melatonin and the stone of madness. Switzerland:
Springer; 2016.
22. Roelfsema F, Pereira AM, Keenan DM, Veldhuis JD, Romijn JA. Thyrotropin secretion by thy-
rotropinomas is characterized by increased pulse frequency, delayed diurnal rhythm, enhanced
basal secretion, spikiness, and disorderliness. J Clin Endocrinol Metab. 2008;93:4052–7.
23. Leproult R, Copinschi G, Buxton O, Van Cauter E. Sleep loss results in an elevation of cortisol
levels the next evening. Sleep. 1997;20:865–70.
24. Mazzoccoli G, Giuliani F, Sothern RB. A method to evaluate dynamics and periodicity of
hormone secretion. J Biol Regul Homeost Agents. 2011;25:231–8.
25. Trivedi H, Szabo A, Zhao S, Cantor T, Raff H. Circadian variation of mineral and bone param-
eters in end-stage renal disease. J Nephrol. 2015;28:351–9.
26. Steiger A, Dresler M, Kluge M, Schussler P. Pathology of sleep, hormones and depression.
Pharmacopsychiatry. 2013;46(Suppl 1):S30–5.
27. Tache Y, Brunnhuber S. From Hans Selye’s discovery of biological stress to the identification
of corticotropin-releasing factor signaling pathways: implication in stress-related functional
bowel diseases. Ann N Y Acad Sci. 2008;1148:29–41.
References 241

28. Rasia-Filho AA. Is there anything “autonomous” in the nervous system? Adv Physiol Educ.
2006;30:9–12.
29. Bartness TJ, Liu Y, Shrestha YB, Ryu V. Neural innervation of white adipose tissue and the
control of lipolysis. Front Neuroendocrinol. 2014;35:473–93.
30. McEwen BS, Gray JD, Nasca C. 60 years of neuroendocrinology: redefining neuroendocrinol-
ogy: stress, sex and cognitive and emotional regulation. J Endocrinol. 2015;226:T67–83.
31. Contoreggi C. Corticotropin releasing hormone and imaging, rethinking the stress axis. Nucl
Med Biol. 2015;42:323–39.
32. Zouikr I, Bartholomeusz MD, Hodgson DM. Early life programming of pain: focus on neuro-
immune to endocrine communication. J Transl Med. 2016;14:123.
33. Schulz A, Vogele C. Interoception and stress. Front Psych. 2015;6:993.
34. Kanbara K, Fukunaga M. Links among emotional awareness, somatic awareness and auto-
nomic homeostatic processing. Biopsychosoc Med. 2016;10:16.
35. Park CL. Meaning making in the context of disasters. J Clin Psychol. 2016;72:1234–46.
36. Kamphuis J, Lancel M, Koolhaas JM, Meerlo P. Deep sleep after social stress: NREM sleep
slow-wave activity is enhanced in both winners and losers of a conflict. Brain Behav Immun.
2015;47:149–54.
37. Yamagata AS, Mansur RB, Rizzo LB, Rosenstock T, McIntyre RS, Brietzke E. Selfish brain
and selfish immune system interplay: a theoretical framework for metabolic comorbidities of
mood disorders. Neurosci Biobehav Rev. 2017;72:43–9.
38. Zorn JV, Schur RR, Boks MP, Kahn RS, Joels M, Vinkers CH. Cortisol stress reactivity across
psychiatric disorders: a systematic review and meta-analysis. Psychoneuroendocrinology.
2016;77:25–36.
39. Herman JP, Tasker JG. Paraventricular hypothalamic mechanisms of chronic stress adaptation.
Front Endocrinol (Lausanne). 2016;7:137.
40. Stern JE. Neuroendocrine-autonomic integration in the paraventricular nucleus: novel roles for
dendritically released neuropeptides. J Neuroendocrinol. 2015;27:487–97.
41. Amihai I, Kozhevnikov M. The influence of Buddhist meditation traditions on the autonomic
system and attention. Biomed Res Int. 2015;2015:731579.
42. Gold PW. The organization of the stress system and its dysregulation in depressive illness. Mol
Psychiatry. 2015;20:32–47.
43. Maniam J, Morris MJ. The link between stress and feeding behaviour. Neuropharmacology.
2012;63:97–110.
44. Kaur J. A comprehensive review on metabolic syndrome. Cardiol Res Pract. 2014;2014:943162.
45. Fitzgerald PJ. Beta blockers, norepinephrine, and cancer: an epidemiological viewpoint. Clin
Epidemiol. 2012;4:151–6.
46. Jansen L, Hoffmeister M, Arndt V, Chang-Claude J, Brenner H. Stage-specific associations
between beta blocker use and prognosis after colorectal cancer. Cancer. 2014;120:1178–86.
47. Eng JW, Kokolus KM, Reed CB, Hylander BL, Ma WW, Repasky EA. A nervous tumor
microenvironment: the impact of adrenergic stress on cancer cells, immunosuppression, and
immunotherapeutic response. Cancer Immunol Immunother. 2014;63:1115–28.
48. Magnon C. Role of the autonomic nervous system in tumorigenesis and metastasis. Mol Cell
Oncol. 2015;2:e975643.
49. Cole SW, Nagaraja AS, Lutgendorf SK, Green PA, Sood AK. Sympathetic nervous system
regulation of the tumour microenvironment. Nat Rev Cancer. 2015;15:563–72.
50. Romeo HE, Colombo LL, Esquifino AI, Rosenstein RE, Chuluyan HE, Cardinali DP. Slower
growth of tumours in sympathetically denervated murine skin. J Auton Nerv Syst.
1991;32:159–64.
51. Kalantaridou SN, Makrigiannakis A, Zoumakis E, Chrousos GP. Stress and the female repro-
ductive system. J Reprod Immunol. 2004;62:61–8.
52. Boeckxstaens G. Chapter 11 – The clinical importance of the anti-inflammatory vagovagal
reflex. In: Buijs RM, Swaab D, editors. Handbook of clinical neurology. Autonomic nervous
system. New York: Elsevier; 2013, p. 119–34.
53. Chrousos GP. Stress and disorders of the stress system. Nat Rev Endocrinol. 2009;5:374–81.
242 5  Third Level: The Hypothalamus

54. Romeo HE, Cardinali DP. The autonomic nervous system of the cervical region as a channel
of neuroendocrine communication. Front Neuroendocrinol 1991;12:278–97
55. Morton GJ, Meek TH, Schwartz MW. Neurobiology of food intake in health and disease. Nat
Rev Neurosci. 2014;15:367–78.
56. Sohn JW, Elmquist JK, Williams KW. Neuronal circuits that regulate feeding behavior and
metabolism. Trends Neurosci. 2013;36:504–12.
57. Munzberg H, Qualls-Creekmore E, Yu S, Morrison CD, Berthoud HR. Hedonics act in unison
with the homeostatic system to unconsciously control body weight. Front Nutr. 2016;3:6.
58. García-García F, Juárez-Aguilar E, Santiago-García J, Cardinali DP. Ghrelin and its interac-
tions with growth hormone, leptin and orexins: implications for the sleep wake cycle and
metabolism. Sleep Med Rev. 2014;18:89–97.
59. Benarroch EE. Neural control of feeding behavior: overview and clinical correlations.

Neurology. 2010;74:1643–50.
60. Mithieux G. Metabolic effects of portal vein sensing. Diabetes Obes Metab. 2014;16(Suppl
1):56–60.
61. Burdakov D, Luckman SM, Verkhratsky A. Glucose-sensing neurons of the hypothalamus.
Philos Trans R Soc Lond B Biol Sci. 2005;360:2227–35.
62. Anderson EJ, Cakir I, Carrington SJ, Cone RD, Ghamari-Langroudi M, Gillyard T, Gimenez
LE, Litt MJ. 60 years of POMC: regulation of feeding and energy homeostasis by alpha-­
MSH. J Mol Endocrinol. 2016;56:T157–74.
63. Joly-Amado A, Cansell C, Denis RG, Delbes AS, Castel J, Martinez S, Luquet S. The hypotha-
lamic arcuate nucleus and the control of peripheral substrates. Best Pract Res Clin Endocrinol
Metab. 2014;28:725–37.
64. Sobrino CC, Perianes CA, Puebla JL, Barrios V, Arilla FE. Peptides and food intake. Front
Endocrinol (Lausanne). 2014;5:58.
65. Biebermann H, Kuhnen P, Kleinau G, Krude H. The neuroendocrine circuitry controlled by
POMC, MSH, and AGRP. Handb Exp Pharmacol. 2012:47–75.
66. Seoane-Collazo P, Ferno J, Gonzalez F, Dieguez C, Leis R, Nogueiras R, Lopez

M. Hypothalamic-autonomic control of energy homeostasis. Endocrine. 2015;50:276–91.
67. Schlogl H, Horstmann A, Villringer A, Stumvoll M. Functional neuroimaging in obesity and the
potential for development of novel treatments. Lancet Diabetes Endocrinol. 2016;4:695–705.
68. Salem V, Dhillo WS. Imaging in endocrinology: the use of functional MRI to study the endo-
crinology of appetite. Eur J Endocrinol. 2015;173:R59–68.
69. Leibowitz SF, Akabayashi A, Wang J, Alexander JT, Dourmashkin JT, Chang GQ. Increased
caloric intake on a fat-rich diet: role of ovarian steroids and galanin in the medial preoptic and
paraventricular nuclei and anterior pituitary of female rats. J Neuroendocrinol. 2007;19:753–66.
70. Fang P, Yu M, Shi M, Zhang Z, Sui Y, Guo L, Bo P. Galanin peptide family as a modulating
target for contribution to metabolic syndrome. Gen Comp Endocrinol. 2012;179:115–20.
71. Cermakian N, Westfall S, Kiessling S. Circadian clocks and inflammation: reciprocal regula-
tion and shared mediators. Arch Immunol Ther Exp (Warsz). 2014;62:303–18.
72. Penicaud L. The neural feedback loop between the brain and adipose tissues. Endocr Dev.
2010;19:84–92.
73. Bailey SM, Udoh US, Young ME. Circadian regulation of metabolism. J Endocrinol.

2014;222:R75–96.
74. Van Cauter E, Blackman JD, Roland D, Spire JP, Refetoff S, Polonsky KS. Modulation of
glucose regulation and insulin secretion by circadian rhythmicity and sleep. J Clin Invest.
1991;88:934–42.
75. Anothaisintawee T, Reutrakul S, Van CE, Thakkinstian A. Sleep disturbances compared to
traditional risk factors for diabetes development: systematic review and meta-analysis. Sleep
Med Rev. 2016;30:11–24.
76. Patton DF, Mistlberger RE. Circadian adaptations to meal timing: neuroendocrine mecha-
nisms. Front Neurosci. 2013;7:185.
77. Cissé YM, Nelson RJ. Consequences of circadian dysregulation on metabolism. ChronoPhysiol
Therapy. 2016;6:55–63.
References 243

78. Karthikeyan R, Marimuthu G, Spence DW, Pandi-Perumal SR, Bahammam AS, Brown GM,
Cardinali DP. Should we listen to our clock to prevent type 2 diabetes mellitus? Diabetes Res
Clin Pract. 2014;106:182–90.
79. Engeland WC, Arnhold MM. Neural circuitry in the regulation of adrenal corticosterone
rhythmicity. Endocrine. 2005;28:325–32.
80. Buijs RM. Chapter 1 – The autonomic nervous system: a balancing act. In: Buijs RM, Swaab
D, editors. Handbook of clinical neurology. Autonomic nervous system. New York: Elsevier;
2013, p. 1–11.
81. Johnson MS, DeMarco VG, Whaley-Connell A, Sowers JR. Chapter 64 – Insulin resistance
and the autonomic nervous system. In: Biaggioni I, Burnstock G, Low PA, Paton JFR, editors.
Primer on the autonomic nervous system. 3rd ed. San Diego: Academic; 2012. p. 307–12.
82. Tseng CJ, Cheng PW, Tung CS. Chapter 29 – Pharmacology of the nucleous tractus solitarii.
In: Biaggioni I, Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous
system. 3rd ed. San Diego: Academic; 2012. p. 141–4.
83. Corbalan-Tutau MD, Gomez-Abellan P, Madrid JA, Canteras M, Ordovas JM, Garaulet
M. Toward a chronobiological characterization of obesity and metabolic syndrome in clinical
practice. Clin Nutr. 2015;34:477–83.
84. Mecawi AS, Ruginsk SG, Elias LL, Varanda WA, Antunes-Rodrigues J. Neuroendocrine regu-
lation of hydromineral homeostasis. Compr Physiol. 2015;5:1465–516.
85. Antunes-Rodrigues J, Ruginsk SG, Mecawi AS, Margatho LO, Cruz JC, Vilhena-Franco T,
Reis WL, Ventura RR, Reis LC, Vivas LM, Elias LL. Mapping and signaling of neural pathways
involved in the regulation of hydromineral homeostasis. Braz J Med Biol Res. 2013;46:327–38.
86. Firsov D, Bonny O. Circadian regulation of renal function. Kidney Int. 2010;78:640–5.
87. Kamperis K, Hagstroem S, Radvanska E, Rittig S, Djurhuus JC. Excess diuresis and natriuresis
during acute sleep deprivation in healthy adults. Am J Physiol Renal Physiol. 2010;299:F404–11.
88. Arai S, Stotts N, Puntillo K. Thirst in critically ill patients: from physiology to sensation. Am
J Crit Care. 2013;22:328–35.
89. Ferguson AV. Circumventricular organs: integrators of circulating signals controlling hydra-
tion, energy balance, and immune function. Boca Raton: CRC Press; 2014.
90. Zimmerman CA, Lin YC, Leib DE, Guo L, Huey EL, Daly GE, Chen Y, Knight ZA. Thirst neu-
rons anticipate the homeostatic consequences of eating and drinking. Nature. 2016;537:680–4.
91. Kopp UC. Role of renal sensory nerves in physiological and pathophysiological conditions.
Am J Physiol Regul Integr Comp Physiol. 2015;308:R79–95.
92. Johns EJ. Chapter 17 – Autonomic regulation of kidney function. In Buijs RM, Swaab D, edi-
tors. Handbook of clinical neurology. Autonomic nervous system. New York: Elsevier; 2013,
p. 203–14.
93. Jackson EK. Chapter 44 – Autonomic control of the kidney. In: Biaggioni I, Burnstock G,
Low PA, Paton JFR, editors. Primer on the autonomic nervous system. 3rd ed. San Diego:
Academic; 2012. p. 215–20.
94. Song W, Wang H, Wu Q. Atrial natriuretic peptide in cardiovascular biology and disease
(NPPA). Gene. 2015;569:1–6.
95. Mahler B, Kamperis K, Schroeder M, Frokiaer J, Djurhuus JC, Rittig S. Sleep deprivation
induces excess diuresis and natriuresis in healthy children. Am J Physiol Renal Physiol.
2012;302:F236–43.
96. Carey RM. Chapter 45 – Dopamine mechanisms in the kidney. In: Biaggioni I, Burnstock G,
Low PA, Paton JFR, editors. Primer on the autonomic nervous system. 3rd ed. San Diego:
Academic; 2012. p. 221–3.
97. Morrison SF. Central control of body temperature. F1000Res. 2016;5
98. Charkoudian N. Chapter 60 – Hypothermia and hyperthermia. In: Biaggioni I, Burnstock G,
Low PA, Paton JFR, editors. Primer on the autonomic nervous system. 3rd ed. San Diego:
Academic; 2012. p. 287–9.
99. Fealey RD. Chapter 7 – Interoception and autonomic nervous system reflexes thermoregula-
tion. In: Buijs RM, Swaab D, editors. Handbook of clinical neurology, autonomic nervous
system. New York: Elsevier; 2013. p. 79–88.
244 5  Third Level: The Hypothalamus

100. Cannon B, Nedergaard J. Brown adipose tissue: function and physiological significance.
Physiol Rev. 2004;84:277–359.
101. Bargut TC, Aguila MB, Mandarim-de-Lacerda CA. Brown adipose tissue: updates in cellular
and molecular biology. Tissue Cell. 2016;48:452–60.
102. Labbe SM, Caron A, Lanfray D, Monge-Rofarello B, Bartness TJ, Richard D. Hypothalamic
control of brown adipose tissue thermogenesis. Front Syst Neurosci. 2015;9:150.
103. Roman S, Agil A, Peran M, Alvaro-Galue E, Ruiz-Ojeda FJ, Fernandez-Vazquez G, Marchal
JA. Brown adipose tissue and novel therapeutic approaches to treat metabolic disorders.
Transl Res. 2015;165:464–79.
104. Peirce V, Carobbio S, Vidal-Puig A. The different shades of fat. Nature. 2014;510:76–83.
105. Blessing W, Mohammed M, Ootsuka Y. Brown adipose tissue thermogenesis, the basic rest-­
activity cycle, meal initiation, and bodily homeostasis in rats. Physiol Behav. 2013;121:61–9.
106. Sessler DI. Perioperative thermoregulation and heat balance. Lancet. 2016;387:2655–64.
107. Hilz MJ. Chapter 48 – Physiology and pathophysiology of female sexual function. In:
Biaggioni I, Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous
system. 3rd ed. San Diego: Academic; 2012. p. 235–8.
108. Rudolph LM, Bentley GE, Calandra RS, Paredes AH, Tesone M, Wu TJ, Micevych
PE. Peripheral and central mechanisms involved in the hormonal control of male and female
reproduction. J Neuroendocrinol. 2016;28.
109. Wahab F, Shahab M, Behr R. The involvement of gonadotropin inhibitory hormone and kis-
speptin in the metabolic regulation of reproduction. J Endocrinol. 2015;225:R49–66.
110. Ullah R, Shen Y, Zhou YD, Huang K, Fu JF, Wahab F, Shahab M. Expression and actions of
GnIH and its orthologs in vertebrates: current status and advanced knowledge. Neuropeptides.
2016;59:9–20.
111. Shaw ND, Butler JP, McKinney SM, Nelson SA, Ellenbogen JM, Hall JE. Insights into
puberty: the relationship between sleep stages and pulsatile LH secretion. J Clin Endocrinol
Metab. 2012;97:E2055–62.
112. Brabant G, Prank K, Schofl C. Pulsatile patterns in hormone secretion. Trends Endocrinol
Metab. 1992;3:183–90.
113. Zeleznik AJ, Plant TM. Chapter 28 – Control of the menstrual cycle. In: Knobil and Neill’s
physiology of reproduction. 4th ed. San Diego: Academic; 2015, p. 1307–61.
114. Veldhuis JD. Neuroendocrine mechanisms mediating awakening of the human gonadotropic
axis in puberty. Pediatr Nephrol. 1996;10:304–17.
115. Gore AC, Hall JE, Hayes FJ. Chapter 37 – Aging and reproduction. In: Knobil and Neill’s
physiology of reproduction. 4th ed. San Diego: Academic; 2015, p. 1661–93.
116. Penev PD. Association between sleep and morning testosterone levels in older men. Sleep.
2007;30:427–32.
117. Bridges RS. Neuroendocrine regulation of maternal behavior. Front Neuroendocrinol.
2015;36:178–96.
118. Miller TV, Caldwell HK. Oxytocin during development: possible organizational effects on
behavior. Front Endocrinol (Lausanne). 2015;6:76.
119. Stolzenberg DS, Champagne FA. Hormonal and non-hormonal bases of maternal behavior:
the role of experience and epigenetic mechanisms. Horm Behav. 2016;77:204–10.
120. Alves E, Fielder A, Ghabriel N, Sawyer M, Buisman-Pijlman FTA. Early social environment
affects the endogenous oxytocin system: a review and future directions. Front Endocrinol
(Lausanne). 2015;6:32.
121. Grattan DR, Le Tissier P. Chapter 12 – Hypothalamic control of prolactin secretion, and the
multiple reproductive functions of prolactin. In: Knobil and Neill’s physiology of reproduc-
tion. 4th ed. San Diego: Academic; 2015, p. 469–526.
Fourth Level: The Limbic System
6

Abstract
In a simplified view, brain function can be considered the product of the neocor-
tex and the limbic system, which complement each other to generate human
behavior with purpose and objective. Phylogenetically, the limbic system com-
prises the oldest parts of the telencephalon and the subcortical structures that
derive from it. The limbic system is essential for emotionality, motivation, learn-
ing and memory. This Chapter analyzes how emotions comprise feelings and
moods, and their expression in somatic and autonomic behaviors. It discusses the
neurobiological mechanisms of memory and how they vary in the three body
configurations (wakefulness, NREM sleep, REM sleep) during a 24-h cycle.

Keywords
Amygdala • Basal ganglia • Chronotypes • Cognitive memory • Emotionality
• Hippocampus • Learning • Limbic system • Memory • Mesolimbic system
• Nucleus accumbens • Papez circuit

Objectives
After studying this chapter, you should be able to:
• Understand how the limbic system is essential for emotionality, motiva-
tion, learning, and memory.
• Describe the structures, connections, and physiological significance of the
limbic components of the basal ganglia.
• Underline how emotions comprise feelings and moods, and their expres-
sion in somatic and autonomic behaviors.
• Enumerate the major findings in functional neuroimaging of the ANS.
• Identify the chronotypes and linked emotional features to the three body
configurations (wakefulness, NREM sleep, REM sleep) during a 24-h cycle.
• Describe the neurobiological mechanisms of memory and how they vary in
the three body configurations (wakefulness, NREM sleep, REM sleep)
during a 24-h cycle.

© Springer International Publishing AG 2018 245


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1_6
246 6  Fourth Level: The Limbic System

 he Limbic System Is Essential in Emotionality, Motivation,


T
Learning, and Memory

Phylogenetically, the limbic system comprises the oldest parts of the telencephalon
and the subcortical structures that derive from it. In a simplified view, brain function
can be considered the product of the neocortex and the limbic system, which com-
plement each other to generate human behavior with purpose and objective [1]. In
this process of complementation, the neocortex mainly regulates precise spatiotem-
poral communication with the environment and executes cognitive and stereognos-
tic functions, producing precise motor outputs (Fig. 6.1).
The limbic system has a primordial link with emotionality and motivation for
action (reinforcement/reward system) and with the process of learning and memory
(involving a high affective content, remembering only what we are interested in
emotionally) [2]. The limbic system gives the information derived from the inner
and outer world its particular emotional meaning [3]. Hence, its role as a last level
in the autonomic motor hierarchy.
Another aspect to consider is the role of the limbic system as a selective inhibitor
of impulses and basic needs, immediately related to survival. The selective inhibi-
tion of certain circuits of a nontopographic character, but relative to memories
loaded with internal meaning can prevent the activation of (too many) lateral ways

Motor
Learning and output
memory Motor circuit

Hippocampus Cerebral
Thalamus
Amygdala cortex
Basal
ganglia

Accumbens Sensory input


Reinforcement
system

VTA

Homeostatic input

Fig. 6.1  The limbic system participates in two of the three basic functional loops of brain function
(learning and memory and the reinforcement system). Modified with permission from Cardinali [1]
The Limbic System Is Essential in Emotionality, Motivation, Learning, and Memory 247

and thus allows the exclusive creation of relevant temporospatial associations (emo-
tional learning). A lateral dispersion in these highly-interconnected circuits would
lead to phenomena of resonance, overabundance, and/or blockade (obsessive ideas,
epileptic seizures, anxiety, etc.). From a physiological perspective, the limbic sys-
tem is able to carry out tasks of this type, as it repeats the basic scheme, present in
many other brain structures, that different sources of information, complementary
and/or opposite, are confronted in the same structure or nodal point, through inter-
mixed circuits [4].
A cortical portion and a subcortical portion are distinguished in the limbic
system:

• The cortical portion consists of the limbic gyrus, part of the ring-shaped cerebral
cortex on the inner side of each hemisphere, separating the neocortex from the
hypothalamus and the brainstem (Figs. 6.2 and 6.3). In 1878, Broca gave the
name “limbic lobe” to this ring of cortical tissue that surrounds the hilum of each
cerebral hemisphere. The limbic gyrus consists of the parahippocampal, cingu-
late, and subcallosal gyrus, and was called “rhinencephalon,” because it was ini-
tially considered to be exclusively associated with the olfactory function. The
orbitofrontal cortex is also included among the cortical areas of the limbic
system.
• The subcortical portion of the limbic system consists of several nuclei (Fig. 6.4).
They are: amygdala, hippocampus, nucleus accumbens, septal nuclei, epithala-
mus (habenula), olfactory bulb, and areas of the anterior thalamus and

Corpus callosum

Cingulate Fornix
cortex

Parahippocampal
Temporal lobe
cortex

Fig. 6.2  Rhinencephalon or limbic gyrus. Modified with permission from Cardinali [1]
248 6  Fourth Level: The Limbic System

Cingulate cortex Fornix


Corpus callosum

Thalamus

Hippocampus
Mammillary bodies

Mammillothalamic tract

Fig. 6.3  Medial view of the structures of the limbic system. Modified with permission from
Cardinali [1]

Cingulate

Thalamus

Frontal lobe

Septum Fornix
Olfactory bulb Stria terminalis
Hippocampus
Medial
forebrain bundle
Parahippocampal
Hypothalamus
girus
Amygdala
Mammillary body

Temporal lobe

Fig. 6.4  Limbic system components. Modified with permission from Cardinali [1]
The Limbic System Is Essential in Emotionality, Motivation, Learning, and Memory 249

hypothalamus (preoptic area, mammillary bodies), and part of the basal ganglia
(ventral pallidal region, innominate substance) [2].

James Papez’s ideas about the limbic system, enunciated in the 1930s, have been
confirmed by recent neuroimaging studies of brain locations [5]. For Papez, the
limbic system is part of the circuit of emotional expression. As it was known that the
hypothalamus was fundamental for the expression of emotional reaction programs,
Papez postulated that the way in which the cerebral cortex modifies, and where
these programs become conscious, is through corticohypothalamic connections via
the cingulate gyrus and the hippocampus [6]. According to Papez’s hypothesis, the
hippocampus processes emotional information and projects to the mammillary bod-
ies through the fornix. The hypothalamus, in turn, provides information to thalamic
nuclei (through the mammillothalamic tract) and from these to the cingulate gyrus
[7]. Subsequently, MacLean extended this scheme to include in the limbic system
hypothalamic areas, the septal area, the nucleus accumbens, neocortical areas (orbi-
tofrontal cortex), and the amygdala. The circuit of Papez, and the most recent modi-
fications to it, are summarized in Fig. 6.5.
The afferent and efferent connections of the limbic system are extremely com-
plex [8]. As we have mentioned, the most outstanding fact is a massive reciprocal
connection with the hypothalamus. The hypothalamus communicates with the hip-
pocampus and the septum through the fornix, with the amygdala through the stria
terminalis and ventral amygdalofugal pathway, and with the portions of the olfac-
tory brain through the central forebrain bundle.
Although there is no complete agreement about the anatomical composition of
the limbic system, it is accepted that a set of structures located in the medial portion
of the telencephalon, highly interconnected with each other, share direct projections
to the hypothalamus, thus regulating the neuroendocrine, autonomic, and behav-
ioral processes associated with this portion of the diencephalon [5].

Prefrontal cortex Association cortex

Cingulate gyrus

Hippocampus
Anterior
thalamic
nuclei Fornix Amygdala

Fig. 6.5  The Papez circuit. Mammillary body


Modified with permission Hypothalamus
from Cardinali [1]
250 6  Fourth Level: The Limbic System

The hippocampus is a portion of the cerebral cortex that forms a kind of horn
along the curvature of the lateral ventricle (Fig. 6.6). It is subdivided into the hip-
pocampus proper, or Ammon’s horn, the dentate gyrus, and the subicular complex.
The connections to hippocampal formation come from the entorhinal cortex, con-
tralateral hippocampus, subcortical structures such as the medial septum, certain
raphe nuclei, and the locus coeruleus (LC) from the brainstem. The hippocampus
projects back to the subicular region and the hippocampus, in turn, extends over
other cortical areas, the anterolateral thalamus, the mammillary bodies, the ventro-
medial and anterior nuclei of the hypothalamus, and the lateral septum. Through the
fornix, the hippocampus projects over the lateral septum (Fig. 6.5) [2, 4].
Thus, the limbic system presents multiple excitation circuits, neuronal substrates
of importance for both emotionality and memory. As we describe later, the fixation
of memory engrams depends on the simultaneous activation of limbic system path-
ways [3].
At the end of the nineteenth century, Jackson proposed that a function is often
represented at various levels of the nervous system. The higher levels, having
appeared later, mediate the function in a more precise form than the lower levels. In
addition, they are more easily excited and usually have an inhibitory action on lower
levels. Therefore, a lesion of the higher levels releases the inferior levels of inhibi-
tion and allows behavioral expression [6].
With this hierarchical notion in mind, MacLean proposed in the 1950s, that
the brain of modern mammals is the sum of three superimposed brains, acquired
during evolution (Fig. 6.7). For MacLean, to the visceral and appetitive of the
primitive reptiles, an emotional brain was added, whose functions would be

Prefrontal cortex

Septum

Hippocampus
(longitudinal view)
Schaffer’s
fiber
CA1
CA3
Perforating
Mossy fiber
fiber Entorhinal Visual cortex
Dentate cortex
girus Hippocampus Hippocampus
(trasversal view) Brain stem
Dentate
gyrus Spinal cord
CA3 cells CA1 cells

Fig. 6.6  The structure of the hippocampus. Modified with permission from Cardinali [1]
The Limbic System Is Essential in Emotionality, Motivation, Learning, and Memory 251

Neomammalian
brain

Paleomammalian
brain

Protomammalian
brain

Fig. 6.7  MacLean’s “trine brain” concept. Modified with permission from Cardinali [1]

assumed, finally, by the limbic system of birds and, mainly, mammals [6]. There
would be:

• A first reptilian or protomammalian, vegetative or instinctive brain, formed by


the upper portions of the spinal cord and part of the brainstem and the basal gan-
glia, with a function in the instinctive survival behaviors (mating, hunting, etc.)
• A second paleomammalian brain, emotional or limbic, and hierarchically supe-
rior to the previous one, the reason it has the possibility to block activation of
primitive drives
• A third neomammalian brain, formed by neocortical structures, capable of analy-
sis stripped of emotional elements.

However, it would be naive to take this functional division as absolute. In fact,


the SNC operates with a unique behavior resulting from the function of the three
hypothetical levels [1].
The limbic system determines the appearance of an inner world, a concept that is
superimposed in part, but not equivalent, to that of internal environment. The inter-
nal world is not based on the presence of interoceptors or the development of
homeostatic mechanisms, but on the development of internal signals of identity. For
example, being able to inhibit certain desires (avoiding a food source in the
252 6  Fourth Level: The Limbic System

presence of a predator) is the behavioral expression of the existence of internal cir-


cuits capable of generating states in which information from extero- and interocep-
tors is subjected to a scrutiny of memories or plans not merely contingent or
immediate. In this sense, the limbic system is a powerful inhibitor of desires and
needs related to the survival of the individual, depending on the conditions of the
internal environment and the external environment (physical and social) [9].

 he Amygdala Is the Main “Motor Nucleus” of the Limbic


T
System

The amygdala plays a major role in the limbic function [2, 10]. It is a subcortical
structure located at the tip of the temporal lobe and continuous with the uncus of the
parahippocampal gyrus (Fig. 6.8). The amygdala is composed of several nuclei,

Amygdala
Hippocampus

Fig. 6.8 Anatomical
localization of the Amygdala
amygdala
The Amygdala Is the Main “Motor Nucleus” of the Limbic System 253

reciprocally connected with the hypothalamus, hippocampus, neocortex, and thala-


mus (Fig. 6.9). Despite the important olfactory input it receives, the amygdala is not
essential for olfactory discrimination [3].
In 1939, Klüver and Bucy described the bilateral lesion of the amygdala nuclei
and part of the anterior pole of the temporal lobe as producing a behavioral syn-
drome in monkeys characterized by: (a) psychic blindness; (b) exaggerated oral
exploratory behavior; (c) temerity or loss of fear, because, for example, they play
with snakes, which they normally fear; (d) excessive indiscriminate eating behavior
(hyperbulimia); (e) increased sexual behavior (self, homo-, and heterosexual); (f)
hypermetamorphosis, or a tendency to react to any visual stimulus. This clinical
picture suggests that, under normal conditions, the amygdala functions as an inhibi-
tory center, preventing reckless or inappropriate behaviors in relation to feeding,
sex, and exploration of the environment [11].
Given the diversity of the afferents received by the amygdala from the limbic
cortex and the rest of the association cortex, in addition to its dense projection on
the hypothalamus nuclei, it is easy to accept the guiding role of the amygdala for the
correct structuring of most available sensory information. Thus, the pioneering
study by Klüver and Bucy showed the involvement of the amygdala in numerous
emotional processes, in which a complex elaboration and association of stimuli
from different sensorial sources are carried out.
The electrical stimulation of the amygdala produces effects on the ANS similar
to those induced by stimulation of the hypothalamus [7]. Primarily, stimulation of
the central amygdala produces changes in BP and heart rate, motility, and gastroin-
testinal secretions, mydriasis, piloerection, etc. Stimulation of the corticomedial
amygdala produces an increase in the secretion of ACTH and gonadotrophins,
whereas the stimulation of the basolateral portion in some cases inhibits it.
Stimulation of the amygdala also induces motor phenomena such as contralateral
head spin, masticatory and swallowing movements, or clonic and rhythmic move-
ments that may become convulsive if the stimulus is prolonged [12].

Amygdala Association cortex


Thalamus
Parieto-
Basolateral temporooccipital
Basal ganglia Frontal
Limbic
Brain stem Central
Raphe nuclei
Locus coeruleus

Hypoph. Hypothal. Corticomedial


Olfactory bulb

Fig. 6.9  Connections the amygdala. Modified with permission from Cardinali [1]
254 6  Fourth Level: The Limbic System

It is characteristic that the effects of stimulation of the amygdala depend on


the functional status of the animal, its environment, and the levels of endo-
crine, metabolic, and autonomic variables. The same stimulus can increase
ACTH levels if they are low, but decrease them if they were previously
increased. This indicates the important role of context evaluation in emotional
response.
The stimulation of the amygdala in humans produces auras with emotional and
polysensorial content (“déjà vu,” hallucinations, etc.). In animals, the selective
lesion of the amygdala decreases the performance in passive avoidance tests, mainly
because of the loss of fear. Animals with a lesioned amygdala show poor affective
behavior, with loss of hierarchical rank [3].
Functionally, three groups of nuclei are distinguished in the amygdala (Fig. 6.9)
[2]:

• Corticomedial, linked to the regulation of the hypothalamus and containing


receptor sites for corticosteroids and gonadal hormones.
• Central, which projects to the brainstem nuclei, such as periaqueductal gray mat-
ter, the NTS and PBN, and the dorsal motor nucleus of the vagus nerve.
• Basolateral, with connections to the association cortexes.

Thus, the corticomedial amygdala participates in endocrine and behavioral func-


tions related to sexual activity, the central amygdala modulates brainstem nuclei
with somatic and autonomic motor responses, and the basolateral amygdala partici-
pates in processes of sensorial and behavioral association. Through the amygdala,
affective behaviors that have proven to be appropriate on previous occasions are
induced.
The central nucleus is closely related anatomically and functionally to the
lateral hypothalamus and to various structures of the brainstem, such as NTS
and PBN, which, in turn, participate in gustatory, cardiorespiratory, and vis-
ceral functions. The corticomedial portion and the periamygdala cortex receive
afferents from the main and accessory olfactory bulb and project to the olfac-
tory cortex. The basolateral portion, which is more phylogenetically modern,
receives afferences from the association cortex, especially of the inferior
(visual) temporal gyrus, upper temporal (acoustic), and lobe of the insula
(somatosensory). It is also closely related to the prefrontal orbitomedial cortex
and to the dorsomedial nucleus of the thalamus. Altogether, the amygdala
nuclei project, through the stria terminalis and the ventral pathway, to various
areas of the hypothalamus, apart from other cortical and subcortical structures
(Fig. 6.10).
The concept of “extended amygdala” defines the mesolimbic and mesocortical
circuits involved in the hedonic response (pleasure) [13]. The neurons of the amyg-
dala respond preferentially to sensory stimuli loaded with an emotional tone, i.e.,
related to situations of reward or punishment.
The Amygdala Is the Main “Motor Nucleus” of the Limbic System 255

Cortical sensorial and associative areas


(cortical activation)

Meynert’s basal Central Basolateral Prefrontal


nucleus cortex

Amygdala
Medial dorsal
Septum Hypothalamus thalamus
(autonomic and N. Accumbens
endocrine control)

Fig. 6.10  Projections of the central and basolateral amygdaloid nuclei

Declarative Emotional
memory memory

Hippocampus Amygdala

Declarative memory Emotional memory

Cognitive expression Emotion

Fig. 6.11  There are cognitive circuits (based on the hippocampus) and emotional circuits (based
on the amygdala) to mediate the two types of memory: declarative and emotional. Declarative
memory implies what is commonly meant by “memory.” The emotional memory involves those
instinctive behaviors, learned or congenital, that protect life. Modified with permission from
Cardinali [1]

The amygdala participates in the learning process, particularly when it comes


to the association of a stimulus with an emotional response. This function is so
important that we recognize today an “emotional memory,” with mechanisms
different from the “cognitive memory” (Figs. 6.11 and 6.12) [3]. The most con-
clusive contemporary evidence for the involvement of the amygdala and other
limbic structures in emotional behavior has been given by PET and fMRI in
individuals with affective diseases and in normal individuals in situations of
­anxiety [14].
256 6  Fourth Level: The Limbic System

Thalamus

Visual Amygdala
cortex

Heart rate

BP Skeletic muscles

Fig. 6.12  Example of an emotional circuit. The snake’s vision triggers a defense reaction before
the cognitive phenomenon of recognition occurs. Modified with permission from Cardinali [1]

The human amygdala mediates interaction between the body and the brain dur-
ing affective processing. The amygdala supports the perception of fear signals and
threat and its activity correlates with the emotional intensity rating of affective pic-
tures including facial expressions. Outputs from the amygdala innervate hypotha-
lamic and brainstem autonomic circuits to trigger autonomic arousal responses to
emotional challenges, particularly threats [7]. Amygdala-induced autonomic
arousal is expressed as increased sympathetic activity and/or decreased heart rate
variability. The amygdala is also sensitive to feedback from the periphery regarding
the state of bodily arousal.
By means of various procedures (electrophysiological, autoradiographic, immu-
nohistochemical, functional neuroimaging), the connections through which the lim-
bic system, via the amygdala, regulate the expression of anxiety, have been
schematized as follows:

• By projections to the lateral hypothalamus, the amygdala produces sympathetic


activation (tachycardia, change in electrodermal response, mydriasis, increased
BP, etc.).
• By projections to the dorsal nucleus of the vagus and ambiguous nucleus, the
amygdala produces parasympathetic activation (gastrointestinal ulcers, urina-
tion, defecation, bradycardia).
Emotions Comprise Feelings and Moods, and Their Expression 257

• From projections to the nucleus parabrachialis, it produces tachypnea.


• By projections to the LC and tegmental areas, the amygdala produces activation
of the noradrenergic, dopaminergic, and cholinergic activity of the reticular for-
mation, with increased alertness.
• By projections to the motor nuclei of the ventromedial descending pathway in
the reticular formation it produces hyperreflexia.
• From projections to the facial and trigeminal motor neurons, it produces changes
in facial expression.
• By projections to the paraventricular nucleus, it induces CRH release, with
stimulation of the adrenal pituitary axis and the descending autonomic
pathways.

Because it controls emotional behavior, the limbic system controls motivation.


Thus, the limbic system determines the appearance of an internal world that inte-
grates the homeostatic functions based on the presence of interoceptors with an
elaboration of internal signals of identity [5].
The avoidance of a food source in the presence of a predator is the behavioral
expression of the existence of internal circuits capable of generating states in which
the information coming from extero- and interoceptors is subjected to verification
of the opportunity to execute it or not. It corresponds, as we discussed in Chap. 5, to
an allostatic response in which physiological systems fluctuate to meet the demands
of external forces. The limbic system is a powerful inhibitor of desires and needs
related to the survival of the individual, depending on the conditions of the internal
environment and the outside world, and therefore the main regulator of allostatic
responses [2].
During emotion regulation, prefrontal control systems modulate emotion genera-
tive systems, such as the amygdala, which is responsible for the detection of affec-
tively arousing stimuli. More specifically, these prefrontal structures include dorsal
regions of the lateral prefrontal cortex that have been implicated in selective atten-
tion and working memory; ventral parts of the prefrontal cortex implicated in
response inhibition; the anterior cingulate cortex, which is involved in monitoring
control processes; and the dorsomedial prefrontal cortex, which is implicated in
monitoring the affective state [15]. A typical pattern detected when individuals
deliberately regulate affective responses (as in mindfulness meditation) is increased
activation within the prefrontal cortex and decreased activation in the amygdala,
suggesting that prefrontal cortex projections to the amygdala exert an inhibitory
top–down influence [16].

 motions Comprise Feelings and Moods, and Their Expression


E
in Somatic and Autonomic Behaviors

Although it seems difficult to reach an agreement to define what is understood by


“emotion,” it is unanimously accepted that in situations that are tense or commit-
ted for the individual (for example, the startle that happens before the sudden
258 6  Fourth Level: The Limbic System

presence of a predator), a nonspecific activation of the vegetative system (tachy-


cardia, cold sweat, etc.) and the skeletal motor system (expression of terror, fight/
flight) occurs, together with a greater or lesser knowledge of the cause of the
shock [17].
Faced with an emotion, we can consider the internal, personal nature, which in
humans also has a cognitive character. We can also consider an external, behavioral
aspect, which serves as a key signal for members of the same species or related spe-
cies [18].
Of course, the external expression of emotions is a consequence of the internal
aspects. The conflict that arises before the unexpected evolution of a situation, such
as the presence of a predator in an unexpected place, is resolved by the activation of
certain autonomic and somatic motor manifestations that imply a reevaluation of
available sensory data. These motor acts, mediated by both the somatic system and
the ANS, are expressive of the state of the inner world [3].
Poor emotional and social adaptation to an environment in constant change char-
acterize limbic system lesions. Without limbic connections and with an intact hypo-
thalamus, cats or monkeys trigger complex behaviors lacking in objective or normal
content, for example, the “false rabies,” hyperphagia, and hypersexuality (Klüver-
Bucy syndrome) mentioned above.
The limbic system acts through the programs contained in the hypothalamus, as
demonstrated by electrophysiological experiments [19]. The electrical stimulation
of the amygdala in the experimental animal triggers effects such as those observed
after hypothalamic stimulation. Such effects include homeostatic responses and
complex autonomic, endocrine, and somatic behaviors.
Bilateral ablation of the amygdala in monkeys eliminates the possibility of social
functioning of the animal [20]. They cannot recognize the social meaning of the
exteroceptive cues that regulate group behavior, and appear anxious and insecure.
This picture is due to the interruption of the flow of information between the pari-
etal–temporal–occipital association cortex and the hypothalamus, which occurs
through the limbic system (in this case, the amygdala). The result of this alteration
is the suppression of a correct evaluation of the sensorial information in the context
of the affective state. Selective lesioning of the amygdaloid nuclei decreases perfor-
mance in passive-type avoidance tests, probably because of the loss of fear. It should
be remembered that in the basolateral amygdala there are numerous receptors for
opiate and GABA, the destruction of which causes a change in the thresholds for
physical pain and affective reactions. In fact, amygdala-lesioned animals present
very poor affective behavior, losing their hierarchical rank in their group, and finally
being rejected by it [20].
The close link among the parietal–temporal–occipital association cortex, the
hypothalamus, and the limbic system is indicated by the following experiments: (a)
amygdala neurons can be activated by stimulation of sensory neocortical areas; (b)
temporal lobe epilepsy in humans is accompanied by various emotional, autonomic,
and sensorimotor signs. Both functional neuroimaging and clinical observations in
humans indicate that the connection: “parietal-temporal–occipital association cor-
tex–amygdala” contains important neuronal substrates of motivated behaviors and
Emotions Comprise Feelings and Moods, and Their Expression 259

emotions. That is, through this system, the sensory information is compared with
the contents of memory and thus becomes significant.
Expression of emotions is primarily based on neurovegetative reactions, which
are, in part, inherited and typical of the species, and partly acquired during early
postnatal age. Innate emotional reactions serve as signals to the congeners and to
members of other species, and are therefore of very important adaptive and evolu-
tionary value (Fig. 6.13).
In parallel with this innate element of emotional behavior, an acquired compo-
nent is identified, resulting from the first stages of contact of the newborn with his
mother and the environment that surrounds him. It is through this process that the
particularization of emotional responses occurs, and therefore, it influences the type
of pathological condition that, if it occurs, is observed in everyone (Fig. 5.14).
The limbic cortex of a newborn child fixes engrams, depending on the type of
emotional stimulation it receives in the early stages of development. Clearly, this is
an active interface between neuroscience and psychology. The production of emo-
tions is associated with the cognitive capacity of the species, and therefore with the
perception and evaluation of sensorial stimuli in relation to the memory of the lived
experience.
The initial works in experimental psychology carried out by Wundt at the end of
nineteenth century, led to the description of the relationship between the intensity of
the sensorial stimulus and the pleasure or not of the perception. Near the threshold,
the stimulus is perceived as neutral, at higher intensities as pleasurable, and at even
greater intensities, as unpleasant. That is, the sensorial and hedonic intensity of a
given stimulus, for example, a certain taste, are not linearly related [1].
To this hedonic theory of emotion, cognitive factors were later added. According
to this interpretation, the intensity of the emotion depends on the level of adaptation
of the subject that perceives and its expectation before the stimulus. A relative dis-
crepancy between these elements generates the opposite effect. Studies based on the

Object
Thalamus Primary sensory
cortex
Learned
Secondary
emotions
sensory cortex

Limbic cortex
Fig. 6.13 Diagram
describing the relationships
Amygdala Hippocampus
between the different
Primary
components involved in the
emotions Hypothalamus Reticular F.
congenital and acquired (innate)
emotional behaviors
A, A', A'' Vegetative
described in the text.
Hypothalamic feedback
Modified with permission
from Cardinali [1] behaviors
260 6  Fourth Level: The Limbic System

Fig. 6.14  Darwin’s drawing to exemplify the instinctive aspects of aggressive interspecies behav-
ior. Modified with permission from Cardinali [1]

analysis of facial expression indicated a three-dimensional aspect for emotion:


pleasant–unpleasant, attention–inattention, and intensity [21].
One of the most persistent influences on the concept of emotion was that of
Charles Darwin, who emphasized its genetic components. Darwin suggested that
emotional expressions are evolutionary remnants of previously adaptive behaviors
that persist, albeit of no use, in a moderate form (e.g., grinding teeth as a sign of
aggression; Fig. 6.14).
William James was the first to propose that emotion consists of bodily changes
originating in the perception of the stimulus (Fig. 6.15) [22]. This theory was called
James–Lange, because of the contribution made independently by a Danish physi-
cian, Carl Lange, to its formulation. According to the James–Lange theory, emo-
tional quality is the result of perceived changes in bodily activity triggered because
of sensory perception.
The elucidation of the structure and function of the ANS marked a fundamental
milestone in the study of the visceral correlates of emotion. It was Walter Cannon
who first established the direct link between emotional activity and sympathetic
function. In what Cannon called an “emerging theory of emotion,” he described the
sympathetic division of the ANS as the mediator of the reaction to stress. For
Cannon, it is the CNS that triggers the emotions and not the bodily changes
(Fig. 6.15) [23]. More recently, a cognitive view developed that combines both posi-
tions: there is a central effect of production of the emotional by the limbic system,
which is fed by body correlates of the emotions (Fig. 6.15) [17]. Improved anatomi-
cal and functional description of bidirectional interactions between body and brain
has advanced our understanding of emotional and, for some emotions, there is good
evidence for specific coupling with autonomically mediated changes in peripheral
physiology [24].
The mind and body are intrinsically and dynamically coupled. Perceptions,
thoughts, and feelings change, and respond to, the state of the body [10].
Neuroimaging techniques are beginning to detail the neuronal substrates mediating
these interactions between mental and physiological states, implicating cortical
regions (specifically insular and cingulate cortices) alongside subcortical (amyg-
dala) and brainstem (notably dorsal pons) in these mechanisms [10, 21]. For exam-
ple, by combining fMRI with carotid stimulation in healthy participants, it was
shown that manipulating afferent cardiovascular signals alters the central
Emotions Comprise Feelings and Moods, and Their Expression 261

Cortex Cortex

Thalamus Thalamus

Limbic syst. Stimulus Limbic syst. Stimulus

Body changes Body changes

James-Lange’s theory Cannon-Bard’s theory


Cortex

Thalamus

Limbic syst. Stimulus

Body changes
Cognitive theory

Fig. 6.15  The different theories of the production of emotions. The current (cognitive) view is
eclectic between two opposing positions. Modified with permission from Cardinali [1]

processing of emotional information (fearful and neutral facial expressions) [25].


Carotid stimulation attenuated activity across cortical and brainstem regions.
Modulation of emotional processing was apparent as a significant expression-by-
stimulation interaction within the left amygdala, where responses during appraisal
of fearful faces were selectively reduced by carotid stimulation. Moreover, activity
reductions within insula, amygdala, and hippocampus correlated with the degree of
stimulation-­evoked change in the explicit emotional ratings of fearful faces. Across
participants, individual differences in autonomic state, as assessed by heart rate
variability, predicted the extent to which carotid stimulation influenced neural
(amygdala) responses during appraisal and subjective rating of fearful faces [25].
Thus, cortical locations of emotional function in man have begun to be defined
using PET and fMRI [26–29]. This methodology is also employed to determine the
sites and effects of pharmacological treatments [30].
For the physiological detection of emotion, a sensitive, non-invasive indicator is
the electrical response of the skin, also known as the psychogalvanic reflex, or elec-
trodermal reflex. The potential or electrical resistance (or conductance) of any part
262 6  Fourth Level: The Limbic System

of the body can be quantified through electrodes on the skin. The skin conductance
response is a remarkably powerful and informative psychophysiological index [31].
Because it is relatively easy to measure, and provides reliable indices of a wide
variety of psychological states and processes, skin conductance response has been
one of the most popular aspects of ANS activity used to study human cognition and
emotion [31, 32]. The analysis of the variability of the heart rate also allows the
evaluation of the sympathetic and parasympathetic response at the thoracic level
before different emotional situations (Fig. 4.13) [33].
Muscle tone is another general peripheral indicator for emotions such as anxiety or
fear, particularly at the level of the face and neck muscles. The startle reflex is considered
a phenomenon that is closely related to the emotional state of the individual. This reflex
consists of an initial blink, with a latency close to 0.04 s. Then, contraction of the skeletal
muscles ensues, with a latency of 0.1 s. Finally, after 1 s, more complex signs appear
(changes in skin potential, increased BP and heart rate). This sequential motor program
is an example of stereotyped reactivity of ANS and the somatic motor system [34].
It must be noted that the different physical and chemical indicators of emotional-
ity used so far simply reflect an overall level of emotional tension and do not dis-
criminate between types of emotions.

Limbic Components of the Basal Ganglia

The main function of the basal ganglia is to select a particular movement or sequence
of thoughts or an autonomic response that is most appropriate for the situation, sup-
pressing any possible other ones [1]. Thus, the basal ganglia play an important role
in limbic function.
There are five main components of the basal ganglia (Figs. 6.16 and 6.17): (a)
three subcortical nuclei: caudate, putamen, and globus pallidus; (b) a diencephalic
component: the subthalamic nucleus of Luys; (c) a mesencephalic component: the
substantia nigra and the ventral tegmental area [35].
The caudate and putamen have the same embryological origin, identical cellular
types and are fused by their anterior part (to form the striatum). The ventral part of
the striatum (ventral striatum or nucleus accumbens) has a functional identity
because of its connection with the limbic system. The striatum comprises the entry
nuclei to the circuit of the basal ganglia.
The globus pallidus is a diencephalic structure divided into two segments, inter-
nal and external (or medial and lateral). The substantia nigra, which is the largest
nucleus of the midbrain, comprises a compact dorsal portion (“pars compacta”) of
pigmented dopaminergic cells, and a ventral, reticular portion (“pars reticulata”) of
nonpigmented GABAergic neurons.
The substantia nigra pars reticulata and the medial (or internal) globus pallidus
form a functional unit as the exit sector of the basal ganglia. In these connections,
and in the functional relation with the limbic system (ventral striatum), another set
of dopaminergic neurons (ventral tegmental area, VTA, or A10), adjacent to the
substantia nigra, participates [35].
Limbic Components of the Basal Ganglia 263

Thalamus
Internal
capsule
Lateral ventricle

Caudate

Putamen

External pallidum
Internal pallidum

Subthalamic
Substantia
nucleus
nigra, VTA
Amygdala

Fig. 6.16  Components of the basal ganglia. Modified with permission from Cardinali [1]

Cortex
+

Thalamo- + +
cortical path
Cortico-striatal
projections

Thalamo-
striatal path Striatum
+

IL Indirect
Direct pathway
pathway DA (D2-R)
(D1-R) –

– External
Thalamus
– pallidum
+
Basal
+ –
ganglia
+
output Subthalamic
SNc – nucleus
+
VTA SNr
Internal pallidum

Fig. 6.17  Connections (direct and indirect) of the basal ganglia. SNc substantia nigra pars com-
pacta, SNr substantia nigra pars reticulata, VTA ventral tegmental area. Modified with permission
from Cardinali [1]
264 6  Fourth Level: The Limbic System

Cerebral Cortex

Motor and
associative Limbic

Glutamate Glutamate

GPe
D2
GABA
D1 enkephalin

Dopamine Striatum
GABA
GABA
substance P
dynorphin
Glutamate NTS
SNpc
VTA

SNpr
GPi Thalamus
GABA

Fig. 6.18  Neurotransmitters in the direct and indirect pathways of the basal ganglia. Modified
with permission from Cardinali [1]

The main entry to the basal ganglia is at the level of the striated body (caudate–
putamen) and the ventral striatum (accumbens; Figs. 6.17 and 6.18). Both the cau-
date and the putamen receive an important dopaminergic projection of the substantia
nigra pars compacta (nigrostriatal pathway). The ventral striatum receives projec-
tions from the dopaminergic neurons of the VTA.
The efferent pathway of basal ganglia has two main origins: (a) the globus palli-
dus (medial portion); (b) the substantia nigra pars reticulata. Both sets of
GABAergic neurons project to the thalamus (on specific ventral lateral and ventral
anterior nuclei and on the association nuclei), from where projections arise to the
cerebral cortex. The globus pallidus also projects to thalamic intralaminar nuclei
(which in turn project to the striatum). The thalamic intralaminar nuclei send and
receive glutamatergic projections from various areas of the cerebral cortex and
from subcortical areas. As discussed in Chap. 2, the state of this circuit determines
the three body configurations: wakefulness, slow-wave sleep, and REM sleep
found in a 24-h cycle.
The connections of the basal ganglia are organized in two main ways (Figs. 6.17,
6.18, and 6.19): (a) a direct pathway, involving the projection of the striatum to the
globus pallidus (medial portion)/substantia nigra pars reticulata and from there to
the thalamus; (b) an indirect pathway, comprising the striatum, globus pallidus (lat-
eral portion), and the subthalamic nucleus, and from there, to the substantia nigra
pars reticulata/globus pallidus (medial portion).
Limbic Components of the Basal Ganglia 265

Motor Cognitive Limbic


circuit circuit circuit
D4 D4
Motor D2 Association Limbic D2
D1 D1
cortex cortexes cortex

(Glu) (Glu) (Glu) (DA)


D1 D1 D3 D2 D1
D2
(DA) D2
Putamen Caudate Accumbens

(GABA) (GABA) (GABA)


SNpc VTA
Internal globus D2 Substantia Ventral globus D2
pallidus VTA nigra pars pallidus
D2 reticulata
(GABA) (GABA) (GABA)

Ventrolateral Anteroventral Dorsomedial


thalamus thalamus thalamus
(Glu) (Glu) (Glu)

Fig. 6.19  Homologies among the motor, cognitive, and limbic circuits of the basal ganglia.
Modified with permission from Cardinali [1]

In both pathways, the basal ganglia and thalamus form closed circuits that receive
information from large portions of the cortical territory, and project to cortical areas
of motor planning (mainly, the supplementary motor area), cognitive cortex (frontal
association cortex), and limbic association cortex.
After stimulation of the cerebral cortex, the inhibitory projections of the striatum are
stimulated with two consequences: (a) the direct pathway reduces the inhibitory activ-
ity of the basal ganglia nuclei (medial globus pallidus/substantia nigra pars reticulata)
on the thalamus; (b) the indirect pathway, via reduction of the inhibition exerted by the
lateral globus pallidus on the subthalamic nucleus, increases the excitation given by the
subthalamic glutamatergic projections on the medial globus pallidus/substantia nigra
pars reticulata. In this case, the inhibitory activity on the thalamus increases [36].
From the functional point of view, there is a definite anatomical segregation in
the striatum: the putamen is linked to the motor functions, the caudate is primarily
linked to cognitive functions (receiving the thalamic-striatal projection from the
intralaminar nuclei as input) and the ventral striatum is associated with the limbic
system (Fig. 6.19).
The function of the striatal–pallidal and striatal–nigral connections is to trans-
form the excitatory input of the cortex into a balanced antagonism on the major exit
neurons of the basal ganglia, i.e., the medial globus pallidus/substantia nigra pars
reticulata GABAergic neurons. The dopaminergic input modulates this balance. DA
acts on excitatory D1-type receptors in striatal neurons of the direct pathway, and on
inhibitory D2-type receptors in striatal neurons of the indirect pathway.
266 6  Fourth Level: The Limbic System

As stated, the main function of the basal ganglia circuit is to select a movement
or sequence of thoughts or an autonomic response that is most appropriate for the
situation, suppressing any others. To achieve this, three circuits are built with func-
tional common features and sequelae in the alterations that compromise them
(Fig. 6.19) [35]. These three circuits are:

• The motor circuit originates in the regions of the motor and premotor cortex, and
in the somatosensory cortex. It passes through the putamen, dorsolateral globus
pallidus, and the ventrolateral nucleus of the thalamus to project back to the
supplementary motor cortex. The alteration of this circuit produces hypokinetic
sequelae (such as bradykinesia of Parkinson’s disease) or hyperkinetic sequelae
(such as Huntington’s chorea).
• The cognitive circuit originates in the dorsolateral prefrontal cortex, projects to
the dorsolateral portion of the caudate nucleus and from here to the dorsolateral
globus pallidus and the ventral anterior and dorsomedial thalamic nuclei, to close
the circuit in the dorsolateral prefrontal cortex. Lesions of this circuit (equivalent
to bradykinesia in the motor circuit) produce executive deficit (apathy), with dif-
ficulty in working memory and action. The obsessive–compulsive disorder is the
hyperkinetic equivalent in cognitive circuit lesions, in which stereotyped behav-
iors are repetitively performed (e.g., washing hands dozens of times) similar to
motor tics. In Tourette’s syndrome, motor tics coincide with compulsions and
obsessions.
• The limbic circuit originates in the inferior and lateral portion of the frontal
cortex (orbitofrontal) and projects to the ventromedial region of the caudate
nucleus, the nucleus accumbens, and the dorsomedial region of the globus
pallidus, to return to the cortex via the anterior and dorsomedial ventral thala-
mus. This circuit is especially relevant for functions of personality, socializa-
tion, restriction of impulses, empathy, etc. Abnormal hyperactivity in this
circuit results in addictive behavior, irritability, impulsivity, and disinhibition.
Abnormal hypoactivity is manifested by anhedonia, that is, the inability to
experience pleasure.

The mesolimbic system consists of dopaminergic projections from the mid-


brain to the cortex (mainly the prefrontal cortex, mesocortical portion) and the
ventral striatum (mesencephalic portion). The dopaminergic neurons are in the
VTA (A10), adjacent to the substantia nigra, which send their axons to the cortex
and the ventral striatum (nucleus accumbens, the ventral part of the caudate, and
the putamen). Dopaminergic neurons of the VTA discharge in the presence of
reward or of stimuli that predict reward. These projections are targets of drugs
capable of generating addiction and their circuits are altered in diseases such as
schizophrenia [13].
The ventral striatum receives most of its excitatory afferents from the hippocam-
pus, basolateral amygdala, and prefrontal cortex. Another part includes the cingu-
late circuit that originates in the region of the anterior cingulate, and projects to the
Limbic Components of the Basal Ganglia 267

nucleus accumbens, the olfactory tuber, regions of the ventromedial caudate, and
the putamen. The circuit returns to the cortex through the lateral globus pallidus and
to the anterior cingulate via the dorsomedial thalamus.
Given a pattern of cortical information input, the striatum selects the most appro-
priate behavioral action repertoire that is triggered by the activation of the direct
path. Simultaneously, and through the activation of the indirect pathway, the stria-
tum suppresses the execution of inappropriate behavioral actions. The activity of
dopaminergic neuron signals predicts environmental events of potential importance
for the individual. The dopaminergic neurons of the substantia nigra or the VTA are
activated only by sensorial stimuli that have a motivational meaning. By these
mechanisms, the basal ganglia participate in the learning and selection of the most
appropriate behavioral patterns for a determined environmental and motivational
context [35].
It must be noted that routine automation reduces the computational burden of the
cerebral cortex by enabling it to process other types of information “in parallel.”
Thus, the basal ganglia deal with the automatic (implicit) processing of information
whereas the cerebral cortex deals with complex tasks of consciousness (explicit
processing of information). The cortico-striatal circuits are the basis for the trans-
formations that convert a cognitive frame of reference into an appropriate sequence
of actions.
In 1954, the areas of reward and punishment were identified, mostly located in
the limbic system. This was done by implanting stimulation electrodes in the central
forebrain bundle and found reinforcement of induced behavior, i.e., the animal con-
centrated all its effort in self-stimulating without paying attention to other meaning-
ful stimuli such as food. Other positive reinforcement points identified were the
VTA, nucleus accumbens, prefrontal cortex, and lateral hypothalamus. In all these
cases, the animals self-stimulated until emaciation, ceasing the effect if the dopami-
nergic neurons of the VTA were destroyed. The idea that the mesolimbic dopami-
nergic system was substantial in defining the hedonic characteristics of a stimulus
was thus consolidated (Fig. 6.20).
Two circuits are distinguished in this reward system [13]: (a) a mesolimbic cir-
cuit, composed of projections of the cell bodies of the ventral tegmental area to the
nucleus accumbens, amygdala, and hippocampus, which is involved in acute rein-
forcing effects, memory, conditioned responses, and emotional changes of the with-
drawal syndrome; (b) a mesocortical circuit, which includes projections of the
ventral tegmental area to the prefrontal cortex, orbitofrontal cortex, and cingulate
cortex. This circuit is involved in the conscious experience of drug effects, “crav-
ing,” and the compulsion to use drugs [13].
The mesolimbic and mesocortical circuits operate in parallel and are affected
reciprocally and with other areas, forming what has been called the “extended
amygdala.” In the extended amygdala, there is interaction of the mesolimbic and
mesocortical circuits by means of projections of the GABA neurons of the nucleus
accumbens to the VTA and to the prefrontal cortex and glutamatergic projections
from the prefrontal cortex to the nucleus accumbens and to the VTA.
268 6  Fourth Level: The Limbic System

Amphetamines, cocaine, opiods,


cannabis, phencyclidine

5-HT Glu
Amygdala Glu
Opioid DA DA
GABA
GABA Prefrontal
Opioid cortex
GABA
DA
GABA Accumbens
NE DA
Locus
5-HT coeruleus
Opioid
Raphe nuclei
Opiods, ethanol, barbiturates,
Ventral tegmental area benzodiazepines

Fig. 6.20  The mesolimbic and mesocortical circuits that make up the “extended amygdala,”, their
neurotransmitters, and the drugs of abuse that act upon them. Modified with permission from
Cardinali [1]

The dopaminergic projection modulates the flow of information in the “extended


amygdala”, providing a signal indicative of the presence of an important event that
requires attention. This mesolimbic/mesocortical system participates not only in
mechanisms of reward, but also in aversive responses, including stress.
Addictive substances induce pleasant states (euphoria in the initiation phase) and
a decrease in pain. Its continued use produces adaptive changes in the CNS that lead
to tolerance, physical dependence, sensitization, craving, and relapse in use. Both
natural pleasures (food, drink, sex) and addictive drugs stimulate the release of DA
from neurons in the VTA that project to the nucleus accumbens. This is translated
into euphoria and reinforcement of behavior. In the case of natural pleasurable stim-
uli, there is a very rapid adaptive change (habituation). In the case of the response
to addictive drugs there is no habituation (each dose of the drug stimulates the dis-
charge of DA). During the abstinence syndrome associated with the drug abuse
listed in Fig. 6.20, there is a significant decrease in DA release in the nucleus accum-
bens. In individuals with high limbic DA levels, there is a poor initial experience
with less possibility of continued use of abuse drugs. In contrast, individuals with
low levels of limbic DA, there is an increased risk of drug use because of the intense
initial response [13].

Functional Neuroimaging of ANS

As discussed in previous chapters, a basic concept is that the function of the ANS is
given by a highly interconnected hierarchy of neuronal structures involving brain
areas from the neocortex to the brainstem, in addition to the cerebellum and the
Functional Neuroimaging of ANS 269

basal ganglia. The way in which those structures operate in a network was first
demonstrated by neuroanatomical and neurophysiological techniques [37, 38], and
more recently by the application of neuroimaging techniques, mainly fMRI [39].
In Chap. 7, we analyze several autonomic tests available to clinically assess ANS
function (Table 7.2). Some can be adapted to the MRI environment, whereas others
are impractical inside an MRI environment (such as orthostatic changes elicited
either by a tilt table or by changing posture from sitting to standing). Procedures
involving a static body position with minimal electrical equipment are most com-
monly used. Valsalva maneuver has been the focus of attention as a simple chal-
lenge that elicits a strong autonomic reaction (Chap. 7). The Valsalva maneuver can
be performed in the supine position and can be repeated multiple times within a
typical fMRI protocol [40]. A hand grip, which also involves sympathetic activity
increases, is another autonomic test feasible to be used in fMRI.
Functional MRI is less suitable for identifying changes in state that last several
minutes, such as the quantitative sudomotor axon reflex test. However, electrical
stimulation of muscle or nerves has been performed.
By using fMRI, the pathways for sympathetic outflow in the ventral medulla were
described, and the temporal patterns for such medullary activation on fMRI to foot cold
pressor and Valsalva maneuver were readily apparent in healthy adolescents and adults.
Concerning the central autonomic network and limbic regions, neuroimaging has
confirmed the original findings derived from recording, lesion, stroke, and physio-
logical studies, demonstrating that cortical brain regions and other rostral brain
areas participate in autonomic regulation [41], and have extended the regions we
now know to be involved in autonomic regulation.
For example, the insula participates in BP challenges in a significant fashion.
Forehead cold pressor, lower body negative pressure, the Valsalva, and the related
forced expiratory loading all lead to insular activation. Hand grip and maximal
inspiratory loading similarly recruit the anterior and posterior insula. The insula has
inhibitory projections to the hypothalamus and its functional organization of the
insula is asymmetrical, with the right side being preferentially active during sympa-
thetic increases and the left side during parasympathetic action [39].
Other regions involved in autonomic regulation are the cingulate, the ventrome-
dial prefrontal cortex, basal ganglia, and hypothalamus, along with the amygdala
and hippocampus [42]. As described in Chap. 5, the hypothalamus plays a major
role in regulating autonomic outflow, with substantial projections from other limbic
structures and efferent projections to the brainstem. The hypothalamus shows fMRI
signal responses under some conditions, but as the structure is small, differentiating
local responses of the multiple subnuclei of the hypothalamus by fMRI is difficult.
The ventromedial prefrontal cortex, amygdala, and hippocampus play significant
roles in the sequencing of responses to BP and other ANS challenges.
By recording muscle sympathetic nerve activity at the same time as performing
fMRI of the brain, the cortical structures involved in central cardiovascular control
in awake human subjects can be best identified. Signal intensity and muscle sympa-
thetic nerve activity correlated positively in the left mid-insula, bilateral dorsolateral
prefrontal cortex, bilateral posterior cingulate cortex, and bilateral precuneus.
270 6  Fourth Level: The Limbic System

In addition, muscle sympathetic nerve activity covaried with signal intensity in the
left dorsomedial hypothalamus and bilateral ventromedial hypothalamus (VMH).
Construction of a functional connectivity map revealed coupling between activity in
the VMH and the insula, the dorsolateral prefrontal cortex, the precuneus, and in the
region of the left and right rostroventrolateral medulla [43].
In thermoneutral conditions, resting skin muscle sympathetic nerve activity is
related to the level of arousal and emotional state. The identified brain regions
responsible for the generation of spontaneous muscle sympathetic nerve activity
include the left thalamus in the region of the ventromedial nucleus, the left posterior
and right anterior insula, the right orbitofrontal cortex, the right frontal cortex, and
bilaterally in the mid-cingulate cortex and precuneus [44]. Functional connectivity
analysis revealed a strong positive coupling between the right orbitofrontal cortex
and the right anterior insula. Signal intensity changes within the precuneus were
temporally coupled with the left anterior and posterior insula, cerebellum, cingulate
cortex, and thalamus. Presumably, these brain regions monitor the internal state of
the body and may regulate emotional state changes [3].
One important finding in fMRI studies has been to discover the cerebellar contri-
butions to ANS regulation (Chap. 4). The cerebellar cortex responds regionally to
BP changes, including respiratory loading, lower body negative pressure, Valsalva
and Mueller maneuver, cold pressor, end-expiratory breath hold, and static hand
grip [39]. The data are consistent with a dampening or coordinating role for the
cerebellum in the presence of significant changes in BP, which could be similar to
the motor coordination role traditionally associated with the structure [45].
Autonomic functions in the brain are lateralized [42], in a manner reminiscent of
other functions, including motor, sensory, and language systems. The cold pressor
and hand grip challenges show multiple structures with lateralized responses to the
challenges, notably in the mid and posterior insula. The amygdala, hippocampus,
and ventral cerebellum show opposite responses to a Valsalva maneuver on the left
and right sides. The insular cortex is of interest with respect to lateralized autonomic
function, as the left-side function appears to be preferentially parasympathetic and
the right-side preferentially sympathetic. Hence, resection of the left insula led to
minimal autonomic changes, but resection of the right led to less sympathetic and
more parasympathetic activity [46]. The lateralization of function has obvious
implications for stroke or other injury, as unilateral damage would have an impact
on the extent and timing of BP regulation.
The vermis participates in fear learning and memory mechanisms related to the
expression of autonomic and motor responses of emotions. In humans, the cerebel-
lar hemispheres are also involved at a higher emotional level [47].
In rodents, the reversible inactivation of the vermis during the consolidation or
the reconsolidation period hampers the retention of the fear memory trace. In this
region, there is a long-term potentiation of both the excitatory synapses between the
parallel fibers and the Purkinje cells and of the feed-forward inhibition mediated by
molecular layer interneurons (Fig. 4.19). This concomitant potentiation ensures the
temporal fidelity of the system. Additional contacts between mossy fiber terminals
and Golgi cells provide morphological evidence for the potentiation of another
Chronotypes, 24-h Rhythms and Emotion 271

feed-forward inhibition in the granular layer. Imaging experiments show that in


humans, the cerebellum is also activated during mental recall of emotional personal
episodes and during learning of a conditioned or unconditioned association involv-
ing emotions [45].

Chronotypes, 24-h Rhythms and Emotion

A Gaussian distribution of the acrophase in the body temperature 24-h rhythm


occurs, with a mean at 18:00 h and with two-thirds of the population within a ±1 h
range. Five % of the population are out of phase 2 h before or 2 h after the mean,
and these subgroups are called “larks” (morning type) and “owls” (evening type)
respectively. Thus, the people with different chronotypes, the morning or larks and
the evening or “owls,” form the limits of the normal distribution in the human popu-
lation [48].
The acrophases of the circadian rhythms of body temperature, mental perfor-
mance, and sleep–wake cycles in the morning and evening groups occurs consider-
ably before or after respectively what is considered the norm. These differences can
be due to differences in sensitivity of the phase adjustment process of the circadian
clock and/or to the lifestyle. Differences in lifestyle influence exposure to zeitgeber,
as those individuals who go to bed and get up early are exposed much earlier to the
environmental zeitgebers. However, the phase differences persist, even though the
individuals studied are maintained in the same sleep–wake cycle, or in a constant
routine, which would rather point to an endogenous origin, related to the synchroni-
zation mechanisms of the circadian pacemaker. The chronotypes are associated
with genetic variations, lifestyle differences, mood states, cognitive function, and
risks of health problems (sleep disorders, depression) [48].
The “larks” are active in the morning, reach their maximum performance during
the noon hours, and enjoy little any nocturnal obligations, at which time they show
tiredness and a predisposition to sleep. The “owls” rise late, are gaining energy dur-
ing the day, and reach their maximal performance toward the night-time; they pre-
fer, therefore, to prolong the vigil period.
The inclination of humans to sleep at night is probably linked to the dependence
of the primate on vision, not smell, as the dominant sense. An in-day wakefulness
and overnight sleep program must have been advantageous in a primitive world
plagued by dangerous nocturnal predators. As a corollary, sleep must have been a
positive influence on natural selection for primates, being beneficial during the
hours of darkness. From this evolutionary point of view, there is a sense of popula-
tion distribution in “larks” and “owls,” as the existence of individuals with different
sleep–wake rhythms should have allowed greater efficacy in surveillance before the
possible nocturnal predators of the hominids.
Several studies in twins have shown genetic links with aspects such as circadian
time and sleep/wake preferences. The “owls” suffer from a conflict between internal
and external time (“social jet lag” [49]) as they have more difficulty adapting to the
demand for morning activity than the “larks” or intermediate chronotypes.
272 6  Fourth Level: The Limbic System

Studies on fluorine-19 nuclear magnetic resonance imaging have revealed differ-


ences in the metabolic function of the brain among “owls” compared with “larks”
[50]. These metabolic differences were discovered in limbic regions and may be a
reason why the nocturnal chronotype has a higher risk of depression related to
insomnia. The diminished whiteness of the CNS white matter can be the result of
chronic social jet lag because of the stress it entails: people who are willing to stay
late and sleep late are often in constant conflict with the schedule that surrounds
them.
A common genetic variant has been identified that affects almost the entire
human population, and is responsible for the difference of more than 1 h a day in the
tendency to be an early riser or a night owl [51]. A single nucleotide in Per1 varied
between two groups that differed in their sleep–wake behavior. At this site of the
genome, 60% of individuals have the nucleotide base adenine (A) and 40% have the
nucleotide base guanine (G). Because one has two sets of chromosomes, in any
given individual, there is a 36% chance of having two A’s, a 16% chance of having
two G’s, and a 48% chance of having a mixture of A and G on this site. People who
have the AA genotype wake up about 1 h earlier than people who have the GG geno-
type and the AG wake up almost exactly in the middle.
There is evidence for the relations between chronotype and cognitive ability.
Nocturnity was positively related to cognitive ability, but negatively correlated with
academic performance indicators. On the contrary, morningness had a negative rela-
tionship with cognitive ability and a positive relationship with academic indicators.
When school performance and inductive intelligence were measured, evening sub-
jects scored higher than morning subjects in inductive reasoning (a good estimate of
general intelligence and one of the strongest predictors of academic achievement).
In general, the nocturnal chronotype is more common in creative and extroverted
individuals (poets, artists, inventors), whereas the morning chronotype is in indi-
viduals with logical, systematic, and deductive thinking (accountants). It is interest-
ing that inductive reasoning is linked to innovative thinking, higher-profile
occupations, and higher incomes. In general, studies have revealed that school per-
formance of “owls” is usually lower than that of “larks” because the late chronotype
is adversely affected by school schedules.
The nocturnal chronotype is also more prone to the indiscriminate use of tobacco
and alcohol and to a greater caloric intake. Several studies have documented how
the timing of sleeping and eating can affect body weight (Chap. 5) [52].
If our preferences for sleep and wakefulness are strongly influenced by genetics
and biology, what can we do when these inclinations do not match the demands and
responsibilities of our lives? Indeed, limiting exposure to artificial light at night and
increasing exposure to sunlight during the day can change sleep–wake cycles, even
in owls. Adequate sleep habits, not consuming alcohol before bedtime and main-
taining a dark bedroom and no interfering technology can help to reinforce the sleep
schedule, even if it does not perfectly align with the natural tendencies of the subject
in question.
There are better times of day for the execution of tasks depending on the cogni-
tive or physical abilities needed [53]. To know them and to try as far as possible to
Chronotypes, 24-h Rhythms and Emotion 273

select the optimal schedules to develop the activities facilitate and optimize the
performance. At the end of the day, the ability to concentrate and to make a material
to be stored in the long-term memory are superior, whereas the ideal time to face a
test occurs at noon, when the working memory and the feeling of activation and
well-being are better. These daily fluctuations of behavioral parameters cannot be
considered trivial, as the total change detected is about 10% and the variation in the
efficiency of execution is equivalent in magnitude to the effect of sleeping only 3 h
or of ingesting the legal limit of alcohol [53].
As for the measurement of biological parameters, there are numerous variables
that influence them, minimizing or amplifying the presence of rhythmicity in the
subjects’ performance skills. These include the registration status (habitual activity
or controlled in the laboratory), food intake, consumption of psychoactive sub-
stances and environmental conditions (temperature, light, noise, etc.), or social
interaction. Thus, the study of circadian variations in cognitive and physical perfor-
mance presents methodological difficulties added to those of the specific tasks per-
formed [53]. If we value the response of individuals to a task, we must consider that
their performance may be more influenced by the learning process and/or by their
own fatigue or boredom that involves doing it repeatedly than by the very effect of
the time of recording. It is impossible to carry out continuous measurements of
cognitive and physical performance, selecting an interval between records that
allows the recovery of the individual as per the demand, type, and duration of the
tests used. A sampling interval every 2–3 h is usually adequate and, wherever pos-
sible, reliability increases by logging more than 1 day. In addition, the levels of
motivation and stress of individuals, not always easy to evaluate and therefore to
control, are factors that have a decisive influence in the evaluations of execution. For
example, highly motivated subjects perform better and show minimal diurnal
variability.
There are also considerable variations in the phase and amplitude of circadian
rhythms with age, especially in childhood and in senescence. In newborn infants,
circadian rhythms are poorly developed, although the existence of a circadian
rhythm of low-amplitude body temperature has been described. Circadian rhythms
increase in amplitude throughout the first months of life, which is believed to be due
to a more robust circadian signal from the suprachiasmatic nuclei (SCN), probably
because of the increase in synaptic connections among the cells of these nuclei.
During this stage, the sleep pattern and the feeding pattern also mature; the subjects
are more active during wakefulness and their interest in the environment around
them is aroused. All these contribute to the development of circadian rhythms, as
some of these variables act as zeitgebers and promote circadian clocks, all of which
increase exogenous influences on circadian rhythms (masking).
In the elderly, the rhythms are also less marked [50]. There is a decrease in the
clock output signals and an increase in the phase instability of the circadian rhythms
in the population as a whole, alterations that persist during “constant routine” pro-
tocols. SCN neuronal degeneration itself may contribute to these modifications.
With age, the ability to maintain sleep is also diminished, and the increase in the
frequency of urination at night, as renal circadian rhythms deteriorate, results in a
274 6  Fourth Level: The Limbic System

decrease in the production of urine. The phase of the circadian rhythms moves
toward earlier hours of the day, and the older individuals tend to behave more like
“larks” [50].
Generally, the variations in the phase and amplitude mentioned above do not
cause serious problems to the individuals. Even when difficulties appear, they
appear to be rather a reflection of a general pattern of development and deterioration
of the body at various stages of life; thus, we can hardly call them “anomalous.”
However, there are examples, in both young and old, where the departure from the
norm is more marked, and for which the term “anomalous” would seem more
appropriate. Some children have nocturnal enuresis. It may be due to inappropriate
development of the circadian rhythm of AVP secretion: very little hormone is
released at night. In some elderly people, getting up at night to urinate also becomes
a problem, in this case because of the decreased ability of the bladder to store large
volumes of urine and an altered rhythm of diuresis.
Sleep disturbances are a symptom that causes great concern for the individual
who suffers them, and in fact, some types of insomnia seem to be produced by
alterations of the circadian system. One of the most frequent of these circadian
alterations is the delayed sleep phase syndrome (DSPS) and related to it, although
less frequent, is the advanced sleep phase syndrome (APSP) [54]. In these altera-
tions, individuals wish to sleep at hours that are either too late (e.g., 04:00 h in the
case of DSPS) or too early (e.g., at 19:00 h in the case of ASPS) for a conventional
work and life schedule. In both cases, the phases of all circadian rhythms appear to
be affected in the same way (temperature, melatonin, cortisol), although it is pre-
cisely the problem with sleep that generates more discomfort and therefore the one
most likely to be detected in the patient. When these individuals can sleep at will,
they do not seem to have problems sleeping normally (provided their rooms are
isolated from the noise and outside light that normally exists during sleep during the
day). The problem is derived from an anomalous phase relationship between the
circadian clock and the environmental synchronizers, although the motifs are
unknown. A mutation in certain clock genes may be behind these alterations [54].
In some individuals, circadian alterations occur because the circadian clock is
unable to adjust to the solar day. This problem occurs occasionally in people who can
see, but it is quite common in blind people, a fact that supports the idea that the normal
light–dark cycle is an important synchronizer (Chap. 2). In this alteration, all circadian
rhythms are maintained, including those that affect the ability to sleep and be awake,
but show a period longer than 24 h. As the pace of the body rhythms shifts from that
which would be appropriate for a normal lifestyle, problems with night-time sleep loss
and fatigue during the day worsen, and become more marked at the time that body
rhythms and lifestyle are maximally out of phase (i.e., 12 h; periodic insomnia) [54].

24-h Rhythms and Learning and Memory

The most important acquired determinant to modify human behavior is learning,


and the consequence of its persistence, or memory. These processes are more per-
sistent the earlier in life they are acquired. Hence, the relationship of the newborns
24-h Rhythms and Learning and Memory 275

with their mothers is of importance. A child with a bad or insufficient affective


relationship in the early stages of his development is prone to presenting alteration
in his emotional and ANS reactivity as an adult [1].
The concepts of memory and learning are closely related. Learning is a process
through which new information is acquired. Memory refers to the persistence of
what is learned, in a state that can be evoked later. In this sense, memory is the result
of learning [55].
One commonly used classification of memory considers its persistence: (a)
short-term memory; (b) long-lasting memory. The former refers to systems that
retain information temporarily, whereas the latter imply permanent information
(Fig. 6.21) [56].
The short-term memory is also called working memory. It is characteristic of the
prefrontal association cortex and comprises a workspace in which information is
maintained while it is processed. The information may come from long-term mem-
ory or from newly acquired information that is being incorporated or being used for
short periods of time and then discarded.
The long-term memory is divided into two processes (Fig. 6.21): (a) declarative
memory; (b) reflexive or procedural memory.
Declarative memory involves the cognitive mechanisms by which a past event is
recalled, with the possibility or not of verbal expression. It constitutes the memory,
which when lost, results in an individual commonly recognized as an amnesiac.
The reflexive or procedural memory refers to the learning process of the subcor-
tical type and that does not require participation of cognition. It may include motor
and ANS components. In the case of motor learning, it implies the different nuances,

Short term / Working memory Reflexive / Procedural (knowing how)


Characteristics: Characteristics:
o sort of scratch pad which allows for o includes skills such as biking or how to tie
temporary storage of information shoelaces
o involves tonic activity of neurons in the o established slowly by practice
frontal lobe o one is not conscious of remembering the
o has a very limited capacity. skill
o starts to develop at birth
o not affected in amnesia
o occurs in much of the CNS.
Declarative (knowing that)
Characteristics:
o representations of objects and events
o involves associations e.g. name with face
o affected by amnesia
o conscious of memory starts after the age
of 2 years
o often established in one trial
o memory formation (learning) requires the
o hippocampus.

Fig. 6.21  Short-term, declarative, and reflexive memories. The figure was prepared in part using
image vectors from Servier Medical Art (www.servier.com), licensed under the Creative Commons
Attribution 3.0 Unported License (http://creativecommons.org/license/by/3.0/)
276 6  Fourth Level: The Limbic System

fixed by experience and repetition, of a motor action. Both motor and ANS learning
are linked to cerebellar function.
How are these types of memory revealed? A suitable test for analyzing proce-
dural memory is to train a person to read inverted words (i.e., reflected in a mir-
ror). Normal individuals require on average two sessions to acquire this capacity,
which is maintained for about 30 days and then extinguishes. The patient com-
monly recognized as an amnesiac (i.e., who suffers from declarative amnesia)
performs normally on the reverse reading test, although he does not even remem-
ber that he has participated in the training sessions. That is, the mechanisms of
procedural and declarative memory differ from each other and can be affected
independently [57].
Classical conditioning is the basis of reflexive or procedural memory. The
declarative memory, on the other hand, implies the mechanisms of fixation of the
experience usually recognized as “memory” [58, 59].
Engram defines the set of neural changes that occur during the memory process.
The engrams are the result of learning, and comprise biochemical and structural
changes in the participating neural circuits. In general, they represent a modification
of the synaptic efficacy of such circuits.
The memory lacks a cerebral location (there is no “memory center”). On the
contrary, memory is the result of information processing and is a change, perma-
nent, in the same neural circuits that process sensory information. For example, in
the visual system, the inferotemporal cortex (one of the last areas in the process of
form analysis) is, in addition to a secondary sensory cortex, a place for the storage
of visual engrams.
One way to demonstrate the storage of engrams in high-order sensory cortical
areas is by microelectrode stimulation of the cortex. Intraoperative stimulation of
the primary auditory cortex (Brodmann areas 41 and 42) produces noises, that is,
elemental sensory sensations. When the secondary auditory cortex, i.e., area 22
(Wernicke’s area in the dominant hemisphere) is stimulated, complex sensations
occur (audible words in the dominant hemisphere, melodies in the nondominant
hemisphere, etc.).
The physiological reason for the fixation of memory engrams, which constitute
a very small portion of the mass of information circulating in the sensory processing
areas, is that the simultaneous activation of the (limbic) motivational system occurs.
In fact, one remembers what has had a certain emotional, conscious or unconscious
meaning [60].
When studying the correlation between the clinical picture and the underly-
ing pathological condition in amnesiac patients, it can be verified that bilateral
damage of certain brain areas makes it impossible to establish new memories
(anterograde memory) and to remember (retrograde memory), although this lat-
ter type of amnesia disappears after a certain time. These brain areas are: (a) the
medial temporal lobe area (hippocampus, amygdala); the medial area of the
24-h Rhythms and Learning and Memory 277

diencephalon (hypothalamic mammillary body, mediodorsal nucleus of the thal-


amus) [61].
The surgical lesion of the hippocampus in humans produces declarative amnesia
with anterograde memory deficit. The afferents to this region come from the ento-
rhinal cortex, the contralateral hippocampus, and from subcortical structures such
as the medial septum, certain raphe nuclei, and the LC from the brainstem (Fig. 6.6).
The dentate gyrus is the area where information enters the hippocampus. Both den-
tate gyrus and hippocampus are phylogenetically old cortical areas consisting of
three layers (allocortex). Located in the dentate gyrus are the molecular layer, the
grain cell layer (which are excitatory and give rise to mossy fibers), and the hilum,
with a population of GABAergic interneurons. The three-layer allocortical structure
of the hippocampus comprises: (a) the molecular layer, where the apical dendrites
of pyramidal neurons reside; (b) the pyramidal or main layer, with its population of
excitatory cells that give rise to the Schaffer collaterals; (c) a polymorphic layer
containing GABAergic interneurons. There are several cornu ammonis (CA) areas:
CA1, CA2, and CA3 (Fig. 6.6) [56].
It was in the hippocampus where the long-term potentiation phenomenon, which
is fundamental to the learning process, was first described [56]. Long-term potentia-
tion consists in changing the efficacy of certain synapses (as already described for
synapses in the dentate gyrus or in the CA1 layer) depending on their previous
activity. Potentiation is induced by brief, high-frequency stimulation of the perforat-
ing pathway (entrance to the hippocampus) both in vivo and in vitro. The process
becomes manifested in a matter of seconds and its effect is maintained for weeks in
the animal under physiological conditions. Long-term potentiation presents as a
characteristic the possibility of associating stimuli from various sources. This prop-
erty allows to the effect of a weak stimulus to be enhanced only when its presence
is associated with that of a more intense stimulus. This situation is very like that of
classical or Pavlovian conditioning.
Figure 6.22 outlines the basis of long-term potentiation. Normal activation
of the neural pathway, which uses Glu as a neurotransmitter, and in the presence
of N-methyl-D-aspartate (NMDA) and non-NMDA glutamatergic receptors in
the postsynapsis (Chap. 3) produces a predominant postsynaptic response
through non-­NMDA receptors. This is because NMDA receptors remain inac-
tive in the presence of Mg2+ at certain membrane potential values. If other depo-
larizing stimuli of pathways affecting the same postsynapsis are simultaneously
associated, the membrane potential value is exceeded to overcome the inhibi-
tion given by Mg2+, which activates the NMDA receptors (Fig. 6.22). This pro-
duces Ca2+ entry with a self-­regenerating depolarization. This flow of Ca2+
through the NMDA receptor is critical for long-term potentiation to appear.
Activation of protein kinase C and calmodulin kinase are the next necessary
steps for this phenomenon. When the conditioned and unconditioned pathways
are activated in a contingent relation, the depolarizing effect allows expression
278 6  Fourth Level: The Limbic System

Presynaptic
membrane

Synaptic cleft
Neurotransmitter
Postsynaptic
lonic channel
membrane

Glutamate

Ca2+

Ca2+
NMDA Na+
receptor
Na+ Second release
Despolarización of glutamate LTP

Fig. 6.22  The proposed ionic mechanism for long-term potentiation (LTP). In the second release
of glutamate, calcium enters the terminal. This results in increased sodium entry during the follow-
ing depolarizations. Modified with permission from Cardinali [1]

of NMDA receptor action (increase in Ca2+ entry) and consequently engram


fixation. If two synapses fire at the same time (synchronously) they produce a
larger depolarization than if they fire at different times (asynchronously). Cells
that fire together wire together, whereas signals from cells that arrive at differ-
ent times become weaker. This is the basis for plasticity or learning throughout
the cortex [56].
The typical electroencephalography rhythm of the hippocampus is theta
(4–10 Hz). This rhythm is related to projections of the medial septum and the
diagonal band of Broca and coincides with low activity in the CA3 and CA1 layers
of the hippocampus. When theta rhythm disappears, there are irregular waves in
CA3 and CA1 concomitant with consummatory behaviors in animals, such as
drinking or eating.
The electrical activity of CA3 pyramidal cells increases as a function of the nov-
elty of the stimulus. The CA1 pyramidal cells are activated to the location of the
animal in a specific and known place of its spatial environment. Through studies in
24-h Rhythms and Learning and Memory 279

animals with an injured hippocampus, it has been possible to establish the function
of this region in spatial memory.
As stated above, short-term memory is intact in declarative amnesiacs. It is the
long-time memory that disappears. Thus, for example, the amnesiac patient can
normally hold a list of numbers for several minutes if he keeps his attention on the
test, but loses it immediately if he is distracted. In lesions that produce declarative
amnesia, there is no modification of the memory already acquired, but there is
impairment of fixation of new engrams. Proof of this is that in patients with a bilat-
eral lesion of the hippocampus, the stored memory does not change, nor does it alter
the short-term memory, which implies mechanisms independent of the limbic
system.
The capacity for the long-term storage of engrams depends both on plastic
neural changes in the higher sensory processing area and on the integrity of the
motivational circuits linked to the limbic system (the subject remembers only
what was fixed with some emotional content). The sensory information above
the cerebral cortex is fixed as a declarative engram if the simultaneous activa-
tion of the motivational system occurs. The constituent parts of the limbic
system that most influence the memory process are the amygdala, the septum,
the reticular formation, certain portions of the hypothalamus, and hippocampal
formation [3].
The flow of information in this motivational circuit is modulated by the cholin-
ergic projection from Meynert’s basal nucleus to the neocortex and by the middle
septal nucleus to the hippocampus (septo-hippocampal projections). This choliner-
gic projection is compromised in senile dementia or AD and these patients also
show significant chronobiological and sleep disturbances.
Likewise, noradrenergic projections from LC, and beta endorphins from the
hypothalamus, are involved in the process of memory fixation. Hormonal influ-
ences, such as those provided by ACTH, AVP, oxytocin or circulating catechol-
amines, modulate memory by action at the level of the reticular formation [60, 62].
We have already discussed how corticosteroids have a deleterious effect on hippo-
campal circuits (Fig. 5.19).
The homeostatic value of forgetting is essential for a normal life. It would be
absurd to recall every detail of our experience, as we would not have time to live in
the present! The Argentine writer Jorge Luis Borges exemplified this fact in his
story “Funes, El Memorioso,” an individual unable to forget a second of every sec-
ond lived, and who remembered the last of the veins of the leaves of each tree he had
seen…
Another fundamental fact is the maintenance of the engrams. We again recount
Borges, who describes his memories of Adrogué, a place near Buenos Aires, where
he spent vacations as a child:
…When I think of Adrogué, I do not think of the present Adrogué, deteriorated by progress,
radiotelephony and motorcycles, but in that lost and tranquil labyrinth of country houses,
squares and streets that converged and diverged, full of Eucalyptus trees… Wherever I am
in the world, the smell of the Eucalyptus is enough for me to return to that lost Adrogué,
which now exists only in my memory…
280 6  Fourth Level: The Limbic System

How did the association between the aroma of Eucalyptus and Adrogué in
Borges’ brain did not fade? It should be noted that the neocortex is an area in con-
tinuous plastic change and that if the neural circuits are not activated, they disap-
pear. The most logical moment for this activation is during the “hallucinosis” of
REM sleep. Therefore, REM sleep is presumably responsible for the “service” of
memory [1].
The role of sleep in the process of learning and memory is thus of great impor-
tance. It is noteworthy that during sleep the activation of areas used in the learning
process in the previous wake period is repeated. In experiments of cortical neuro-
plasticity induced by monocular deprivation in the cat during the critical period of
development (30 days of age) it was observed that for such plastic changes to occur
REM sleep must be present. It is possible that REM sleep (by its visual imagery)
provides an input analogous to the visual cortex stimulus given by the normal visual
input.
There is evidence for a specific output of information from the hippocampus to
the neocortex during dreaming. This phenomenon is due to cholinergic activation,
which is maximal in wakefulness and REM sleep and minimal in slow-wave sleep.
During wakefulness, the exit information from the neocortex is projected onto the
entorhinal association cortex and hippocampus, whereas the reverse flow from the
hippocampus to the cortex is attenuated (Fig. 6.23) [63].
During slow-wave sleep, information flows mainly from the hippocampus to the
neocortex where, over time, memory engrams originate. During REM sleep, the exit
from the hippocampus to the neocortex is blocked as in wakefulness, the flow of
information from the neocortex to the hippocampus being maintained (Fig. 6.23).
Thus, during slow-wave sleep the hippocampus consolidates unstable memories
and transfers information to the cortex for long-term storage (these are the most
favorable conditions for long-term potentiation and for synaptic plasticity).
There is evidence that the various types of memory are related differently to the
type of sleep. The function of REM sleep is to reinforce the cortical plasticity
involved in procedural memory and high-level cognitive processes (semantic mem-
ory), with little or no function in episodic memory [64].
In contrast, episodic memory depends on NREM sleep. The sources of memory
in this period are mainly episodes experienced the previous day (episodic memory),
whereas, in contrast, during REM sleep, memory sources are a mixture of episodic
memories and semantics with emotional content. There are currently two hypothe-
ses regarding the link between memory and sleep:

• For the dual process hypothesis, NREM and REM sleep facilitate different mem-
ory processes: slow-wave sleep facilitates declarative or explicit memory whereas
REM sleep facilitates procedural memory, nondeclarative memory, and semantic
and emotional memory. In favor of this hypothesis, NREM sleep deprivation
selectively decreases performance in word association tasks or spatial memory
tests whereas REM sleep deprivation damages performance in procedural tasks
such as inverted writing in a mirror.
24-h Rhythms and Learning and Memory 281

In waking and REM sleep information flows from


the neocortex to the hippocampus

Hippocampus
In slow wave sleep information flows
from the hippocampus to neocortex

Hippocampus

Fig. 6.23  The flow of information to and from the hippocampus is different in sleep and wakeful-
ness. Cholinergic activation of an aminergically demodulated cortex during REM sleep, or the
combined cholinergic and aminergic activation during wakefulness, block the passage of informa-
tion from the hippocampus to the neocortex. In slow-wave sleep, this flow is facilitated by the
aminergic effect in the presence of inhibition of the cholinergic input. Modified with permission
from Cardinali [1]

• For the sequential hypothesis, the different sleep phases consolidate a memory
trail in a consecutive and complementary way. There would be no functional dif-
ferentiation in the role of each type of sleep in the various categories of
memories.

Finally, a very important aspect regarding memory occurs during REM sleep.
We suddenly remember things that happened to us a long time ago and that had
“disappeared” from our daily memory. It is known that no neural network
remains if unused. A patient with amputation of one of its members experiences
it as a “phantom limb” for a while until the neuroplastic changes leading to the
disuse of the neural circuits representing the amputated limb sensibility are
completed.
How then are the distant memories kept? In part or in whole, their engrams must
be activated periodically so that they do not disappear. Dreaming guarantees this
“service” of memory, contributing both to its maintenance and its reworking (we
remember in general only what is pleasant; thus, “all past times were better”;
Fig. 6.24).
Neuroimaging methods are currently used experimentally to understand the
intricacies of the memory and plasticity processes, in addition to clinically assess-
ing the decline of several cognitive processes that result from changes in defined
282 6  Fourth Level: The Limbic System

Input: Working Output:


sensations memory actions
Limited
1
capacity
Frontal lobe

2
Limbic emotional 4
tone Consolidation
hippocampus Remembering
(amygdala)

5 3
Long term
REM sleep Memory
High
“technical capacity
service”of Secondary and
memory association
cortexes

Fig. 6.24  Sensory information flowing through working memory ① is only consolidated in the
hippocampus by simultaneous activation of the motivational system (limbic system) ②. Long-term
memory, located in high-order or secondary sensory and associated cortexes, requires periodic
maintenance by the implementation of parts or the entire engram ③. This “memory service” is
most likely effected during REM sleep ⑤. Remembering ④ no longer requires the hippocampus.
Modified with permission from Cardinali [1]

neural circuits in the aging brain [65–67]. Advanced magnetic resonance imaging
techniques, such as functional connectivity magnetic resonance imaging, diffusion
tensor imaging, and magnetic resonance spectroscopy, and molecular imaging tech-
niques, such as 18F fluoro-deoxy glucose PET, amyloid PET, and tau PET, are avail-
able for experimental and clinical use.
By using these methodologies, the functional neural architecture of working
memory is described as an interaction of the frontal-parietal control network and
more posterior areas in the ventral visual stream. In addition, several studies have
demonstrated a link between age-related episodic memory decline and the hippo-
campus during active mnemonic processing, which is further supported by studies
of hippocampal functional connectivity in the resting state. The hippocampus inter-
acts with anterior and posterior neocortical regions to support episodic memory, and
alterations in hippocampus–neocortex connectivity have been shown to contribute
to impaired episodic memory.
American and European guidelines recommend imaging to exclude treatable
causes of dementia, such as tumor, hydrocephalus, or intracranial hemorrhage, and
to distinguish between different dementia subtypes, the commonest of which is
Alzheimer’s disease. As the hallmark feature of dementia is that of irreversible
References 283

cognitive decline, usually affecting memory, and impaired activities of daily living,
intervention at the preclinical stages before irreversible brain damage occurs is cur-
rently the best hope of reducing the impact of dementia [65–67].

References
1. Cardinali DP. Neurociencia Aplicada. Sus fundamentos. Editorial Médica Panamericana:
Buenos Aires; 2007.
2. Waxman SG. Chapter 19. The Limbic system. In: Clinical neuroanatomy. 27th ed. New York:
The McGraw-Hill Companies; 2013.
3. LeDoux JE, Damasio AR. Emotions and feelings. In: Kandel E, Schwartz J, Jessell T,
Siegelbaum S, Hudspeth AJ, editors. Principles of neural science. 5th ed. New York: McGraw-­
Hill; 2012. p. 1079–94.
4. Ropper AH, Samuels MA, Klein JP. Chapter 25. The Limbic lobes and the neurology of emo-
tion. In: Adams and Victor’s principles of neurology. 10th ed. New York: The McGraw-Hill
Companies; 2014.
5. Lovblad KO, Schaller K, Vargas MI. The fornix and limbic system. Semin Ultrasound CT MR.
2014;35:459–73.
6. Roxo MR, Franceschini PR, Zubaran C, Kleber FD, Sander JW. The limbic system conception
and its historical evolution. ScientificWorldJournal. 2011;11:2428–41.
7. LeDoux JE. Emotion circuits in the brain. Annu Rev Neurosci. 2000;23:155–84.
8. Calhoon GG, Tye KM. Resolving the neural circuits of anxiety. Nat Neurosci.
2015;18:1394–404.
9. Bello-Morales R, Delgado-Garcia JM. The social neuroscience and the theory of integrative
levels. Front Integr Neurosci. 2015;9:54.
10. Critchley HD, Eccles J, Garfinkel SN. Chapter 6 – Interaction between cognition, emotion, and
the autonomic nervous system. In: Buijs RM, Swaab D, editors. Handbook of clinical neurol-
ogy, autonomic nervous system. New York: Elsevier; 2013. p. 59–77.
11. Lippe S, Gonin-Flambois C, Jambaque I. The neuropsychology of the Kluver-Bucy syndrome
in children. Handb Clin Neurol. 2013;112:1285–8.
12. Bergstrom HC. The neurocircuitry of remote cued fear memory. Neurosci Biobehav Rev.
2016;71:409–17.
13. Koob GF, Volkow ND. Neurobiology of addiction: a neurocircuitry analysis. Lancet Psychiatry.
2016;3:760–73.
14. Brooks SJ, Stein DJ. A systematic review of the neural bases of psychotherapy for anxiety and
related disorders. Dialogues Clin Neurosci. 2015;17:261–79.
15. Modinos G, Ormel J, Aleman A. Individual differences in dispositional mindfulness and brain
activity involved in reappraisal of emotion. Soc Cogn Affect Neurosci. 2010;5:369–77.
16. Holzel BK, Lazar SW, Gard T, Schuman-Olivier Z, Vago DR, Ott U. How does mindfulness
meditation work? Proposing mechanisms of action from a conceptual and neural perspective.
Perspect Psychol Sci. 2011;6:537–59.
17. Damasio AR. Descartes’ error: emotion, reason and the human brain. New York: Grosset/
Putnam; 1994.
18. Bader O. Attending to emotions is sharing of emotions – a multidisciplinary perspective to
social attention and emotional sharing. Comment on Zahavi and Rochat (2015). Conscious
Cogn. 2016;42:382–95.
19. Reznikov R, Binko M, Nobrega JN, Hamani C. Deep brain stimulation in animal models of
fear, anxiety, and posttraumatic stress disorder. Neuropsychopharmacology. 2016;41:2810–7.
20. Murray EA, Gaffan D, Mishkin M. Neural substrates of visual stimulus-stimulus association
in rhesus monkeys. J Neurosci. 1993;13:4549–61.
284 6  Fourth Level: The Limbic System

21. Dricu M, Fruhholz S. Perceiving emotional expressions in others: activation likelihood esti-
mation meta-analyses of explicit evaluation, passive perception and incidental perception of
emotions. Neurosci Biobehav Rev. 2016;71:810–28.
22. James W. Physical basis of emotion (reprinted in 1994. Psychol rev 101: 205–210). Psychol
Rev. 1894;1:516–29.
23. Cannon WB. The James-Lange theory of emotions: a critical examination and an alternative
theory. By Walter B. Cannon, 1927. Am J Psychol. 1987;100:567–86.
24. Harrison NA, Singer T, Rotshtein P, Dolan RJ, Critchley HD. Pupillary contagion: central
mechanisms engaged in sadness processing. Soc Cogn Affect Neurosci. 2006;1:5–17.
25. Makovac E, Garfinkel SN, Bassi A, Basile B, Macaluso E, Cercignani M, Calcagnini G, Mattei
E, Agalliu D, Cortelli P, Caltagirone C, Bozzali M, Critchley H. Effect of parasympathetic stim-
ulation on brain activity during appraisal of fearful expressions. Neuropsychopharmacology.
2015;40:1649–58.
26. Lavina B. Brain vascular imaging techniques. Int J Mol Sci. 2016;18:70.
27. Im JJ, Namgung E, Choi Y, Kim JY, Rhie SJ, Yoon S. Molecular neuroimaging in posttrau-
matic stress disorder. Exp Neurobiol. 2016;25:277–95.
28. Biotteau M, Chaix Y, Blais M, Tallet J, Peran P, Albaret JM. Neural signature of DCD: a criti-
cal review of MRI neuroimaging studies. Front Neurol. 2016;7:227.
29. Herting MM, Sowell ER. Puberty and structural brain development in humans. Front

Neuroendocrinol. 2017;44:122–37.
30. Wandschneider B, Koepp MJ. Pharmaco fMRI: determining the functional anatomy of the
effects of medication. Neuroimage Clin. 2016;12:691–7.
31. Vinik AI, Nevoret ML, Casellini C. The new age of sudomotor function testing: a sensitive and
specific biomarker for diagnosis, estimation of severity, monitoring progression, and regres-
sion in response to intervention. Front Endocrinol (Lausanne). 2015;6:94.
32. Guinjoan SM, Bonanni Rey RA, Cardinali DP. Correlation between skin potential response and
psychopathology in patients with affective disorders. Neuropsychobiology. 1995;31:24–30.
33. Pérez Lloret S, Diez JJ, Domé MN, Alvarez Delvenne A, Braidot N, Cardinali DP, Vigo
DE. Effects of different “relaxing” music styles on the autonomic nervous system. Noise
Health. 2014;16:279–84.
34. Nonnekes J, Carpenter MG, Inglis JT, Duysens J, Weerdesteyn V. What startles tell us about
control of posture and gait. Neurosci Biobehav Rev. 2015;53:131–8.
35. Wichmann T, DeLong MR. The basal ganglia. In: Kandel E, Schwartz J, Jessell T, Siegelbaum
S, Hudspeth AJ, editors. Principles of neural science. 5th ed. New York: McGraw-Hill; 2012.
36. Waxman SG. Control of movement. In: Clinical neuroanatomy. 28th ed. New York: McGraw-­
Hill Education; 2017.
37. Saper CB. The central autonomic nervous system: conscious visceral perception and auto-
nomic pattern generation. Annu Rev Neurosci. 2002;25:433–69.
38. Horn JP, Swanson LW. The autonomic motor system and the hypothalamus. In: Kandel E,
Schwartz J, Jessell T, Siegelbaum S, Hudspeth AJ, editors. Principles of neural science. 5th ed.
New York: McGraw-Hill; 2012. p. 1056–78.
39. Macey PM, Ogren JA, Kumar R, Harper RM. Functional imaging of autonomic regulation:
methods and key findings. Front Neurosci. 2015;9:513.
40. Wu P, Bandettini PA, Harper RM, Handwerker DA. Effects of thoracic pressure changes on
MRI signals in the brain. J Cereb Blood Flow Metab. 2015;35:1024–32.
41. Benarroch EE. The central autonomic network: functional organization, dysfunction, and per-
spective. Mayo Clin Proc. 1993;68:988–1001.
42. Harper RM, Bandler R, Spriggs D, Alger JR. Lateralized and widespread brain activation
during transient blood pressure elevation revealed by magnetic resonance imaging. J Comp
Neurol. 2000;417:195–204.
43. James C, Macefield VG, Henderson LA. Real-time imaging of cortical and subcortical control
of muscle sympathetic nerve activity in awake human subjects. Neuroimage. 2013;70:59–65.
44. James C, Henderson L, Macefield VG. Real-time imaging of brain areas involved in the gen-
eration of spontaneous skin sympathetic nerve activity at rest. Neuroimage. 2013;74:188–94.
References 285

45. Lisberger SG, Thacht WT. The cerebellum. In: Kandel E, Schwartz J, Jessell T, Siegelbaum
S, Hudspeth AJ, editors. Principles of neural science. 5th ed. New York: McGraw-Hill; 2012.
p. 960–81.
46. De Morree HM, Rutten GJ, Szabo BM, Sitskoorn MM, Kop WJ. Effects of insula resection on
autonomic nervous system activity. J Neurosurg Anesthesiol. 2016;28:153–8.
47. Strata P. The emotional cerebellum. Cerebellum. 2015;14:570–7.
48. Roenneberg T, Keller LK, Fischer D, Matera JL, Vetter C, Winnebeck EC. Human activity and
rest in situ. Methods Enzymol. 2015;552:257–83.
49. Wittmann M, Dinich J, Merrow M, Roenneberg T. Social jetlag: misalignment of biological
and social time. Chronobiol Int. 2006;23:497–509.
50. Gaggioni G, Maquet P, Schmidt C, Dijk DJ, Vandewalle G. Neuroimaging, cognition, light and
circadian rhythms. Front Syst Neurosci. 2014;8:126.
51. Lim AS, Chang AM, Shulman JM, Raj T, Chibnik LB, Cain SW, Rothamel K, Benoist C,
Myers AJ, Czeisler CA, Buchman AS, Bennett DA, Duffy JF, Saper CB, De Jager PL. A
common polymorphism near PER1 and the timing of human behavioral rhythms. Ann Neurol.
2012;72:324–34.
52. Roenneberg T, Allebrandt KV, Merrow M, Vetter C. Social jetlag and obesity. Curr Biol.
2012;22:939–43.
53. Thun E, Bjorvatn B, Flo E, Harris A, Pallesen S. Sleep, circadian rhythms, and athletic perfor-
mance. Sleep Med Rev. 2015;23:1–9.
54. Baron KG, Reid KJ. Circadian misalignment and health. Int Rev Psychiatry. 2014;26:139–54.
55. Mesulam MM. Aphasia, memory loss, and other focal cerebral disorders. In: Kasper D, Fauci
A, Hauser S, Longo D, Jameson JL, Loscalzo J, editors. Harrison’s principles of internal medi-
cine. 19th ed. New York: McGraw-Hill Education; 2015.
56. Barrett KE, Barman SM, Boitano S, Brooks HL. Learning, memory, language. In: Ganong’s
review of medical physiology. 25th ed. New York: McGraw-Hill Education; 2016.
57. Waxman SG. Chapter 21. Higher cortical functions. In: Clinical neuroanatomy. 27th ed.
New York: The McGraw-Hill Companies; 2013.
58. Sweatt JD. Neural plasticity and behavior – sixty years of conceptual advances. J Neurochem.
2016;139(Suppl 2):179–99.
59. Connor DA, Gould TJ. The role of working memory and declarative memory in trace condi-
tioning. Neurobiol Learn Mem. 2016;134 Pt B:193–209.
60. Mammarella N, Di DA, Palumbo R, Fairfield B. Noradrenergic modulation of emotional mem-
ory in aging. Ageing Res Rev. 2016;27:61–6.
61. Ropper AH, Samuels MA, Klein JP. Chapter 21. Dementia, the amnesic syndrome, and the
neurology of intelligence and memory. In: Adams and Victor’s principles of neurology. 10th
ed. New York: The McGraw-Hill Companies; 2014.
62. Kirsch P. Oxytocin in the socioemotional brain: implications for psychiatric disorders.

Dialogues Clin Neurosci. 2015;17:463–76.
63. Walker MP. The role of slow wave sleep in memory processing. J Clin Sleep Med.

2009;5:S20–6.
64. Carskadon MA, Dement WC. Chapter 2 – Normal human sleep: an overview. In: Principles
and practice of sleep medicine. 5th ed. Philadelphia: WB Saunders; 2011. p. 16–26.
65. Narayanan L, Murray AD. What can imaging tell us about cognitive impairment and demen-
tia? World J Radiol. 2016;8:240–54.
66. Febo M, Foster TC. Preclinical magnetic resonance imaging and spectroscopy studies of mem-
ory, aging, and cognitive decline. Front Aging Neurosci. 2016;8:158.
67. Rana M, Varan AQ, Davoudi A, Cohen RA, Sitaram R, Ebner NC. Real-time fMRI in neuro-
science research and its use in studying the aging brain. Front Aging Neurosci. 2016;8:239.
Clinical Implications of the Enlarged
Autonomic Nervous System 7

Abstract
Since the ANS is extensively involved in the function of almost every organ sys-
tem, the clinical manifestations of autonomic dysfunction are diverse. Indeed,
the ANS is involved in most diseases. Any structural pathological process affect-
ing the brain (whether infectious, inherited, neoplastic, or degenerative in nature)
can result in an autonomic syndrome. This Chapter describes the ANS semiol-
ogy and the different classifications of the ANS disorders, emphasizing that
derived from the hierarchical organization of the ANS.

Keywords
Amyloidotic autonomic failure • Autoimmune autonomic ganglionopathy •
Autonomic disturbances in spinal cord injuries • Autonomic dysfunction •
Autonomic dysfunction associated with aging • Autonomic dysfunction in pri-
mary sleep disorders • Autonomic tests • Fibromyalgia and chronic fatigue •
Guillain–Barré syndrome • Hereditary autonomic neuropathies • Paraneoplastic
autonomic dysfunction • Peripheral neuropathies • α-Synucleinopathies

Objectives
After studying this chapter, you should be able to:
• Describe the ANS semiology, including bedside evaluation and the most
important autonomic tests to perform.
• Describe the different classifications of the ANS disorders, emphasizing
that derived from the hierarchical organization of the ANS discussed in
this book.
• Give examples of autonomic entities, describing the functions of the ANS
affected.

© Springer International Publishing AG 2018 287


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1_7
288 7  Clinical Implications of the Enlarged Autonomic Nervous System

Semiological Aspects of ANS Disorders

As discussed in previous chapters, the ANS is extensive and is involved in the func-
tion of almost every organ system. Therefore, the clinical manifestations of auto-
nomic dysfunction are diverse. Indeed, the ANS is involved in most diseases. Any
structural pathological process affecting the brain (whether infectious, inherited,
neoplastic, or degenerative in nature) can result in an autonomic syndrome.
Although autonomic disorders are a well-defined group of conditions affecting
the central and/or peripheral autonomic pathways, autonomic symptoms and signs
are often seen in many other medical conditions or may be isolated manifestations
of a limited autonomic instability (Table 7.1) [1–3]. To evaluate these conditions, it
is necessary to cover each autonomic sector (cardiovascular, gastrointestinal, geni-
tourinary, secretomotor, sudomotor, neuroimmunoendocrine), define the temporal
profile, identify associated symptoms, recognize atypical symptoms as expression
of autonomic dysfunction, and exclude other conditions that could mimic its mani-
festations. To properly study these disorders, multiple tests are needed, in addition
to supportive data that often include neuroimaging, sleep studies, specific blood and
urine testing, and tissue diagnosis [4].

Table 7.1  Clinical signs of Cardiovascular Tachycardia at rest


autonomic dysfunction Orthostatic hypotension
Arterial hypertension
Arrhythmias
Syncope
Gastrointestinal tract Dysphagia, regurgitation
Gastroparesis, vomiting
Constipation
Episodes of diarrhea
Sudomotor Hypohidrosis, anhidrosis
Hyperhidrosis
Gustatory sweating
Eye Anisocoria
Nyctalopia
Close blurred vision
Tunnel vision
Double vision
Genitourinary tract Bladder dysfunction
Urinary retention
Incontinence
Impotence
CNS Anxiety
Insomnia
Chronic fatigue
Brain fog
Vertigo
Dizziness
Weakness
Semiological Aspects of ANS Disorders 289

A preliminary evaluation can be done at the bedside, at least for the most disabling
symptoms of autonomic dysfunction. Significant decrement in BP without compensa-
tory tachycardia is much worse prognostically than marked tachycardia without sig-
nificant BP changes (Chap. 4). Secretomotor and sudomotor functions can be assessed
by observation of the mucosae and by appreciating the presence of moisture on the
skin by palpation. Hyperhidrosis is easily appreciated as sweat droplets over the skin
or visibly wet garments. Occasionally, simultaneous ECG monitoring can identify an
ictal bradycardia or asystole, or provide confirmation that convulsive manifestations
commonly seen in syncope are secondary to the hemodynamic changes.
Symptoms suggestive of autonomic dysfunction include, but are not limited to,
postural hypotension, digestive discomfort, altered intestinal, bladder or sexual
function, decreased or increased sweating, dry mucous membranes, and cooling or
discoloration of the extremities. Insomnia, anxiety, brain fog and chronic fatigue
may be present (Table 7.1). In clinical practice, the symptoms of autonomic dys-
function are often underestimated, because they are subjective, frequently transient
in healthy subjects, of slow onset and evolution, of little disability in the patient (at
least in the initial stages), and difficult to treat.
Table 7.2 summarizes the most important autonomic tests that allow the various
autonomic functions to be explored [1–3, 5]. Autonomic tests usually consist of
physical stimuli or actions that elicit changes in sympathetic and parasympathetic
activity, often reflected as BP and heart rate alterations. The Valsalva maneuver is
one autonomic challenge with multiple response phases that influences, to varying
degrees, both sympathetic and parasympathetic outflow. Despite consisting of a
relatively simple somatomotor task, namely exhaling against a resistance to a pre-
determined pressure (30–40 mmHg) for a defined period (15–20 s), the cardiovas-
cular response is separated into four distinct patterns occurring over periods of time,
not including preliminary inhalation:

1. Initial BP rise: on application of expiratory force, pressure rises inside the chest
forcing blood out of the pulmonary circulation into the left atrium. This causes a
mild rise in stroke volume during the first few seconds of the maneuver.
2. Reduced venous return and compensation: return of systemic blood to the heart is
impeded by the pressure inside the chest. The output of the heart is reduced and
systolic volume falls. The fall in systolic volume reflexively causes blood vessels to
constrict with some rise in BP (15–20 s). This compensation can be quite marked,
with BP returning to near or even above normal, but the cardiac output and blood
flow to the body remain low. During this time, a compensatory tachycardia occurs.
3. Pressure release: the pressure on the chest is released, allowing the pulmonary
vessels and the aorta to re-expand, causing a further initial slight fall in systolic
volume (20–23 s) because of the decreased left atrial return and increased aortic
volume respectively. Venous blood can once more enter the chest and the heart
and cardiac output begins to increase.
4. Return of cardiac output: blood return to the heart is enhanced by the effect of
the entry of blood which had been dammed back, causing a rapid increase in
cardiac output (From 24 s on).
290 7  Clinical Implications of the Enlarged Autonomic Nervous System

Table 7.2  Some tests of ANS function


Main part of the reflex arc
Test Parameter tested
Cardiovascular system
RR interval during Heart rate Vagal afferent and efferent
respiration limbs
Heart rate variability Heart rate Vagal afferent and efferent
limbs
Valsalva maneuver Heart rate, BP Afferent and efferent limbs
Mueller maneuver Heart rate, BP Afferent and efferent limbs
BP response to standing or BP Afferent and sympathetic
vertical tilt efferent limbs
Heart rate response to Heart rate Vagal afferent and efferent
standing limbs
Handgrip Heart rate, BP Sympathetic efferent limb
Cold pressor test Heart rate, BP Sympathetic efferent limb
Radiant heating of the trunk Hand blood flow Sympathetic efferent limb
Immersion of the hand in Blood flow of the opposite hand Sympathetic efferent limb
hot water
Emotional stress Heart rate, BP Sympathetic efferent limb
Baroreflex sensitivity Heart rate, BP Vagal afferent and efferent
limbs
Doppler Blood flow Sympathetic efferent limb
Plasma norepinephrine Rises on tilting from horizontal to Sympathetic efferent limb
levels vertical
Plasma arginine Rise with induced hypotension Afferent limb
vasopressin levels
Sudomotor system
Sweat tests Sweat Sympathetic efferent limb
Quantitative sudomotor Evaporation rate Sympathetic efferent limb
axon reflex test
Sympathetic skin response Potentials Sympathetic efferent limb
Pupil
Pharmacological tests of Pupil diameter Afferent and efferent limbs
pupillary innervation
Pupil cycle time Pupil diameter Afferent and efferent limbs
Pupillometry Pupil diameter and latency Afferent and efferent limbs
Other
Microneurography Potentials Sympathetic efferent limb
Sympathetic neuroimaging 123
I-Metaiodobenzylguanidine Sympathetic efferent limb
uptake by NE vesicles in cardiac
sympathetic nerves
Skin biopsy Peripheral adrenergic and Sympathetic efferent limb
cholinergic fibers innervating sweat
glands and arrector pili muscles

A hand grip, which also involves sympathetic activity increases, is another com-
mon autonomic test. Baroreceptor unloading tasks include lower body negative
pressure, which requires a specialized suit, or the simpler to implement Müller’s
maneuver, which is the reverse of the Valsalva manoeuver, consisting in a negative
pressure in the chest and lungs after a forced expiration following by an inspiration
with closed mouth and nose [5].
Semiological Aspects of ANS Disorders 291

The cold pressor is a passive autonomic test that has the advantage of being a
consistent stimulus across subjects, although the pain component may be a con-
founder to BP manipulation. Other tests identify changes in state that last several
minutes, such as the quantitative sudomotor axon reflex test.
Head-up tilt table testing is a provocative test designed to simulate orthostatic
stress and downward gravitational fluid shifts. It is classically used to diagnose neu-
rally mediated (reflex) syncope. Tilt testing can also be used to aid in the diagnosis
of other disorders of orthostatic intolerance, including postural tachycardia syn-
drome and orthostatic hypotension [6].
Changes in cardiac autonomic function can be tracked by several techniques.
The simplest measure of cardiac autonomic status is the resting heart rate. Greater
autonomic dysfunction is associated with increasing resting heart rates over time. A
more robust measure of autonomic function is heart rate variability (HRV), mea-
sured using continuous heart rate monitoring. HRV is a set of parameters that
reflects interval fluctuations between sequential beats of the heart [7–9]. Measures
derived from interval differences between successive beats reflect parasympatheti-
cally modulated changes in heart rate. Other HRV measures reflect the combined
signaling of the two arms of the autonomic nervous system and reflect both intrinsic
(e.g., baroreflex, renin–angiotensin, sleep cycles, circadian) and extrinsic (activity,
rest) rhythms. In general, decreased or decreasing HRV would be a signal for worse
cardiac autonomic dysfunction. However, a higher, but more disorganized, HRV
pattern, detectable by certain “nonlinear” HRV measures also reflects greater car-
diac autonomic dysfunction [10]. Ideally, HRV is measured using 24-h ambulatory
monitoring which can capture both daytime heart rate patterns and heart rate pat-
terns during sleep, providing insights into circadian rhythm, sleep quality, and pos-
sible sleep-disordered breathing or periodic limb movements, all of which affect
cardiac autonomic functioning. However, significant clinical information can be
obtained from shorter recordings performed, perhaps, at the time of clinical visits
and in association with standard bedside autonomic tests.
As mentioned in Chap. 6, changes in skin conductance occur with eccrine sweat-
ing and constitute a relatively pure assay of sympathetic activity (sympathetic skin
response) [11]. Alterations in sweating are mediated by cholinergic nerves and are
not affected by β-adrenoceptor blockers, allowing evaluation of the sympathetic
system. It consists of a potential generated by sweating in the skin in response to
different stimuli, including those that produce emotional reactions [12]. An altera-
tion in the resistance of the skin is caused and a potential is obtained.
Microneurography is a unique method of recording postganglionic sympa-
thetic neural traffic directly from human peripheral nerves. For sympathetic
microneurography, tungsten microelectrodes are inserted through the skin onto
an underlying peripheral nerve, innervating either skin or skeletal muscle.
Sympathetic fibers are spontaneously active and many fibers discharge in syn-
chronized bursts of impulses. Usually, action potentials are recorded simultane-
ously from several fibers (multifiber activity) and presented in a mean voltage
(integrated) neurogram [13].
Sympathetic neuroimaging provides an important supplement to physiological,
neurochemical, and neuropharmacological approaches in the evaluation of patients
with clinical autonomic disorders. Sympathetic neuroimaging to date has involved
292 7  Clinical Implications of the Enlarged Autonomic Nervous System

visualization of noradrenergic innervation in the left ventricular myocardium [14].


Sympathetic imaging depends on the radiolabeling of vesicles in sympathetic
nerves. The most commonly used imaging agent to assess cardiac sympathetic
innervation is 123I-metaiodobenzylguanidine. Cardiac sympathetic neuroimaging
and postmortem neuropathological findings have linked α-synucleinopathy with
noradrenergic denervation in Lewy body disease. Patients with familial Parkinson’s
disease from abnormalities of the gene encoding α-synuclein have cardiac sympa-
thetic denervation.
Cutaneous punch biopsies are widely used to evaluate nociceptive C fibers in
patients with suspected small-fiber neuropathy [15]. Peripheral adrenergic and cho-
linergic fibers innervate several cutaneous structures, such as sweat glands and
arrector pili muscles, and can easily be seen with punch skin biopsies. Skin biopsies
allow for both regional sampling, in diseases with patchy distribution, and the
opportunity for repeated sampling in progressive disorders.

Classification of ANS Disorders

There are several ways to categorize autonomic dysfunctions, depending on the


points of view from which they are considered [1, 2]. Clinical disorders of ANS can
be conceptualized as focal (e.g., Horner’s syndrome, Chap. 3) or generalized, affect-
ing several autonomic segments (such as progressive autonomic failure). Another
possible pathophysiological classification of autonomic dysfunctions is based on
their primary or secondary nature, as summarized in Table 7.3. In addition, it is

Table 7.3  Pathophysiological classification of autonomic dysfunctions on the bases of their


­primary or secondary nature
Primary
Pure autonomic failure
Multisystem atrophy (Shy–Drager syndrome)
Parkinson’s disease
Pan-dysautonomic neuropathy
Paraneoplastic autonomic neuropathy
Secondary
General diseases: diabetes, alcoholism, renal failure, amyloidosis, neoplasms, dysautonomia of
aging
Autoimmune diseases: acute and chronic inflammatory polyneuropathy, Lambert–Eaton
syndrome, rheumatoid arthritis, lupus erythematosus
Metabolic diseases: porphyria, Tangier disease, Fabry disease, pernicious anemia
Hereditary disorders: familial dysautonomia, hereditary motor and sensory neuropathies,
sensorial and autonomic congenital neuropathy, Friedreich’s ataxia
Infections: Chagas disease, AIDS, botulism, leprosy, syphilis
Diseases of the central nervous system: medullary lesions, strokes, tumors, multiple sclerosis,
amyotrophic lateral sclerosis, Adie syndrome
Intoxications: by acrylamide, heavy metals, organic solvents
Pharmacological: by antineoplastics, antidepressants, sedatives, hypotensives, adrenergic
blockers, cholinergic blockers
Some Clinical Autonomic Entities 293

necessary to consider whether the manifestations are predominantly due to a sym-


pathetic, parasympathetic, or mixed dysfunction (pandysautonomia).
Some characteristic patterns, based on temporal evolution and the constellation
of semiology constellation, are also important in the differential diagnosis of auto-
nomic neuropathies. Finally, there are some neurological disorders that affect the
ANS, but in most cases, they are associated with somatic nervous system involve-
ment. ANS dysfunctions may be due to increases or decreases in autonomic control
activity and may occur because of brain, spinal or peripheral nerve injuries.
The enlarged view of ANS discussed in this book leads to a classification of
autonomic disorders compatible with that found in popular clinical textbooks, e.g.,
Low and Engstrom (Table 7.4) [3]:

• Level 1 disorders (spinal) are systematized in those where there is involvement


of the spinal cord (traumatic quadriplegia, syringomyelia, multiple sclerosis,
amyotrophic lateral sclerosis, tumors spinal cord) and in various autoimmune
autonomic neuropathies, paraneoplastic, Guillain–Barré syndrome, botulism,
and porphyria. Autonomic neuropathy by amyloidosis, diabetes, or nutritional
deficiency, and dysautonomia of aging are also included in this group. Frequently,
they co-exist with orthostatic intolerance disorders (syncope, postural orthostatic
tachycardia syndrome, etc.).
• Level 2 disorders include alterations of the brainstem and cerebellum, such as
vertebrobasilar and Wallenberg syndromes, syringobulbia, and Arnold–Chiari
malformation. Disorders in BP control (hypertension, hypotension) and of heart
rate and central sleep apneas are included in this group.
• Level 3 disorders consist of the alteration of specific hypothalamic behaviors
with their autonomic, neuroendocrine, and behavioral consequences. They
include Wernicke–Korsakoff syndrome, malignant neuroleptic syndrome, fatal
familial insomnia, alterations of AVP release and of temperature regulation
(hyperthermia, hypothermia), disrupted sexual function (Klüver–Bucy syn-
drome, Chap. 6) and appetite disturbances.
• Level 4 disorders involve the abnormal function of limbic and paralimbic circuits
such as autonomic seizures or limbic encephalitis, and various disorders with
involvement of the cerebral cortex (complex partial seizures, cerebral infarction
of the insula).
• Multilevel ANS disorders, such as multiple system atrophy, Parkinson’s disease,
and diffuse Lewy body disease.

Some Clinical Autonomic Entities

Peripheral Neuropathies with Dysautonomia

Peripheral nerves are susceptible to toxic damage, metabolic disorders, trauma, or


neoplasms (neuropathies). In some cases, the axon is the primary focus of injury. The
myelin sheath, or both the sheath and the axon, may be involved. In some cases, the
294 7  Clinical Implications of the Enlarged Autonomic Nervous System

Table 7.4  Clinical classification of autonomic disorders by organizational level (modified from
Low and Engstrom [3])
Level 1 disorders (spinal cord)
Autonomic disorders with spinal cord involvement
   a. Traumatic quadriplegia
   b. Syringomyelia
   c. Multiple sclerosis and neuromyelitis optica
   d. Amyotrophic lateral sclerosis
   e. Tetanus
   f. Spinal cord tumors
Autonomic neuropathies
   A. Acute/subacute autonomic neuropathies
     1. Subacute autoimmune autonomic ganglionopathy
      a. Subacute paraneoplastic autonomic neuropathy
      b. Guillain–Barré syndrome
      c. Lambert–Eaton syndrome
      d. Botulism
      e. Porphyria
      f. Drug-induced autonomic neuropathies
      g. Toxin-induced autonomic neuropathies
   B. Chronic peripheral autonomic neuropathies
     1. Distal small fiber neuropathy
     2. Combined sympathetic and parasympathetic failure
      a. Amyloid
      b. Diabetic autonomic neuropathy
      c. Autoimmune autonomic ganglionopathy (paraneoplastic and idiopathic)
      d. Sensory neuronopathy with autonomic failure
      e. Familial dysautonomia (Riley–Day syndrome)
      f. Uremic or nutritional deficiency
      g. Dysautonomia of old age
     3. Disorders of reduced orthostatic intolerance: idiopathic orthostatic hypotension, reflex
syncope, postural orthostatic tachycardia syndrome, associated with prolonged bedrest,
associated with space flight, chronic fatigue
Level 2 disorders (brainstem and cerebellum)
   a. Vertebrobasilar and lateral medullary (Wallenberg) syndromes
   b. Posterior fossa tumors
   c. Syringobulbia and Arnold–Chiari malformation
   d. Horner’s syndrome
   e. Disorders of blood pressure control (hypertension, hypotension)
   f. Cardiac arrhythmias
   g. Baroreflex failure
   h. Central sleep apnea
   i. Brainstem encephalitis
Level 3 disorders (hypothalamus)
   a. Thiamine deficiency (Wernicke–Korsakoff syndrome)
   b. Diencephalic syndrome
   c. Neuroleptic malignant syndrome
   d. Serotonin syndrome
   e. Fatal familial insomnia
   f. Arginine vasopressin syndromes (diabetes insipidus, inappropriate arginine vasopressin
secretion)
Some Clinical Autonomic Entities 295

Table 7.4 (continued)
   g. Disturbances of temperature regulation (hyperthermia, hypothermia)
   h. Disturbances of sexual function
   i. Disturbances of appetite
Level 4 disorders (focal central nervous system disorders)
   a. Frontal cortex lesions causing urinary/bowel incontinence
   b. Focal seizures (temporal lobe or anterior cingulate)
   c. Cerebral infarction of the insula
   d. Shapiro syndrome (agenesis of the corpus callosum, hyperhidrosis, hypothermia)
   e. Autonomic seizures
   f. Limbic encephalitis
Multilevel autonomic nervous system disorders
Multisystem degeneration: autonomic failure clinically prominent
   a. Multiple system atrophy (Shy–Drager syndrome)
   b. Parkinson’s disease with autonomic failure
   c. Diffuse Lewy body disease (some cases)
Multisystem degeneration: autonomic failure clinically not usually prominent
   a. Parkinson’s disease
   b. Other extrapyramidal disorders (inherited spinocerebellar atrophies, progressive
supranuclear palsy, corticobasal degeneration, Machado–Joseph disease, fragile X syndrome)

proximal portion of the nerve is affected, whereas in others, it is the distal portion.
Neuropathies of a single nerve are called mononeuropathies, in contrast to polyneu-
ropathies, referred to diffuse neural damage throughout the body. The most common
forms of neuropathies are diabetic, renal failure, alcoholic or nutritional cirrhosis,
autoimmune and traumatic diseases [16].
Many neuropathies are characterized by Wallerian degeneration of the segment
distal to the lesion. Those of toxic origin produce a degeneration of the neural distal
segment, especially in the limbs. This explains why the first signs of neuropathy are
detected in the fingers or toes.
The demyelination processes can be primary or secondary, the abnormal demy-
elinated segments being localized or scattered along the axon. Indicating the
importance of the myelin sheath for the transmission of the nerve impulse, the
neural conduction is slowed down, or in some cases halted. These conduction
abnormalities can be identified electrophysiologically by determining the neural
conduction time.
On clinical neurology standpoints, it is common to consider negative or positive
signs or symptoms, that is, those resulting from the inhibition or disappearance of a
function, or from its increase or externalization of an abnormal function. The most
common signs of peripheral neuropathies are negative: loss of strength and sensitiv-
ity. In axonal neuropathies, the signs begin in the feet and progress centrally, cor-
relating with a symptomatology of predominance in the distal portions of the
peripheral nerves. In demyelinating neuropathies, there is also a distal predomi-
nance, with a greater tendency to present foci of demyelination in the long nerves.
In addition to muscle weakness, muscle atrophy and fasciculations, both signs of
denervation, are usually found. Under conditions that affect small nerve fibers, pain
and heat sensitivity are preferentially lost. When neuropathy involves large myelin
296 7  Clinical Implications of the Enlarged Autonomic Nervous System

fibers, there is loss of proprioceptive and vibratory sensitivity. Another early nega-
tive sign is the loss of muscle reflexes [5].
The compromise of the unmyelinated fibers leads to sympathetic alterations,
such as loss of sweating, cardiac arrhythmias, or alteration of cardiovascular control
mechanisms. These alterations are seen only in neuropathies that attack small
myelin fibers (Guillain–Barré syndrome) or unmyelinated fibers (diabetic neuropa-
thy). Another negative symptom of peripheral neuropathies is incoordination
(ataxia), of sensory origin [16].
The most prominent positive symptoms of neuropathies are sensory, or paresthe-
sias, i.e., tingling or other sensations (burns, punctures, vibrations, etc.) originating
without apparent stimulus. Spontaneous painful sensations are more frequent in
small diameter fiber neuropathy.
Peripheral neuropathies may develop acutely or chronically. It may take years for
the appearance of symptoms in the case of metabolic, degenerative, or genetic neu-
ropathies. This is the case of neuropathies seen in diabetes mellitus to lead to renal
failure or chronic exposure. In contrast, neuropathies that are rapidly evolving are
less frequent, such as Guillain–Barré syndrome, which is an autoimmune degenera-
tion of peripheral nerves that evolves in 1–2 weeks.
One of the challenges of managing patients with peripheral neuropathy is that
there are over 100 etiological causes, and not every patient will have an identifiable
etiology, even after extensive investigation. Pain is always an important symptom
to elucidate, as severe burning neuropathic pain is often indicative of C-fiber
involvement and may be the only complaint in patients with small fiber neuropathy
[17]. Patients with painful idiopathic small-fiber neuropathy may have coexistent
autonomic symptoms, but typically the autonomic involvement is subclinical or
minimal at presentation. Physical examination and neurophysiological testing are
helpful in determining whether large fiber or small fiber (C) involvement is pre-
dominant. Formal autonomic testing is helpful in establishing the diagnosis of
peripheral neuropathy in small-fiber predominant neuropathies where nerve con-
duction testing is normal.

α-Synucleinopathies

Several components of the ANS network are affected in neurodegenerative disor-


ders characterized by the presence of intracellular inclusions containing α-synuclein,
a protein that plays a role in maintaining synaptic vesicles in presynaptic terminals
[18]. α-Synucleinopathies include multiple system atrophy, characterized by the
accumulation of glial cytoplasmic inclusions, and Lewy body disorders (i.e.,
Parkinson’s disease, dementia with Lewy bodies, and pure autonomic failure).
Orthostatic hypotension is a major autonomic sign in α-synucleinopathies.
Patients with orthostatic hypotension include: (a) those with pure autonomic
failure, characterized by isolated peripheral autonomic dysfunction and
decreased NE synthesis; (b) multiple system atrophy with symptoms of a central
Parkinson-like syndrome and normal resting plasma NE; (c) Parkinson’s
Some Clinical Autonomic Entities 297

disease, with lesions in postganglionic noradrenergic neurons and signs of auto-


nomic dysfunction [19].
Pure autonomic failure is a sporadic, adult-onset, slowly progressive neurode-
generative disorder. It is characterized pathologically by the abnormal accumula-
tion of α-synuclein in peripheral autonomic neurons and clinically by symptomatic
orthostatic hypotension, variable bladder dysfunction, and sexual dysfunction,
with no somatic neurological deficit. Pure autonomic failure patients usually have
very low plasma NE levels when recumbent, whereas plasma NE levels when
recumbent are normal in multiple system atrophy and variable in Parkinson’s dis-
ease. Pure autonomic failure selectively involves efferent autonomic neurons, with
the postganglionic neurons mainly affected. Afferent pathways are spared. It
occurs sporadically, and progresses slowly with a relatively good prognosis.
However, some cases of pure autonomic failure may develop a central neurodegen-
erative disorder [19].
Multiple system atrophy is a sporadic and fatal α-synuclein-linked oligodendro-
gliopathy manifesting with progressive autonomic failure, poorly l-Dopa-­responsive
parkinsonism, and cerebellar ataxia, in any combination [20]. This combined par-
kinsonian and autonomic disorder is also referred to as the Shy–Drager syndrome.
In multiple system atrophy, involvement of the rostral ventrolateral medulla is pri-
marily responsible for orthostatic hypotension and involvement in the pontine mic-
turition area; the sacral preganglionic nucleus is responsible for neurogenic bladder,
and involvement of the pre-Bötzinger complex and medullary raphe contributes to
sleep-related respiratory abnormalities (Chap. 4). In contrast, early involvement of
the enteric nervous system and cardiac sympathetic ganglia characterize Lewy body
disorders. The dorsal motor nucleus of the vagus is affected both in multiple system
atrophy and in the early stages of Parkinson’s disease [18].
Parkinson’s disease is a neurodegenerative disorder defined by its motor fea-
tures: asymmetric resting tremor, rigidity, bradykinesia, and postural instability.
Autonomic symptoms are present to varying degrees in patients with Parkinson’s
disease, but very severe dysautonomia in patients with mild motor findings suggests
a diagnosis of multiple system atrophy.
There are several differing criteria in use to diagnose Parkinson’s disease and
parkinsonian syndromes. In Parkinson’s disease, symptoms and signs of autonomic
failure occur commonly, especially in cardiovascular, gastrointestinal, and genito-
urinary domains. Most patients with Parkinson’s disease have neuroimaging evi-
dence for cardiac sympathetic denervation. In Parkinson’s disease, orthostatic
hypotension can be an early finding and is associated with extracardiac noradrener-
gic denervation, reduced cardiovagal baroreflex, and abnormal sympathoneural
responses [21]. Histopathology of postmortem tissue is typically required for defini-
tive diagnosis. Commonly, separating Parkinson’s disease from parkinsonian syn-
dromes rests largely upon the response of the physical examination to l-Dopa
replacement therapy.
Dementia with Lewy bodies is a fatal α-synucleinopathy characterized by
severely affected cognition, visual hallucinations, and parkinsonism. It represents
the second most common cause of neurodegenerative dementia in the elderly after
298 7  Clinical Implications of the Enlarged Autonomic Nervous System

AD. Progressive cognitive decline with deficits of visuospatial ability and frontal


executive function is accompanied by only mildly to moderately severe parkinson-
ism, which is often bilateral akinetic-rigid without the classic rest tremor.
Management of patients with dementia with Lewy bodies is based on a multidi-
mensional approach considering the cognitive decline and dementia that form the
core clinical syndrome. Patients with dementia with Lewy bodies have a pro-
nounced cholinergic deficit. Orthostatic hypotension may be a disabling feature
that if present frequently exacerbates the disability arising from progressive motor
disturbance [22].
Autonomic nerve system dysfunction in the form of REM sleep behavior dis-
order (lack of motor atonia during REM sleep) is an early manifestation of disease
in α-synucleinopathies, commonly occurring in the prodromal period. This is
likely due to the proximity of the cholinergic REM sleep nuclei and the autonomic
nuclei in the brainstem. In the Braak model of neurodegeneration, these nuclei
become impaired before the motor nuclei are affected, as the deposition of
α-synuclein progresses in a rostral–caudal fashion from the lower brain stem to
the cortex. REM sleep behavior disorder is a common prodromal manifestation of
Parkinson’s disease occurring about 12–15 years in advance.

Diabetes Mellitus Autonomic Dysfunction

In diabetes mellitus patients, perturbations in autonomic function underlie most


pathophysiology, from abnormalities in pupillary function to gastroparesis, intesti-
nal dysmotility, diabetic diarrhea, genitourinary dysfunction, among others [23].
Sympathetic nerves, which dilate the iris, show earlier and more extensive impair-
ment than the parasympathetic nerves, which constrict the iris. Signs and symp-
toms of gastric emptying abnormalities in diabetes include early satiety, nausea,
vomiting, large fluctuations in blood glucose, and weight loss. The prevalence of
gastroparesis is 30–50% of patients with long-standing type I and type II diabetes
mellitus (T2DM). Diarrhea and the development of gallstones are symptoms of
gallbladder atony. Gallstones are much more likely to develop in patients with
hypercholesterolemia, and are often seen in patients with diabetes. The most com-
mon gastrointestinal symptom of autonomic neuropathy is constipation. Both
afferent and efferent nerves to the bladder can be affected in T2DM patients.
Afferent neuropathy results in the inability to feel the need to void. Therefore,
there is a decrease in the frequency of urination, which may be misinterpreted as
an improvement in glucose control.
Cardiac autonomic neuropathy in diabetes mellitus has been linked to resting
tachycardia, postural hypotension, orthostatic bradycardia and orthostatic tachycar-
dia, exercise intolerance, decreased hypoxia-induced respiratory drive, loss of baro-
receptor sensitivity, enhanced intraoperative or perioperative cardiovascular lability,
increased incidence of asymptomatic ischemia, myocardial infarction, and decreased
rate of survival after myocardial infarction and congestive heart failure [24].
Autonomic dysfunction can affect the daily activities of individuals with diabetes
Some Clinical Autonomic Entities 299

mellitus and may provoke potentially life-threatening outcomes. For example,


intensification of glycemic control in the presence of autonomic dysfunction (more
so if combined with peripheral neuropathy) increases the likelihood of sudden death
and is a caveat for aggressive glycemic control [23, 25].

Autoimmune Autonomic Ganglionopathy

Autoimmune autonomic ganglionopathy consists of a panautonomic failure


caused by antibodies to ganglionic acetylcholine (ACh) receptors. The clinical
syndrome is characterized by significant postural hypotension, diffuse choliner-
gic and adrenergic impairment, gastrointestinal dysmotility, urinary retention,
and pupillary dysfunction. Sicca complex (dryness of the mucous membranes, as
of the eyes and mouth, in the absence of a connective tissue disease) and hypohi-
drosis also occur. Serological testing for ganglionic ACh receptor antibodies
helps to confirm the diagnosis. These antibodies cause a similar phenotype of
autonomic failure in animal models, indicating that an antibody-mediated func-
tional impairment of ganglionic transmission is the underlying etiology [26]. The
usual course of autoimmune autonomic ganglionopathy is monophasic worsen-
ing followed by incomplete recovery. Some patients experience a chronic pro-
gressive course or stable dysautonomia without recovery. The diagnosis of
idiopathic autoimmune autonomic ganglionopathy is suspected in cases of
acquired autonomic failure without somatic neuropathy, when toxic or paraneo-
plastic causes have been excluded.

Paraneoplastic Autonomic Dysfunction

Patients with paraneoplastic autonomic neuropathy may be clinically indistinguish-


able from those with idiopathic autoimmune autonomic ganglionopathy until a can-
cer, usually small-cell carcinoma of the lung, is detected [27, 28]. Immune-mediated
paraneoplastic autonomic dysfunction comprises a spectrum of neurological dys-
function that may range from isolated autonomic involvement to dysautonomia with
recognizable syndromes including limbic encephalitis, subacute sensory neuronop-
athy, neuromyotonia, and Lambert–Eaton myasthenia syndrome. The neurological
dysfunction may be multifocal and widespread owing to evolving immune response
to multiple onconeural antigens in a single tumor. Somatic neurological findings are
of variable severity. Urgent evaluation is necessary to prevent progressive neuronal
loss, especially in CNS syndromes [27, 28].

Amyloidotic Autonomic Failure

Autonomic failure is an important feature of immunoglobulin amyloidosis and hered-


itary systemic amyloidosis. Mutant forms of transthyretin cause the most common
300 7  Clinical Implications of the Enlarged Autonomic Nervous System

type of autosomal-dominant hereditary systemic amyloidosis–familial amyloidotic


polyneuropathy [29]. It must be considered in all patients with familial or paraprotein-
emic neuropathies and autonomic dysfunction. The amyloid fibrils in immunoglobu-
lin amyloidosis amyloidopathy are composed of immunoglobulin light chain proteins
or their degradation products. Signs of autonomic failure include orthostatic hypoten-
sion with inappropriate heart rate response, impotence, dry mouth, gastrointestinal
autonomic disturbances, sluggish pupillary action, and impairment of sweating.
Autonomic dysfunction, rather than deposition of amyloid in the mucosa, seems to be
a more frequent cause of gastrointestinal symptoms in these patients.

Autonomic Dysfunction in Primary Sleep Disorders

The ANS is integrally related to sleep initiation, maintenance, and disruption. When
such disruptions become frequent or chronic, autonomic impairment may follow
dysfunction [30]. In the short term, such autonomic impairment may lead to
increased sympathetic drive and a sensation of hyperarousal, further perpetuating
the sleep disturbance. If sustained, this impairment may result in significant morbid-
ity and even mortality. The most prevalent sleep disorders are insomnia, sleep dis-
ordered breathing, and restless legs syndrome.

Insomnia
Insomnia is among the most prevalent health concerns in the general population and
it is present at particularly high rates among people with comorbid health problems.
Approximately one-third of adults have occasional difficulty with insomnia and
about 10–15% have persistent problems achieving sufficient sleep. Although the
word insomnia may be used colloquially to represent poor sleep, specific diagnostic
criteria for insomnia disorders are detailed in the Diagnostic and Statistical Manual,
5th Edition (DSM-5) [31] and the International Classification of Sleep Disorders,
3rd Edition (ICSD-3) [32]. Essentially, insomnia disorder is a persistent difficulty in
falling asleep or remaining asleep, or awakening earlier than desired in the context
of an adequate opportunity for sleep, and is associated with daytime consequences
(for example, fatigue, poor concentration).
Like patients with obstructive sleep apnea (OSA), many patients with insomnia
have high BP and are nondippers. i.e., they do not show the normal decrease in BP
at night. In addition, frequent arousals, either spontaneous or secondary to an under-
lying sleep disorder, can result in increased sympathetic tone. There is a typical
cardiac response observed during an arousal: an initial tachycardia, which often
precedes the electrocortical arousal by several seconds, followed by bradycardia. If
the arousals are frequent enough, the elevation in sympathetic tone can persist long
after the patient has returned to sleep (Chap. 4). This association of high BP and
insomnia remained after adjusting for age, race, gender, smoking, obesity, diabetes,
alcohol consumption, depression, and other sleep disorders such as OSA and peri-
odic limb movement disorder [30].
Some Clinical Autonomic Entities 301

Brain imaging studies support the increased nocturnal cardiac sympathetic drive
in insomnia patients. Increased activation and hypermetabolism in the arousal net-
works of the hypothalamus and brainstem, in addition to their efferent projections
in the medial prefrontal cortex and amygdala, have been demonstrated in insomniac
patients by PET. In addition, electroencephalography studies indicate a beta (14–
35 Hz) and gamma (35–45 Hz) activity, frequencies typically associated with the
cortical activity of the waking state. Sleep-deprived patients without insomnia (e.g.,
shift workers) also exhibit signs of autonomic dysfunction and are at a greater risk
of developing cardiovascular disorders, even if young and otherwise healthy [33].

 leep-Related Breathing Disorders


S
This defines a wide spectrum of abnormalities of respiration during sleep, including
abnormal respiratory pattern (e.g., apneas, hypopneas, or respiratory effort-related
arousals) or abnormal reduction in gas exchange (e.g., hypoventilation) during
sleep. The ICSD-3 has defined four major categories of sleep-related breathing dis-
orders: (1) OSA disorders, (2) central sleep apnea disorders, (3) sleep-related
hypoventilation disorders, and (4) sleep-related hypoxemia disorder [32]. For this
summary, we only consider OSA. which is the most common sleep-related breath-
ing disorder, affecting an estimated 5–10% of the general population.
In OSA patients, repeated apneas or hypopneas can have an impact on the ANS
and lead to significant consequences. When a patient with OSA experiences an
obstructive respiratory event during sleep, pulmonary autonomic afferents are
largely inhibited because of the prolonged increase in negative intrathoracic pres-
sure, the result of inspiring against a closed or partially closed glottis (Chap. 4).
Thus, hyperventilation is prevented, baroreceptors are stimulated, and sympathetic
vasomotor tone increases, leading to peripheral vasoconstriction. Another frequent
concomitant phenomenon is hypoxemia, which activates chemoreceptors in the
carotid bodies, and exacerbates sympathetic vasomotor tone increases. Moreover,
the diving reflex is triggered by hypoxemia, with an increase in cardiac vagal tone
helping to preserve blood flow to the heart and brain while limiting cardiac oxygen
demand [30].
If a sleep-related hypoventilation syndrome occurs, the augmented PCO2 stimu-
lates central chemoreceptors and further increases sympathetic tone via a similar
process. BP increases because the venous return increases during apnea and periph-
eral vasoconstriction persists for some time after the patient initiates a recovery
breath and resumes normal breathing, resulting in large BP surges. In susceptible
individuals tachyarrhythmias may occur.
Concerning BP, patients with OSA tend to be “nondipping” or can also exhibit
“reverse dipping,” whereby BP increases during sleep, indicative of increased sym-
pathetic tone. Moreover, this increase, together with a diminished baroreceptor
reflex sensitivity are also found in wakefulness and may contribute to the increased
incidence of cardiovascular events in OSA patients. There is evidence that OSA is
an independent risk factor for arterial hypertension, cardiovascular disease, and
ischemic stroke. Most cardiovascular and cerebrovascular events occur in the early
302 7  Clinical Implications of the Enlarged Autonomic Nervous System

hours of the morning, either out of sleep or shortly after awakening. This may be
related to the increased sympathetic tone at this time, as the frequency and duration
of phasic REM periods increase [30].

 eriodic Limb Movement Disorder and Restless Legs Syndrome


P
Periodic limb movements are mainly characterized by relatively simple, usually
stereotyped, movements that disturb sleep or its onset. Periodic leg movements in
sleep are a frequent finding in PSG; its prevalence is estimated to be 4–11% in
adults. Periodic limb movement disorder is defined by the PSG demonstration of
periodic limb movements of >5/h in children and >15/h in adults that cause signifi-
cant sleep disturbance or impairment of functioning.
Some patients with otherwise unexplained insomnia, fatigue, or hypersomnia
have PSG results that reveal an elevated number of periodic leg movements. In sleep
studies, periodic leg movements are most frequently found in restless legs syn-
drome. They also often occur in narcolepsy, sleep apnea syndrome, and REM sleep
behavior disorder.
The autonomic arousal response discussed earlier, that of a rapid rise in heart rate
and arterial BP followed by rapid bradycardia and a return of BP to baseline values,
has been demonstrated before the onset of periodic limb movements during sleep
[30]. In fact, even without an arousal, periodic limb movements have been associ-
ated with autonomic cardiovascular response, although the magnitude of the
response is greater when an arousal is present. It seems feasible that periodic limb
movements result from the loss of subcortical inhibition to pacemaker cells in the
spinal cord or brainstem that have phasic control of autonomic, motor, and arousal
networks. Like patients with OSA and insomnia, patients with periodic limb move-
ment disorder are at an increased risk of cardiovascular disease.
Restless legs syndrome is a common disorder affecting an estimated 5–10% of
the population. Unlike periodic limb movement disorder, restless legs syndrome is
a clinical syndrome that includes an urge to move the legs, symptoms that worsen
with rest or inactivity, are relieved with movement, and occur predominantly in the
evening hours. The most commonly ANS symptoms were sialorrhea, constipation,
early satiety, heat intolerance, and orthostatic intolerance. Periodic limb movements
are present in approximately 80–90% of patients with restless legs syndrome; how-
ever, not all patients with periodic limb movements have symptoms of restless legs
syndrome. Although the pathophysiology of restless legs syndrome is yet to be
elucidated, and is likely multifactorial, one theory involves a reduction in dopami-
nergic outflow to the preganglionic sympathetic neurons in the dorsal horn of the
spinal cord. DA inhibits preganglionic sympathetic neurons; therefore, a reduction
in DA may in turn increase sympathetic outflow [30].

 EM Sleep Behavior Disorder


R
Rapid eye movement sleep behavior disorder is characterized by partial arousal and
abnormal behaviors emerging during REM sleep that may cause injury. It is a condi-
tion whereby a patient loses the protective muscle atonia that normally occurs dur-
ing REM sleep. These patients are thereby free to act out their dreams, many times
Some Clinical Autonomic Entities 303

with injurious consequences. REM behavior disorder is more common in individu-


als with neurodegenerative disorders: 50–80% of patients presenting idiopathic
REM behavior disorder go on to develop a synucleinopathy over 10–15 years.
Clinical manifestations in idiopathic REM behavior disorder are identical to those
seen in cases secondary to Parkinson’s disease, multiple system atrophy and Lewy
body disease and has been associated with mitochondrial disorders, brain tumors,
and many other diseases, and clinical conditions that may damage brainstem mech-
anisms involved in generating REM sleep atonia. REM behavior disorder is proba-
bly the strongest nonmotor predictor of future disease, with an estimated median
latency of motor symptoms of 12–14 years.
Dysfunction of the ANS is an early manifestation of disease in these patients,
commonly occurring in the prodromal period [30]. REM sleep behavior disorder
patients suffer from common symptoms of autonomic impairment, with the greatest
deficits in gastrointestinal, urinary, and cardiovascular function. Patients have
decreased HRV during REM sleep and the Valsalva ratio, an indicator of cardiova-
gal function, was significantly lower compared with healthy controls. A postgangli-
onic cardiac sympathetic denervation, as measured by 123I-metaiodobenzylguanidine
scintigraphy, further support the concept of prodromal autonomic impairment in
patients with REM sleep behavior disorder [30].

Autonomic Dysfunction in Cardiovascular Disorders

The sympathetic nervous system participates in the development and progression of


the essential hypertensive state, as shown by increased circulating plasma levels of
NE, elevated NE spillover rate, and augmented sympathetic nerve traffic discharge
detected in the high BP state. In addition, the sympathetic overdrive participates in
the development of the metabolic disarray and in target organ damage [34]. The
above-mentioned sympathetic abnormalities explain why adrenergic overdrive rep-
resents an important therapeutic target in the treatment of hypertension.
Hypertension is characterized by autonomic abnormalities that include sympa-
thetic activation and parasympathetic impairment [34]. These abnormalities occur
in the earlier phases of the disease, participating in high BP development and pro-
gression. In addition, these autonomic abnormalities participate in the pathogene-
sis of hypertension-related end-organ damage and in the development of cardiac
hypertrophy, of large, medium, and small artery structural changes, and of renal
dysfunction. The autonomic abnormalities characterizing high BP are potentiated
in the presence of metabolic alterations, such as insulin resistance, obesity, and
metabolic syndrome.
Time and frequency domain estimates of tonic and reflex vagal HRV have dem-
onstrated consistently the loss of parasympathetic tone at the onset of heart failure
[35]. As heart failure advances, the loss of sino-atrial responsiveness to neutrally
released and circulating catecholamines blunt further heart rate variation, which
becomes a marker of reduced life expectancy. With regard to the ventricles, vagus
nerve activity is generally antiarrhythmic, as it inhibits the profibrillatory effects of
304 7  Clinical Implications of the Enlarged Autonomic Nervous System

sympathetic nerve activation, whereas atrial arrhythmias generally derive from


heightened levels of both vagus and sympathetic nerve activity [36].
Because heart failure was conceptualized initially as a primarily hemodynamic
disorder, attention was directed first at potential baroreceptor reflex-mediated con-
tributions to sympathetic activation and vagal withdrawal [37]. Augmented periph-
eral chemosensitivity to hypoxia, present in about half of patients with advanced
congestive failure, is associated with higher plasma NE concentrations, loss of tonic
and reflex heart rate modulation, oscillatory patterns in breathing, and ventricular
arrhythmias. Each of these abnormalities is exacerbated by exercise and anticipates
premature death. The central sympathoexcitation induced by sleep apnea may be
both causal and contributory to heart failure and its progression [30].
The absence of proper regulation of mean arterial pressure can have significant
pathophysiological consequences. Low mean arterial pressure can cause inadequate
blood flow to organs, syncope, and shock. On the other hand, elevated mean arterial
pressure contributes to increased oxygen demand by the heart, ventricular remodel-
ing, vascular injury, end organ damage, and stroke.
The arterial baroreflex system is a key controller of mean arterial pressure and is
a complex system (Fig. 4.12). It can be considered in its entirety as an integrative
physiological system or in terms of its regulated component parts. Baroreflex sensi-
tivity for the control of heart rate is consistently decreased in numerous pathological
states including chronic hypertension, coronary artery disease, postmyocardial
infarction, heart failure, diabetes mellitus, obesity, and aging [38]. The prominent
feature of baroreflex failure is volatile arterial hypertension. Hypertensive episodes
can be explained by sympathetic activation that is unopposed by the baroreflex.
Episodes can be triggered by factors such as psychological stress, physical exercise,
and pain [39]. A few patients with baroreflex failure present with episodes of hypo-
tension and bradycardia that can be observed when patients are resting and cortical
input is diminished.
Postural tachycardia syndrome is a syndrome of orthostatic tachycardia associ-
ated with symptoms of cerebral hypoperfusion and/or autonomic activation [40].
Patients are usually women aged 20–50 years and some have limited autonomic
neuropathy. Evidence of peripheral denervation of sudomotor fibers includes sweat
loss from the legs on the thermoregulatory sweat test or the quantitative sudomotor
axon reflex test (Table 7.2). Peripheral adrenergic denervation may be present,
resulting in impairment of reflex vasoconstriction with baroreflex unloading.
In hyperadrenergic postural tachycardia syndrome, sympathetic tone is increased,
and manifests as orthostatic hyperadrenergic response and sometimes as a spontane-
ous episode of excessive sympathetic activity. Patients are characterized by supine
vasoconstriction, increased peripheral venous pressure, reduced stroke volume and
cardiac output, blunted orthostatic vasoconstrictive responses, often supine tachycar-
dia, and reduced blood volume of uncertain origin. There is increased supine cardiac
output compared with healthy volunteers. Relative lower extremity vasodilation per-
sists during orthostatic stress, causing venous pooling in the legs [40].
The postural tachycardia syndrome is not a single disease. It is best viewed as a
“disorder” or a syndrome in which excessive orthostatic tachycardia can be a final
Some Clinical Autonomic Entities 305

common pathway of many underlying pathophysiological processes. These include


hyperadrenergic postural tachycardia syndrome, NE transporter deficiency, mast
cell activation disorder, and neuropathic postural tachycardia syndrome. These
pathophysiological mechanisms are not mutually exclusive, but can coexist in an
individual patient with postural tachycardia syndrome [41]. Some postural tachy-
cardia syndrome patients have a form of dysautonomia, with preferential denerva-
tion of sympathetic nerves innervating the lower limbs [40].
Patients with delayed orthostatic hypotension show a fall in systolic BP of
≥20 mmHg or diastolic BP of ≥10 mmHg lasting longer than 3 min of standing or
upright tilt table testing [19]. Delayed orthostatic hypotension can be caused by a
variety of pathogenetic mechanisms, including extra-adrenal pheochromocytoma,
hyperadrenergic orthostatic hypotension, hyperbradykininism, primary hyperepi-
nephrinemia, and hypoaldosteronism with baroreflex impairment. There were
greater abnormalities of both phase II and IV of the Valsalva maneuver in patients
with earlier and greater decreases in BP compared with those with later and milder
reductions in BP. Abnormalities were also more likely detected in HRV in patients
with earlier and more severe decreases in BP. Symptoms of orthostatic intolerance
in patients with delayed orthostatic hypotension are similar to those with orthostatic
hypotension that occurs within the first 3 min of upright posture [19].
In pathophysiological states, such as hypertension, heart failure, and chronic
renal disease, there may be an inappropriate sympathoexcitation causing sodium
retention, which exacerbates the disease process. The contribution of the renal sym-
pathetic nerves to these cardiovascular diseases is demonstrated by the long-term
normalization of BP found in resistant hypertensive patients subjected to renal
denervation (Chap. 5) [42].
Pregnancy increases sympathetic nerve firing and decreases both basal parasym-
pathetic activity and baroreflex function [43]; these changes are exaggerated in
women with preeclampsia. Preeclampsia is a potentially fatal hypertensive disorder
of pregnancy that is initiated by reduced placental perfusion. Increased sympathetic
tone contributes to the hypertension, as basal muscle sympathetic nerve activity is
clearly increased above the levels observed in normal pregnant women. The changes
in basal autonomic tone may counteract to some degree the profound vasodilation
that is a hallmark of normal pregnancy [44].

Autonomic Dysfunction Associated with Aging

Healthy human aging is associated with several abnormalities in ANS function that
can impair an older person’s adaptation to the stresses of everyday life [45].
Medications that acutely lower BP may also contribute to hypotension in elderly
patients, particularly benzodiazepines, diuretics, antihypertensives, α-blockers used
for prostatic obstruction, l-Dopa, tricyclic antidepressants, and neuroleptics.
Orthostatic hypotension is an important symptom of autonomic failure, commonly
associated with diabetes, malignancy, amyloidosis, Parkinson’s disease, multiple
system atrophy, Lewy body dementia, pure autonomic failure, and other syndromes
306 7  Clinical Implications of the Enlarged Autonomic Nervous System

in elderly patients [45]. The increase in plasma NE is primarily due to an increase


in NE spillover at sympathetic nerve endings and secondarily due to a decrease in
clearance. When plotted continuously over time, the beat-to-beat heart rate or BP
signal is highly irregular because of the interactions of multiple autonomic control
systems operating over different time scales.

Guillain–Barré Syndrome

Guillain–Barré syndrome is an acute immune-mediated demyelinating neuropathy


that may cause profound weakness and respiratory failure over a period of a few
weeks. Although the disease affects motor fibers preferentially, paresthesias and
pain are common sensory manifestations. Autonomic neuropathy of some degree,
particularly involving cardiovascular and gastrointestinal function, is found in two-­
thirds of patients and may be a life-threatening complication of the disease [46].
Autonomic dysfunction in Guillain–Barré syndrome can affect the sympathetic and
the parasympathetic nervous system. Autonomic manifestations comprise a combi-
nation of autonomic failure and autonomic over-reactivity, the latter most com-
monly being manifested as sinus tachycardia and systemic hypertension. Marked
arrhythmias and wide fluctuations in BP may also occur. Gastrointestinal dysmotil-
ity is common, but rarely progresses to adynamic ileus. Typically, autonomic neu-
ropathy improves in concert with motor and sensory nerve function.

Hereditary Autonomic Neuropathies

Inherited neuropathies with autonomic involvement include Fabry disease (angio-


keratoma corporis diffusum), an X-linked, inherited, slowly progressive metabolic
disorder with nonspecific clinical manifestations [47]. Unlike many other lysosomal
storage diseases, most patients remain clinically silent during the first few years of
life. Cerebrovascular disease secondary to multifocal abnormalities of large and
small vessels, including transient ischemic attacks and stroke, may also occur.
Acute hepatic porphyrias (acute intermittent porphyria, variegate porphyria,
and hereditary coproporphyria) are a group of autosomal dominant, inherited met-
abolic disorders that manifest as acute or subacute, severe, life-threatening motor
neuropathy, abdominal pain, autonomic dysfunction, and neuropsychiatric symp-
toms [48]. The neurological manifestations of all forms of acute porphyrias are
identical. Symptoms of an acute attack include severe abdominal pain, nausea,
vomiting, constipation, diarrhea, urinary frequency and hesitancy, urine discolor-
ation, labile hypertension, tachycardia, excessive sweating, pain in the limbs and
back, and convulsions.
Familial dysautonomia (Riley–Day syndrome) or hereditary sensory and auto-
nomic neuropathy type III, is an autosomal recessive disease caused by mutations in
the gene that encodes for I-κ-B kinase complex-associated protein [49]. Affected
patients have a complex neurological phenotype. First, because of a congenital
Some Clinical Autonomic Entities 307

abnormality in the afferent baroreflex pathways, BP is extremely labile, with severe


episodic hypertension and orthostatic hypotension. In addition to the autonomic
cardiovascular abnormalities, patients with familial dysautonomia also have
decreased pain and temperature perception, an impaired sense of taste, abnormal
swallowing, gait ataxia, decreased/absent myotatic reflexes and decreased ventila-
tory responses to hypoxia and hypercapnia [49].

Autonomic Disturbances in Spinal Cord Injuries

In cervical and high thoracic transection, the entire or a large part of the sympathetic
outflow, together with the sacral parasympathetic outflow, is separated from cerebral
control. Autonomic malfunction may affect the cardiovascular, thermoregulatory,
sudomotor, gastrointestinal, urinary, and reproductive systems [5]. Basal heart rate is
usually below normal. In patients with high cervical lesions, who need artificial ven-
tilation because of diaphragmatic paralysis, severe bradycardia and cardiac arrest
may occur during tracheal stimulation. Plasma NE levels are low and do not rise with
head-up postural change, unlike normal subjects. There is a marked rise in levels of
plasma renin, aldosterone, and AVP, which may contribute to the recovery of BP and
account for other symptoms, such as reduced urine output. The reverse, severe hyper-
tension, may occur during autonomic dysreflexia following stimulation below the
level of the lesion, predominantly, but not always, through noxious stimuli [5].

Drug-Induced Autonomic Dysfunction

Medications frequently alter BP regulation, which may interfere with daily activities
or contribute to cardiovascular complications. Hypotension is frequently accompa-
nied by symptoms, but hypertension may go unrecognized until serious complica-
tions develop, prompting medical attention. Medications taken for anti-­inflammatory,
decongestant, or anorexic effects are frequently associated with an increase in
BP. Although patients may take these medications to promote a healthy lifestyle, the
frequent use of over-the-counter, off-label, and herbal medications may contribute to
drug-induced dysregulation of BP control. A thorough medication history is needed,
or an opportunity to prevent complications could be missed [3].

 utonomic Disorder in Fibromyalgia, Chronic Fatigue Syndrome


A
and Chronic Pain

Fibromyalgia is a disabling disease affecting 2–5% of the population, mostly young


and middle-aged women (7 to 1 vs men). Its major clinical signs are: generalized
musculoskeletal pain and allodynia, poor sleep quality, significant levels of fatigue,
and cognitive alterations, particularly problems with concentration and memory. The
etiology of fibromyalgia remains unknown; however, there are often characteristic
308 7  Clinical Implications of the Enlarged Autonomic Nervous System

alterations in sleep patterns and changes in neuroendocrine transmitters that sug-


gest that the pathophysiology of the syndrome might be associated with autonomic
and neuroendocrine regulation. Central sensitization, dampening of inhibitory pain
pathways, and changes in neurotransmitters can lead to abnormal processing of sen-
sory signals in the CNS, as such lowering the pain threshold and amplifying sensory
signals causing constant pain (allodynia) (Chap. 4) [50].
Frequent comorbidity of fibromyalgia and mood disorders suggests that stress
response and neuroendocrine abnormalities might play roles in the disease process.
In fibromyalgia, the stress–adaptation response is disrupted, leading to stress-­induced
symptoms. Patients with fibromyalgia also often suffer from comorbid anxiety disor-
ders. Neuroimmunoendocrine alterations in fibromyalgia include: (a) decrease in
cortisol secretion, which has been associated with the presence of chronic or intense
physical or psychological stress; (b) decrease in GH secretion; (c) chronic activation
of various cytokines by stress (higher levels of IL-8 and IL-10, which has been linked
to sleep problems); (d) decreased melatonin levels. fMRI studies support the hypoth-
esis of increased central pain, showing that patients with fibromyalgia have a
decreased volume of gray matter in the prefrontal cortex, amygdala, and anterior
cingulate cortex. Duration of pain or functional disability due to pain do not correlate
with gray matter volumes. One possible finding is that reductions in gray matter
volume may be a precondition for central sensitization in fibromyalgia [51].
Chronic fatigue syndrome is a disabling disorder characterized by persistent or
relapsing unexplained fatigue, accompanied by characteristic physical, constitu-
tional, and neuropsychological symptoms lasting at least 6 months [52]. The fatigue
is not relieved by rest, nor explained by medical or psychiatric conditions, and it is
accompanied by a range of cognitive and somatic symptoms. The pathophysiology
of chronic fatigue syndrome remains unclear. A hyperserotonergic state and hypo-
activity of the hypothalamic–pituitary–adrenal axis have also been indicated, but it
remains uncertain whether these are a cause of or a consequence of chronic fatigue
syndrome. Female gender genetic disposition, certain personality traits and physical
and emotional stressors have been identified as risk factors. Moreover, exposure to
childhood trauma has been found to increase the risk of chronic fatigue syndrome
three- to eightfold. Prevalence of chronic fatigue syndrome in primary care setting
ranges from 3 to 20%.
Many chronic fatigue patients experience clinical features consistent with auto-
nomic dysfunction, such as orthostatic symptoms and tachycardia, increased sweat-
ing, pallor, sluggish pupillary responses, gastrointestinal symptoms, and frequency
micturition. A particularly strong association has been observed between chronic
fatigue syndrome and orthostatic intolerance, including neurally mediated hypoten-
sion and postural tachycardia syndrome.
Both fibromyalgia and chronic fatigue syndrome are chronic clinical conditions
characterized by intractable fatigue and nonrestorative sleep [52]. In fibromyalgia,
the prominent diffuse musculoskeletal discomfort is the distinguishing symptom.
However, in general clinical terms the two conditions cannot be clearly distinguished.
Fibromyalgia and chronic fatigue syndrome, together with irritable bowel syn-
drome and migraine are clinical entities showing an abnormal central processing of
Some Clinical Autonomic Entities 309

pain [50]. Usually, nonpainful stimuli are amplified and experienced as painful (allo-
dynia) and sleep abnormalities are present because patients are unable to achieve
deep NREM sleep. Participation of the ANS in the pathogenesis of chronic pain is
indicated by the decreased high-frequency HRV, i.e., the decreased parasympathetic
activity, found in chronic pain. ANS dysfunction depends on pain conditions. For
example, fibromyalgia, unlike irritable bowel syndrome, is characterized by greatly
decreased parasympathetic activity. Fibromyalgia dysautonomia is also character-
ized by basal hyperactive sympathetic activity with hyporeactivity to stress. This
hyporeactivity to stress is less marked in localized chronic muscle pain [53].

Hyperthermia, Hypothermia

Hyperthermia is a result of a failure of the regulatory mechanisms to prevent the body


temperature from rising above euthermia values, because of high environmental or met-
abolic heat (Chap. 5). The hyperthermia syndrome consists of several disorders, ranging
from mild to fatal and temporary. The severity of hyperthermia depends on a variety of
endogenous factors that may affect the magnitude and rate of increase in body tempera-
ture to interact with the thermoregulatory system or other control systems.
Several symptoms may accompany an excessive heat load, such as swelling in
the lower limbs, orthostatic dizziness or syncope, exhaustion with headache, vomit-
ing and diarrhea, and, in more severe cases, heat stroke. Heat stroke is the break-
down of thermoregulatory mechanisms resulting in a sudden increase in body
temperature to values above 39–40 °C. Among the various factors leading to heat
stroke, two classes can be distinguished: physiological unsuitability and malfunc-
tion of the effector organs. During exposure to a moderately hot environment, blood
flow of the body surface increased without compromising the blood supply to other
tissues, a fact that is due to a rise in and/or redistribution of cardiac output [54].
Flow redistribution occurs by reducing the blood supply to tissues considered non-
vital organs to increase that to organs of the splanchnic area, kidney, muscle, and
adipose tissue. However, when the heat stress becomes severe, the thermoregulatory
mechanisms may conflict with the mechanisms regulating blood pressure, as com-
peting demands arise. It is believed that such a conflict in homeostatic systems is the
main causative factor in heat stroke.
Hypothermia usually occurs when the body temperature drops below 35°C and
can be mild (34–35°C), moderate (30–34°C) or severe (below 30°C). There is a
failure of the thermoregulatory mechanisms to maintain euthermia, but they cannot
prevent the decrease in body temperature either. Hypothermia can result from expo-
sure to very cold air, immersion in cold water, metabolic disorders (hypothyroidism,
hypoglycemia, malnutrition) or disorders of the central nervous system (hypothala-
mus), drug effects, alcohol abuse, or even attenuated behavioral thermoregulation
[55]. Hypothermia produces progressive depression of mental functions, apathy
starting with and ending with lethargy, and coma when the body temperature reaches
28–30°C. Except for the adrenergic activity in peripheral vessels (to induce vaso-
constriction) and nerve activity involved in the tremor, hypothermia leads to a
310 7  Clinical Implications of the Enlarged Autonomic Nervous System

gradual reduction in the conduction velocity of nerve impulses. In severe hypother-


mia, carbon dioxide retention induces an acidotic state. It is important to emphasize
that the reduction in temperature has several benefits in conditions of oxygen short-
age. In the clinic, forced hypothermia is being increasingly used as a therapeutic
strategy for some surgical procedures or brain ischemia situations.

References
1. Biaggioni I, Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous sys-
tem. 3rd ed. San Diego: Academic; 2012.
2. Buijs RM, Swaab D, editors. Handbook of clinical neurology. Autonomic nervous system, vol.
117. New York: Elsevier; 2013.
3. Low PA, Engstrom JW. Disorders of the autonomic nervous system. In: Kasper D, Fauci A,
Hauser S, Longo D, Jameson JL, Loscalzo J, editors. Harrison’s principles of internal medi-
cine. 19th ed. New York: McGraw-Hill Education; 2015.
4. Sandroni P. Chapter 78 – Clinical evaluation of autonomic disorders. In: Biaggioni I, Burnstock
G, Low PA, Paton JFR, editors. Primer on the autonomic nervous system. Third ed. San Diego:
Academic; 2012. p. 377–82.
5. Ropper AH, Samuels MA, Klein JP. Chapter 26. Disorders of the autonomic nervous sys-
tem, respiration, and swallowing. In: Adams and Victor’s principles of neurology. 10th ed.
New York: The McGraw-Hill Companies; 2014.
6. Raj SR. Chapter 79 – Tilt table studies. In: Biaggioni I, Burnstock G, Low PA, Paton JFR,
editors. Primer on the autonomic nervous system. Third ed. San Diego: Academic; 2012.
p. 383–7.
7. Vigo DE, Nicola-Siri L, Ladrón de Guevara MS, Martinez-Martinez JA, Fahrer RD, Cardinali
DP, Massoli O, Guinjoan SM. Relation of depression to heart rate nonlinear dynamics in
patients >60 years of age with recent unstable angina pectoris or acute myocardial infarction.
Am J Cardiol. 2004;93:756–60.
8. Vigo DE, Guinjoan SM, Scaramal M, Nicola-Siri L, Cardinali DP. Wavelet transform shows
age-related changes of heart rate variability within independent frequency. Auton Neurosci.
2005;123:94–100.
9. Vigo DE, Domínguez J, Scaramal M, Ruffa E, Solernó J, Nicola-Siri L, Cardinali DP. Nonlinear
analysis of heart rate variability within independent frequency components during thesleep-­
wake cycle. Auton Neurosci. 2010;154:84–8.
10. Vigo DE, Castro MN, Dörpinghaus A, Weidema H, Cardinali DP, Nicola-Siri L, Rovira B,
Fahrer RD, Nogués M, Leiguarda RC, Guinjoan SM. Nonlinear analysis of heart rate vari-
ability in patients with eating disorders. World J Biol Psychiatry. 2007;20089(03):183–9.
11. Liu JC, Verhulst S, Massar SA, Chee MW. Sleep deprived and sweating it out: the effects
of total sleep deprivation on skin conductance reactivity to psychosocial stress. Sleep.
2015;38:155–9.
12. Guinjoan SM, Bonanni Rey RA, Cardinali DP. Correlation between skin potential response and
psychopathology in patients with affective disorders. Neuropsychobiology. 1995;31:24–30.
13. Wallin BG. Chapter 80 – Sympathetic microneurography. In: Biaggioni I, Burnstock G,
Low PA, Paton JFR, editors. Primer on the autonomic nervous system. Third ed. San Diego:
Academic; 2012. p. 389–92.
14. Goldstein DS. Chapter 29 – Sympathetic neuroimaging. In: Buijs RM, Swaab D, editors.
Handbook of clinical neurology. Autonomic nervous system. New York: Elsevier; 2013.
p. 365–70.
15. Wang N, Gibbons CH. Chapter 30 – Skin biopsies in the assessment of the autonomic nervous
system. In: Buijs RM, Swaab D, editors. Handbook of clinical neurology. Autonomic nervous
system. New York: Elsevier; 2013. p. 371–8.
References 311

16. Kuritzky L, Espay AJ, Gelblum J, Payne R, Dietrich E. Diagnosing and treating neurogenic
orthostatic hypotension in primary care. Postgrad Med. 2015;127:702–15.
17. Chiang MC, Tseng MT, Pan CL, Chao CC, Hsieh ST. Progress in the treatment of small fiber
peripheral neuropathy. Expert Rev Neurother. 2015;15:305–13.
18. Cersosimo MG, Benarroch EE. Chapter 5 – Central control of autonomic function and involve-
ment in neurodegenerative disorders. In: Buijs RM, Swaab D, editors. Handbook of clinical
neurology. Autonomic nervous system. New York: Elsevier; 2013. p. 45–57.
19. Garland EM, Hooper WB, Robertson D. Chapter 20 – Pure autonomic failure. In: Buijs RM,
Swaab D, editors. Handbook of clinical neurology. Autonomic nervous system. New York:
Elsevier; 2013. p. 243–57.
20. Wenning GK, Krismer F. Chapter 19 – Multiple system atrophy. In: Buijs RM, Swaab D, edi-
tors. Handbook of clinical neurology. Autonomic nervous system. New York: Elsevier; 2013.
p. 229–41.
21. Kaufmann H, Goldstein DS. Chapter 21 – Autonomic dysfunction in Parkinson disease. In:
Buijs RM, Swaab D, editors. Handbook of clinical neurology. Autonomic nervous system.
New York: Elsevier; 2013. p. 259–78.
22. Ropper AH, Samuels MA, Klein JP. Chapter 21. Dementia, the amnesic syndrome, and the
neurology of intelligence and memory. In: Adams and Victor’s principles of neurology. 10th
ed. New York: The McGraw-Hill Companies; 2014.
23. Vinik AI, Erbas T. Chapter 22 – Diabetic autonomic neuropathy. In: Buijs RM, Swaab D, edi-
tors. Handbook of clinical neurology. Autonomic nervous system. New York: Elsevier; 2013.
p. 279–94.
24. Rodriguez-Colon SM, Li X, Shaffer ML, He F, Bixler EO, Vgontzas AN, Cai J, Liao D. Insulin
resistance and circadian rhythm of cardiac autonomic modulation. Cardiovasc Diabetol.
2010;9:85.
25. Cryer PE. Chapter 23 – Hypoglycemia-associated autonomic failure in diabetes. In: Buijs RM,
Swaab D, editors. Handbook of clinical neurology. Autonomic nervous system. New York:
Elsevier; 2013. p. 295–307.
26. Muppidi S, Vernino S. Chapter 25 – Autoimmune autonomic failure. In: Buijs RM, Swaab
D, editors. Handbook of clinical neurology. Autonomic nervous system. New York: Elsevier;
2013. p. 321–7.
27. Chefdeville A, Honnorat J, Hampe CS, Desestret V. Neuronal central nervous system syn-
dromes probably mediated by autoantibodies. Eur J Neurosci. 2016;43:1535–52.
28. Paraschiv B, Diaconu CC, Toma CL, Bogdan MA. Paraneoplastic syndromes: the way to an
early diagnosis of lung cancer. Pneumologia. 2015;64:14–9.
29. Ueda M, Ando Y. Recent advances in transthyretin amyloidosis therapy. Transl Neurodegener.
2014;3:19.
30. Miglis MG. Autonomic dysfunction in primary sleep disorders. Sleep Med. 2016;19:40–9.
31. American Psychiatric Association. Diagnostic and statistical manual of mental disorders. 5th
ed. 2015.
32. American Academy of Sleep Medicine. International classification of sleep disorders. 3rd ed.
Darien: American Academy of Sleep Medicine; 2014.
33. Lavigne G. Chapter 17 – Relevance of sleep physiology for sleep medicine clinicians
A2 – Kryger, Meir H. In: Roth T, Dement WC, editors. Principles and practice of sleep medi-
cine. 5th ed. Philadelphia: W.B. Saunders; 2011. p. 199–200.
34. Mancia G, Grassi G. Chapter 26 – The central sympathetic nervous system in hypertension.
In: Buijs RM, Swaab D, editors. Handbook of clinical neurology. Autonomic nervous system.
New York: Elsevier; 2013. p. 329–35.
35. D’Negri CE, Marelich L, Vigo D, Acunzo RS, Girotti LA, Cardinali DP, Siri LN. Circadian
periodicity of heart rate variability in hospitalized angor patients. Clin Auton Res.
2005;15:223–32.
36. Tan AY, Verrier RL. Chapter 12 – The role of the autonomic nervous system in cardiac arrhyth-
mias. In: Buijs M, Swaab D, editors. Handbook of clinical neurology. Autonomic nervous
system. New York: Elsevier; 2013. p. 135–45.
312 7  Clinical Implications of the Enlarged Autonomic Nervous System

37. Wehrwein EA, Joyner MJ. Chapter 8 – Regulation of blood pressure by the arterial baroreflex
and autonomic nervous system. In: Buijs RM, Swaab D, editors. Handbook of clinical neurol-
ogy. Autonomic nervous system. Elsevier; 2013. p. 89–102.
38. Chapleau MW. Chapter 33 – Baroreceptor reflexes. In: Biaggioni I, Burnstock G, Low PA,
Paton JFR, editors. Primer on the autonomic nervous system. Third ed. San Diego: Academic;
2012. p. 161–5.
39. Jordan J. Chapter 72 – Baroreflex failure. In: Biaggioni I, Burnstock G, Low PA, Paton JFR,
editors. Primer on the autonomic nervous system. Third ed. San Diego: Academic; 2012.
p. 349–53.
40. Jones PK, Shaw BH, Raj SR. Clinical challenges in the diagnosis and management of postural
tachycardia syndrome. Pract Neurol. 2016;16:431–8.
41. Wieling W, Groothuis JT. Chapter 39 – Physiology of upright posture. In: Biaggioni I,
Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous system. Third ed.
San Diego: Academic; 2012. p. 193–5.
42. Johns EJ. Chapter 17 – Autonomic regulation of kidney function. In: Buijs RM, Supowit SC,
editors. Handbook of clinical neurology. Autonomic nervous system. New York: Elsevier;
2013. p. 203–14.
43. Hart EC, Charkoudian N, Joyner MJ. Chapter 54 – Sex differences in autonomic function.
In: Biaggioni I, Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous
system. Third ed. San Diego: Academic; 2012. p. 261–4.
44. Brooks VL, McCully BH, Cassaglia PA. Chapter 55 – Autonomic control during pregnancy.
In: Biaggioni I, Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous
system. Third ed. San Diego: Academic; 2012. p. 265–8.
45. Lipsitz LA, Novak V. Chapter 56 – Aging and the autonomic nervous system. In: Biaggioni I,
Burnstock G, Low PA, Paton JFR, editors. Primer on the autonomic nervous system. Third ed.
San Diego: Academic; 2012. p. 271–3.
46. Anandan C, Khuder SA, Koffman BM. Prevalence of autonomic dysfunction in hospitalized
patients with Guillain-Barre syndrome. Muscle Nerve. 2016 Dec 31. doi: 10.1002/mus.25551.
[Epub ahead of print]
47. El-Abassi R, Singhal D, England JD. Fabry’s disease. J Neurol Sci. 2014;344(15–9):5–19.
48. Arora S, Young S, Kodali S, Singal AK. Hepatic porphyria: a narrative review. Indian
J Gastroenterol. 2016;35:405–18.
49. Chelimsky G, Chelimsky T. Unusual structural autonomic disorders presenting in pediatrics:
disorders associated with hypoventilation and autonomic neuropathies. Pediatr Clin N Am.
2017;64:173–83.
50. Sluka KA, Clauw DJ. Neurobiology of fibromyalgia and chronic widespread pain.

Neuroscience. 2016;338:114–29.
51. Morton DL, Sandhu JS, Jones AK. Brain imaging of pain: state of the art. J Pain Res.
2016;9:613–24.
52. Penfold S, St DE, Mazhar MN. The association between borderline personality disorder, fibro-
myalgia and chronic fatigue syndrome: systematic review. BJPsych Open. 2016;2:275–9.
53. Martinez-Martinez LA, Mora T, Vargas A, Fuentes-Iniestra M, Martinez-Lavin M. Sympathetic
nervous system dysfunction in fibromyalgia, chronic fatigue syndrome, irritable bowel
syndrome, and interstitial cystitis: a review of case-control studies. J Clin Rheumatol.
2014;20:146–50.
54. Periard JD, Travers GJ, Racinais S, Sawka MN. Cardiovascular adaptations supporting human
exercise-heat acclimation. Auton Neurosci. 2016;196:52–62.
55. Perlman R, Callum J, Laflamme C, Tien H, Nascimento B, Beckett A, Alam A. A recommended
early goal-directed management guideline for the prevention of hypothermia-related transfu-
sion, morbidity, and mortality in severely injured trauma patients. Crit Care. 2016;20:107.
Clinical Implications of the Timed
Autonomic Nervous System 8

Abstract
There are many reasons for the lack of sleep in our society that operates 24 h a day,
7 days a week, constantly. The main disturbing factor has been the technological
advance of being able to light our evenings artificially. This Chapter analyzes the
impact of the lack of sleep in the “24/7 Society”, principally the disruption of the
three ANS physiological programs. Typical examples of a desynchronized ANS
are jet lag, shift work and chronodisruption, the metabolic syndrome and mental
illnesses being also examples of a desynchronized ANS. The chronobiological
aspects of normal and pathological brain aging and cancer are discussed.

Keywords
24/7 Society • Adjuvant chronobiological treatment • Chronodisruption
• Chronodisruption in cancer • Chronodisruption in mental illnesses
• Chronodisruption in metabolic syndrome • Chronodisruption of brain aging
• Jet lag • Light at night • Melatonin • Shift work

Objectives
After studying this chapter, you should be able to:
• Describe the mechanisms underlying the inadequate experience of day and
night in the “24/7 Society.”
• Describe the disruption of the three ANS physiological programs in a 24-h
cycle due to the inadequate experience of day and night.
• Describe the physiological changes observed in jet lag, shift work. and
chronodisruption as examples of a desynchronized ANS.
• Understand why the chronobiological treatment of a desynchronized ANS
is needed for full recovery in most cases.
• Describe the metabolic syndrome and mental illnesses as examples of a
desynchronized ANS.
• Understand the chronobiological aspects of normal and pathological brain aging.
• Understand the chronobiological aspects of cancer.

© Springer International Publishing AG 2018 313


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1_8
314 8  Clinical Implications of the Timed Autonomic Nervous System

 ue to the “24/7 Society,” the ANS Has Lost Adequate


D
Experience of Day and Night

There are many reasons for lack of sleep in our society, which operates 24 h a day,
7 days a week, constantly. The early hours of school or work, TV programs increas-
ingly displacing the “prime time” ones late at night, the daily stress, or the wide-
spread use of foods and beverages rich in caffeine are among the precipitating
factors. However, the main factor has been the technological advance being able to
artificially light our evening.
Our hominid ancestor, Homo erectus, used caves as shelter and must have begun
to use fire about a half million years ago. Homo sapiens built artificial shelters pro-
tected from the sun’s rays and manufactured lamps that allowed him to extend the
daily lighting period about 70,000 years ago. The first lamp invented was made of a
shell, hollowed-out rock, or other similar nonflammable objects and was filled with
a combustible material (probably dried grass or wood), sprinkled with animal fat
(the original lighter fluid) and ignited.
In the last 200 years, we have shifted our routines from rural environments to the
cities and from outdoor life to confinement in our homes. Furthermore, with the
advent of electric light we have become progressively isolated from the natural
cycles of light and darkness that shaped our biological rhythms for millions of
years. This is an environmental mutation, with an increasing impact on the quantity
and quality of our sleep.
The internet explosion has added a further complication. Increasingly, individu-
als spend part of their nights in front of lit monitors (LCD, tablets, smartphones),
screens that produce at least two phenomena of concern to the sleep–wake cycle: (a)
plundering the natural period of sleep by reducing it to dangerous levels; (b) adding
a disruptor factor, the monitor light during the circadian period causes phase delays
of the biological clock, which produces a later sleep on subsequent nights, tending
to perpetuate the situation of nocturnal sleep deprivation [1].
The artificial light striking the retina from dusk until dawn exerts a strong inhibi-
tory activity of hypothalamic neurons that induces sleep and a strong excitatory
activity of brain mechanisms that maintains wakefulness. As we have already dis-
cussed in Chap. 2, suppression of the nocturnal release of melatonin occurs, which
is responsible for synchronizing our circadian rhythms and for the “opening of the
sleep gate,” (Fig. 2.12) [2]. If someone tries to sleep at 18:00 or 19:00 h, he or she
would probably take a 1- to 2-h nap. If instead the time to sleep onset shifts to 21:00
or 22:00 h, a 6- to 8-h consolidated sleep tends to occur. This sharp increase in sleep
propensity is due to inhibition of wakefulness promoting action of the SCN given
by melatonin. Ambient light by inhibiting the secretion of melatonin, reduces sleep-
iness, promotes alertness, and interferes with sleep.
In daylight conditions, the SCN exerts maximum wake-up activity toward the
end of the period of wakefulness, a “second wind” that keeps us awake despite
the sleep debt already accumulated in wakefulness. Before the widespread use
of electric light, people experienced this “second wind” in the afternoon (17:00
to 18:00 h) and possibly kept alert until nightfall. However, exposure to artificial
Due to the “24/7 Society,” the ANS Has Lost Adequate Experience of Day and Night 315

24/7 Society

Circadian misaligment

Sleep-wake Sleep disorders Disturbed sociotemporal


disturbances and behavioral patterns

Cronic sleep
Mood disturbances
restriction

Internal
desynchronization

Melatonin suppression

DISEASE

Fig. 8.1  The impact of 24/7 society. Reproduced with permission from Cardinali [75]

light after dark has incorporated a signal indicating to the SCN a time of day
that is not real, deferring the “second wind” and delaying the secretion of mela-
tonin (Fig. 8.1).
Indeed, technology has disconnected us from the natural 24-h day in which our
species has evolved. A comparison of the biological effects of reading an electronic
book on a light-emitting device with reading a printed book in the hours before
bedtime indicated that those reading an electronic book took longer to fall asleep
and had reduced evening sleepiness, reduced melatonin secretion, later timing of
their circadian clock, and reduced next-morning alertness than those reading a
printed book [1].
Because about 20% of electricity consumption worldwide is devoted to the pro-
duction of light, many governments are eliminating traditional incandescent lamps
(emitting in the red) for more efficient LEDs in search of savings. However, this
white light solid state is typically rich in blue light, which is the portion of the spec-
trum that most inhibits photoreceptor retinal ganglion cells and hence the secretion
of melatonin, which further amplifies the disruptive effect on sleep–wake rhythm
[3]. Therefore, sleep topics are not only a medical problem, but substantially influ-
ence the social and organizational frame of society.
Both longitudinal statistics in northern hemisphere countries and regional data in
Latin America indicate that in just 40–50 years, we have reduced our sleeping time
by 25% [4]. Longitudinal statistics from the National Sleep Foundation, USA, indi-
cate that the number of daily hours of sleep has fallen since 1960 to date from 8.2 to
6 h daily. In our study, we verified that 65% of the population, regardless of age,
reported having sleep disorders in the last 12 months; 40% of these disorders were
described as moderate to severe, and there was a sleep deficit of about 2 h a day. All
respondents recognized the negative consequences of poor sleep for health and
quality of life [5].
316 8  Clinical Implications of the Timed Autonomic Nervous System

In the USA, 30% of employed adults and 44% of night workers sleep less than
6 h per night, versus less than 3% of the adult US population 50 years ago [6].
Globally, children sleep about 1.2 h less on weekdays before school activity than a
century ago. It is noteworthy that children tend to become hyperactive instead of
sleepy when they do not get enough rest and have difficulty in concentrating, focus-
ing attention; thus, deficiency of sleep can be confused with hyperactivity/attention
deficit disorder, a condition “over-diagnosed” in many societies [7].

 he Disruption of the Three ANS Physiological Programs


T
(“Body Configurations”) Is a Major Consequence for the “24/7
Society”

We discussed in Chap. 2 how the three different “bodies” (wakefulness, slow-wave


sleep, and REM sleep) necessarily follow each another harmoniously to ensure
health. A 76-year-old man (the current life expectancy in our society) sleeping 8 h
daily will live 50 years in the physiological state of wakefulness, 20 years in slow-­
wave sleep, and 6 years in REM sleep. However, our society has reduced by 25%
the amount of time spent sleeping over the last 40 years. Therefore, the above cal-
culation now changes to a distribution of 55 years of wakefulness, 15 years of slow-­
wave sleep and 6 years of REM sleep (Fig. 8.2).
A timed view of the ANS from clinical standpoints allows interpretation of
symptom intensity and mortality of human diseases, conditions, and syndromes that
exhibit a 24-h pattern and are thus related to the three body configurations. Twenty-­
four-­hour patterns are characteristic of more than 100 acute and chronic common
and rare human diseases (Fig. 8.3) [8]. The worst symptoms of many of these tend
to be expressed either late evening/overnight or early morning, thus significantly
compromising nocturnal sleep, daytime productivity, and overall quality of life, and
are therefore of great relevance clinically to patient management, for example [8]:

• Cardiac—atrial premature beats and tachycardia, paroxysmal atrial fibrillation,


atrial–ventricular block, paroxysmal supraventricular tachycardia, ventricular
premature beats, angina pectoris, acute (nonfatal and fatal) incidents of myocar-
dial infarction, sudden cardiac arrest, acute cardiogenic pulmonary edema, heart
failure
• Vascular and circulatory system—hypertension, acute hypotension/syncope,
intermittent claudication, venous insufficiency, standing occupation leg edema,
arterial and venous branch occlusion of the eye, menopausal hot flash, sickle cell
syndrome, abdominal, aortic, and pulmonary thromboembolism, deep venous
thrombosis, cerebrovascular transient ischemic attack, stroke
• Respiratory—viral and allergic rhinorrhea, reversible (asthma) and nonrevers-
ible (bronchitis and emphysema) chronic obstructive pulmonary disease, cystic
fibrosis, high-altitude pulmonary edema
• Gastrointestinal tract—esophageal reflux, peptic ulcer, cyclic vomiting syn-
drome, biliary colic, hepatic variceal hemorrhage
• Renal—colic, nocturnal enuresis, polyuria
The Disruption of the Three ANS Physiological Programs (“Body Configurations”) 317

Life expectancy: PARASYMPATHETIC Adaptive cellular immunity


76 years SYSTEM ↓Blood pressure

GH, insulin, leptin
Sleep Anorexia
8 h/day nR
EM
sle R
EM
Aw ep
ak sl
ee
e p

50 years 20 years 6 years

SYMPATHETIC
SYSTEM Somatic and
autonomic
Natural immunity (NK); humoral paralysis

Hypercoagulability Declarative Reflexive Emotional


memory memory memory
Vasoconstriction
Memory
Ghrelin, ↑Food intake maintenance

Life expectancy: ↓ Adaptive cellular immunity


PARASYMPATHETIC
76 years ↑ Blood pressure
SYSTEM
↓ GH, insulin, leptin

↓ Anorexia
Sleep nR
6 h/day EM
sle
ep R
EM
Aw sl
ak ee
e p
15 years
55 years 6 years

SYMPATHETIC
SYSTEM Somatic and
autonomic
Inflammation paralysis

Atherothrombosis Declarative Reflexive Emotional


memory memory memory
Arterial hypertension
Memory
Metabolic syndrome, type 2-type 3 diabetes maintenance

Fig. 8.2  The three different “bodies,” wakefulness, slow-wave sleep (NREM sleep) and REM
sleep, must necessarily follow each another harmoniously to ensure health. Upper panel: a 76-year-­
old man sleeping 8 h daily will live 50 years in the physiological state of wakefulness, 20 years in
slow-wave sleep and 6 years in REM sleep. Lower panel: reduction of 25% of sleep over the last
40 years leads to predominance of the wakefulness state and reduction of the slow-wave sleep,
associated with cardiovascular disease, metabolic syndrome, obesity, and type II and type III
diabetes
318 8  Clinical Implications of the Timed Autonomic Nervous System

Gout attacks
Menopause hot flashes Gallbladder attacks

Restless Leg Syndrome GERD, peptic ulcer attacks

Congestive heart failure,


Backache Acute cardiogenic
24 pulmonary edema
22 2
20 4 Cluster and migraine
Increased cholesterol
headaches
synthesis 18 6
Asthma attacks
Osteoarthritis pain worsens 16 8
14 10 Death (all causes)
12
Instestinal ulcer perforation
Alergic rhinitis, Cold and flu

Tension headache onset Reumathoid rthritis

Gastric ulcer perforation Depression worst


Angina pectoris, Myocardial
infarct, Sudden cardiac
death, Stroke

Fig. 8.3  Chronopathology. Acrophases for several clinical entities across a 24-h cycle

• Neural—frontal, parietal, temporal, and occipital lobe seizures, Parkinson’s and


Alzheimer’s disease, hereditary progressive dystonia
• Psychiatric/behavioral—major and seasonal affective depressive disorders, bipo-
lar disorder, dementia-associated agitation, addictive alcohol, tobacco, and her-
oin cravings, and withdrawal phenomena
• Pain—cancer, post-surgical, diabetic neuropathic and foot ulcer, tooth caries,
fibromyalgia, sciatica, multiple sclerosis muscle spasm, and migraine (tension,
cluster, hypnic, paroxysmal hemicranial headache)
• Autoimmune and musculoskeletal—rheumatoid arthritis, osteoarthritis, spondy-
larthrosis, gout, Sjögren’s syndrome, systemic lupus erythematosus
• Infection—susceptibility, fever, mortality
• Skin—atopic dermatitis, urticaria, psoriasis, palmar hyperhidrosis
• Ocular—bulbar conjunctival redness, keratoconjunctivitis sicca, intraocular
pressure, anterior ischemic optic neuropathy, recurrent corneal erosion
syndrome

The imbalance at the expense of slow-wave sleep imposed by the “24/7 Society”
(Fig. 8.2) can be costly in countless aspects of our life and health. The impact of
sleep deprivation is widespread and affects not only the physical, but also the psy-
chological and social wellbeing. With impaired cognitive performance in the areas
of attention, memory, and executive functions, the added emotional and behavioral
consequences of sleep deprivation may explain the exasperated social behavior
found in our present life [9].
The predominance of the sympathetic configuration of wakefulness is a strong
predisposing factor for chronic, low-degree inflammation. The term
Jet Lag, Shift Work, and Chronodisruption as Examples of a Desynchronized ANS 319

“inflammaging” has been coined to denominate the contribution of inflammatory


processes to the progression of aging [10]. However, inflammation is not only a
matter of normal senescence, but is also observed in several diseases that can be
linked to the predominant sympathetic configuration given by prolonged
wakefulness.
Chronic, mild inflammation is multiply intertwined with other potentially dete-
riorating processes, among which mitochondrial dysfunction is of premier impor-
tance [10]. An abnormally prolonged sympathetic predominance, as in sleep
deprivation, implies several immune remodeling processes that include tendencies
toward enhanced proinflammatory signaling (Chap. 4).
Numerous epidemiological studies indicate the association of sleep deprivation
with cardiovascular disease, metabolic syndrome, obesity, and T2DM [4, 11].
Moreover, today the association of these symptoms is emphasized, with dementia,
particularly Alzheimer’s disease, which is often called “type 3 diabetes”. There is
also epidemiological evidence for the link between poor sleep when working shifts
and cancer, especially breast cancer [12].

J et Lag, Shift Work, and Chronodisruption as Examples


of a Desynchronized ANS

Changes in the environmental timing cues (zeitgebers) are associated with altera-
tions in the body’s 24-h rhythms (Chap. 2). Under most circumstances, this adjust-
ment to the environmental light/dark cycle is normal and consistent. However,
mismatches can occur when changes in environmental demands are either sudden or
severe. In shift work, there is a phase shift of the activity/rest cycle with regard to
the light/dark cycle, whereas jet-lagged time zone travelers encounter a pattern of
light and darkness, activity, and social schedules shifting together in time [13]. The
endogenous circadian system is slow to adapt to new time cues, and until the correct
phase relationship between biological rhythms and external zeitgebers is re-­
established, a host of physiological and behavioral problems can manifest. Similar
problems are encountered by shift workers operating under new work schedules out
of phase with the normal light/dark cycle, other competing zeitgebers, and their
endogenous body clock.
Chronodisruption comprises the changes in amplitude and phase of circadian
rhythms that are found as a comorbidity of most acute and chronic diseases. The
observed chronodisruption may be due to a failure in one or more components
of the clock itself, in the output signal to the different systems, in the presence
of synchronizers, or in the transmission of information from the zeitgebers to
the circadian clock (Fig. 8.4). In the clinic, the exact cause of an alteration is
unlikely to be known, and in any case, most are multifactorial (Table 8.1)
[14–17].
As discussed in Chap. 2, almost every physiological function has a circadian
“phase map” consisting of an ordered sequence of peaks and valleys (Fig. 2.1).
320 8  Clinical Implications of the Timed Autonomic Nervous System

Chronodisruption

Disrupted Acute or chronic


sleep phases of most
diseases.
Late wake up Light at night.
(lack of morning exposure to light)

Early desynchronization
(circadian desynchronization)
Insomnia
Genetic
and Sleep
Vicious
epigenetic deficiency
circle
factors Excessive
daytime
Chronic desynchronization sleepiness

Decreased melatonin Reduced physical


levels activity

Fig. 8.4  Chronodisruption found as a comorbidity of most acute and chronic diseases. The
observed chronodisruption may be due to a failure in one or more components of the clock itself,
in the output signal to the different systems, in the presence of synchronizers, or in the transmission
of information from the zeitgebers to the circadian clock

Table 8.1  Etiology, symptoms, and treatment of chronodisruption


Acute or chronic
phases of most
diseases Alteration of the central circadian pacemaker
Sleep deprivation, indicated by the alteration in melatonin
Etiology light at night (LAN) secretion
Symptoms (phase and ANS dysfunction Insomnia, somnolence, metabolic syndrome,
amplitude changes of thoracic-muscular sympathetic predominance,
circadian rhythms) parasympathetic abdominal predominance,
gastrointestinal disorders
Somatic alterations Tiredness, fatigue, stiff neck/back, low back
pain, headache, drowsiness
Cognitive functions Disorientation, loss of sociability and
motivation, poor attention, performance,
memory, or concentration
Behavior Aggressiveness, impulsivity, hyperactivity,
irritability
Psychiatric Depressive symptoms, personality disorders,
disorders anxiety
Therapeutic approach Chronobiological treatment (exposure to morning light, melatonin
supplementation at bedtime)
In the case of acute or chronic diseases, the chronobiological
treatment must accompany the specific treatment for “ad integrum”
recovery
Jet Lag, Shift Work, and Chronodisruption as Examples of a Desynchronized ANS 321

While the period of the phase map for different physiological functions is similar,
the peaks and valleys of the maps generally do not coincide. Phase maps are also
sensitive to environmental changes and can be transiently affected by temporal dis-
ruptions such as jet-lag disorder or shift work, or by the disruption imposed by an
acute or chronic illness, regardless of its severity.

Jet Lag

Flight dysrhythmia, more commonly known as jet lag, comprises a constellation of


symptoms consisting of daytime fatigue, impaired alertness, nighttime insomnia,
loss of appetite, depressed mood, poor psychomotor coordination, and reduced cog-
nitive skills, among other (Table 8.2) [18]. These symptoms are caused by the

Table 8.2  General symptoms of jet lag


Anorexia or loss of appetite
Apathy
Bowel irregularities (constipation or frequent defecation)
Clumsiness
Daytime somnolence
Decreased vigilance and attention domains
Depression
Diminished mental abilities (i.e., cognitive performance, concentration, judgment, decision
making, memory lapses)
Diminished physical performance
Disorientation
Fuzziness
Gastrointestinal symptoms (e.g., bloating and upset stomach)
General feeling of malaise
Generalized fatigue and lethargy
Glucose metabolism dysregulation
Headache
Impaired alertness
Impaired task performance (increased accidents and errors)
Inappropriate timing of defecation and urination
Irritability
Menstrual irregularity
Mood disturbances
Muscular pain
Sleep loss
Sleeping difficulties (inappropriate sleep at local time)
Slowed reflexes
Stress
Tiredness (traveler’s fatigue)
Traveler’s thrombosis (deep vein thrombosis)
Trouble initiating and maintaining sleep
Tumor progression is noted in chronic animal model
Please note that given the inter-individual variability and susceptibility, the core symptoms vary
and not necessarily every individual experiences or exhibits the entire spectrum of symptoms
322 8  Clinical Implications of the Timed Autonomic Nervous System

temporary misalignment between the circadian clock and external time, which
occurs because of rapid travel across time zones. The number of time zones crossed
and the direction of travel influence the severity of jet lag symptoms. Eastward
travel tends to cause difficulty in falling asleep whereas westward travel usually
interferes with sleep maintenance.
In jet lag, the recovery time for restoring the normal rhythm profile (re-­
entrainment) can differ significantly from one physiological function to another.
Although during the period of re-entrainment individual circadian rhythms gen-
erally move in a direction that corresponds with that of the environmental time
shift, in some cases the circadian system moves in a direction that is opposite to
that of the environmental change, giving rise to a phenomenon called “splitting.”
Phase map “partitioning” is then more complex, involving a partial re-entrain-
ment by some phase maps in a direction that is opposite to that of other phase
maps [13].
Jet lag affects the health status of frequent air travelers. The disruptive effect of
jet lag has been documented in experimental animals at the molecular level of clock
genes in the SCN and in the clock genes present in peripheral tissues [19]. Eastbound
travel causes a phase advance in all the body’s circadian rhythms, whereas west-
ward flight has the opposite effect, i.e., it produces a phase delay. Consequently,
travelers tend to synchronize their bodily rhythms at a speed of 1.5 h a day after
westward and 1 h a day after eastward flight, irrespective of whether they travel dur-
ing the day or at night. This difference in adjustment time is usually attributed to the
greater ability of the internal body clock to adapt to a longer rather than to a shorter
day (τ longer than 24 h, Chap. 2).
Jet-lag disorder falls under the category of circadian rhythm sleep disorders. The
symptomatology of jet-lag disorder includes both physiological and psychological
disturbances (Table 8.2). Several studies suggest that chronic or repeated jet-lag
exposure can lead to cognitive decline and temporal lobe atrophy in humans if there
is a short recovery time between flights. Model jet-lag disorder is also associated
with tumor progression in rodents and elevated mortality rate in aged mice [13].
Jet-lag disorder symptoms show considerable inter- and intra-individual vari-
ability. Age is one important factor. In simulated jet-lag disorder, middle-aged male
subjects had more symptoms than younger men. Moreover, those over 60 years are
reported to have greater difficulty in adapting to jet-lag disorder [20]. Another
important variable is the individual’s “chronotype” (Chap. 2). Individuals who are
“morning chronotypes” generally have less difficulty in phase-advancing their body
rhythms (i.e., adjusting after a flight from west to east) than “evening chronotypes,”
and vice versa in the case of a phase delay. Generally, those who had “rigid” sleep
habits had more severe symptoms after a transmeridian flight [21].
Several studies have examined the effects of simulated and real jet lag on physi-
ological and psychological variables in different populations including aircrew
members, and have confirmed that these frequent flyers suffer marked sleep–wake
problems because of jet lag. For example, in a 2-year collaborative field study of
Jet Lag, Shift Work, and Chronodisruption as Examples of a Desynchronized ANS 323

Spanish pilots flying the routes from Madrid, Spain, to Mexico City, Mexico (−7
time zones) or from Madrid to Tokyo, Japan (+8 time zones) we used telemetry to
record pilots’ activity, temperature, and heart rate [22, 23]. Subjective time estima-
tion and other psychological variables such as anxiety, tiredness, and performance
were recorded. Urinary 6-sulphatoxymelatonin and cortisol excretion (determined
in 6-h intervals) were also measured. Activity/rest and heart rate rhythms, linked to
a “weak” or exogenous oscillator, became rapidly synchronized, whereas tempera-
ture or 6-sulphatoxymelatonin excretion rhythms, which are closely regulated by
the biological clock (Chap. 5) showed a more rigid response after the phase shift of
the light/dark cycle [22]. In both young (<50 years old) and old (>50 years old)
pilots arriving in Mexico or Tokyo, the activity/rest rhythm rapidly adjusted to the
new schedule, whereas the acrophase of the temperature rhythm tended to fluctuate
near the original temporal zone. This desynchronization was evident until the
return flight (day 5) and persisted after arrival in Madrid [22]. The sequelae of
desynchronization were less tolerable in older than younger pilots. Skin tempera-
ture rhythm did not become entrained neither on reaching Tokyo nor after the
return flight to Madrid in the group of older pilots [22]. The changes in urinary
6-­sulphatoxymelatonin and cortisol excretion were consistent with these
conclusions.
Systematic and incorrectly planned work schedules of airline pilots produce a
chronic disruptive condition, a higher incidence of stress-related emotional changes,
and a diminished life expectancy. The optimal work strategy for this population is a
compromise between two extreme possibilities: a long rest period at stopovers until
full re-entrainment is achieved, or a short stop accompanied with relative isolation,
maintaining the original “home” local habits to prevent re-entrainment. With the
first strategy, aircrew would be systematically exposed to a re-entrainment process,
with a permanent disruption to the circadian system, whereas the second approach,
although less disturbing to the body clock, would probably not allow pilots to have
the necessary rest and alertness for the return flight [24].
Sleep deprivation produces an allostatic overload that can have deleterious con-
sequences [25]. Restriction of sleep to 4 h per night is associated with increases in
BP pressure, decreases in parasympathetic tone, increases in evening cortisol and
insulin levels, and increases in appetite through the elevation of ghrelin, a pro-­
appetitive hormone, and decreases in the levels of leptin, which has anorexic activ-
ity (Chap. 5). Proinflammatory cytokine levels are also increased, along with
decreases in performance in tests of psychomotor vigilance after a modest sleep
restriction to 6 h per night. Allostatic overload in animal models causes atrophy of
neurons in the hippocampus and prefrontal cortex, the brain regions involved in
memory, selective attention, and executive function. It also causes hypertrophy of
neurons in the amygdala, the brain region involved in fear and anxiety, and aggres-
sion (Chap. 6). Thus, the ability to learn and remember and to make decisions may
be compromised and may be accompanied by increased levels of anxiety and
aggression.
324 8  Clinical Implications of the Timed Autonomic Nervous System

Both long-term and short-term exposure to transmeridian flights have an impact on


cognitive functioning. For example, in a group of individuals who were on a trans-
meridian flight and who underwent functional magnetic resonance imaging study, par-
ticipants from the jet-lag group presented decreased activation in the bilateral medial
prefrontal and the anterior cingulate cortex [26]. The results are suggestive of a nega-
tive impact of jet lag on important cognitive functions such as emotional regulation and
decision-making during the first few days after individuals arrive at their destination.

Shift-Work Disorder

A wide range of work schedules is referred as “shift work.” They include occasional
on-call overnight duty, rotating schedules, and steady, permanent night work
(Fig. 8.5) [6]. Owing to the overlap of these categories, it is difficult to generalize
about shift work disorder. Over 10% of night workers and of rotating workers met
the minimal criteria for shift work disorder.

60

50

40
Percentage

30

20

10

0
es

n
g

s
n

an
ce

le
or

tio
in

tio
ic

Sa
rv

pp

vi

ci

uc
ta
rv

se

er

ni
or

su
se

od
ch
/s
n/

sp

Pr
re
e

re
io

/te
an
tiv

ca
at

ca

rs
ec

Tr
ar

lth

ne
al
ot

ep

ea

on

io
Pr

pr

tit
H

rs

ac
od

Pe

pr
Fo

re
ca
lth
ea
H

Fig. 8.5  Prevalence of shift work. Reproduced with permission from Cardinali [75]
Jet Lag, Shift Work, and Chronodisruption as Examples of a Desynchronized ANS 325

Several studies have now confirmed that there is a relationship between cardiovas-
cular disease and shift work (Fig. 8.6). Shift workers are at a 40% higher risk of
developing ischemic heart disease. The association between shift work and meta-
bolic syndrome, a major risk factor for cardiovascular disease, also occurs. Alternating
shift work has been reported to be a significant independent risk factor for high BP,
an effect that was more pronounced than that of age or body mass index [6].
In 2007, the International Agency for Research on Cancer classified shift work
as a probable human carcinogen (2A). Women who work on rotating night shifts
are reported to be at a moderately increased risk of breast cancer after extended
periods of working night shifts [27], as are female cabin crew [28]. Because mela-
tonin has oncostatic effects, including effects on estrogen and fat metabolism, it
may play a role in both breast and endometrial cancer. Light exposure at night
reduces melatonin levels, but the role of melatonin in both of these cancers remains
to be defined.
Misalignment between the circadian pacemaker and the timing of sleep, wake,
and work occurs in shift workers, and shift work disorder, with insomnia, reduced

SHIFT WORK

interindividual variation
in resilience
Insufficient Biological rhythms
Sleep dyssynchrony
Power
nap

Neuroendocrine stress Immune response Oxidative stress

¯ melatonin  leukocytes ¯ NK cells


 cortisol  proinflammatory cytokines
 catecholamines  C reactive protein

 Pro- ¯Immune  Pro- Antioxidant


inflammatory defense oxidative state
state state

 Blood pressure  Atherogenesis

Cardiovascular risk  Carcinogenesis

Fig. 8.6  Potential physiopathological pathways by which shift work may lead to cardiovascular
disease and cancer. Experimental circadian misalignment and sleep restriction protocols disrupt
and enhance the activity of neuroendocrine stress systems, reduce immune defense (NK cells), and
cause inflammation and oxidative damage. Inter-individual vulnerability to the adverse effects of
sleep restriction and circadian misalignment contribute to a heterogeneous tolerance to shift work.
Prophylactic naps could blunt the stress response, possibly correct stress-dependent immune
changes, and improve the recovery of immune homeostasis. Reproduced with permission from
Cardinali [75]
326 8  Clinical Implications of the Timed Autonomic Nervous System

sleep, and excessive sleepiness, is common. All these impair cognitive function,
alertness, and mood, and increase the risk of accidents. For years, the public, media,
and regulatory authorities have blamed the effects of excessive speed and alcohol as
the main causes of road accidents. However, it is important to note that the lack of
sleep produces the same effects on the ability to drive a vehicle as drinking alcohol.
In psychometric studies, to be awake for 17–18 h disturbs the ability to drive a
vehicle in a similar manner to the effect of an alcohol concentration in the blood of
0.05 g/dL [29]; moreover, both situations may add up to decreased attention.
Between 20 and 25% of road accidents are caused by fatigue and sleepiness of driv-
ers, being more frequent between 0200 and 0800 h. This trend is particularly evi-
dent on motorways and monotonous routes.
In several studies on this subject, high number of drivers refer to often being
drowsy at the wheel. Clearly, there are more sleepy drivers than drunk ones on
roads. For example, in representative samples of public transportation drivers in the
Metropolitan Area of Buenos Aires and long-distance drivers covering various geo-
graphical corridors of the country, we conducted surveys on health and working
conditions and applied objective measures of physiological variables, including
evaluation of the sleep/wake rhythm using actigraphy, circadian rhythmicity by the
peripheral rhythm of body temperature, alertness by determining psychomotor
response to a stimulus, autonomic activity by heart rate variability, and endocrine
response to stress by measuring cortisol in the saliva [30, 31].
In short-distance drivers, a high prevalence of work-related stress, overweight,
obesity, physical inactivity, and hypertension was observed. The quantity and
quality of sleep on weekdays was poor, with partial recovery at the weekend, a
high frequency of daytime sleepiness, and high risk of apnea. The neurohor-
monal weekday pattern was consistent with stress and a significant drop in psy-
chomotor performance was observed during working hours, especially the
morning shift [30, 31].
In long-distance drivers, a high prevalence of cardiovascular risk factors, such as
overweight, physical inactivity, and smoking, was also found. Sleep patterns of poor
quality, with little sleep at home, while traveling, and at the destination, and a
decrease in amplitude circadian rhythms, were observed. The pattern was consistent
with high cortisol levels, with little recovery in the days out of work, and a decrease
in alertness at the end of the return trips [30, 31].
In addition to the known effects on sleep, eating patterns, and alterations in social
life, gastrointestinal disorders are very common in shift workers. The association
between rotational work shifts and gastrointestinal disorders has several causes.
Irregular ingestion habits of workers affect the synchronization of numerous circa-
dian rhythms, in particular, those related to digestive functions and metabolism
[32]. Gastrointestinal disorders may be due to ingestion of food at the “wrong”
times, which induces anomalous patterns of motility and digestive secretions. In
addition, the absence of hot food, which occurs frequently during the night (with a
predominance of snacks), a high carbohydrate intake, caffeine and alcohol, and high
consumption of tobacco have all been proposed as causes of gastrointestinal disor-
ders in shift workers.
Jet Lag, Shift Work, and Chronodisruption as Examples of a Desynchronized ANS 327

Chronodisruption

This term defines the changes in amplitude and phase of circadian rhythms that are
found in acute or chronic illness, even in mild situations such as common flu (“poor
wakefulness, poor sleep”) [14–17]. The clock itself, the information pathways from
synchronizing agents to the clock or the efferent pathways of the clock can be
affected alone or conjointly (Fig. 8.4). The deterioration of circadian rhythms with
age is an example of a change with many components, which implies a decrease in
the effectiveness of many aspects of the circadian system, from a reduced influence
of the synchronizers, to a deterioration of the clock itself, to a decrease in the ability
to obey the clock output signal [33].
The observation that the circadian system is not functioning normally does not
necessarily imply that it is the primary cause of the alteration. Rather, the primary
defect may have originated elsewhere, and its effects on the circadian system be
one of the many changes it causes. In this case, and although treating the circadian
system could improve the individual’s nocturnal sleep and diurnal activity, we
would not be addressing the real cause of the problem. A combined (specific and
chronobiological) treatment is needed for full recovery (Figs. 8.7 and 8.8)
[14–17].

Specific
Treatment

Acute or Chronic
Disease

Adjuvant
Concomitant
Chronobiologic
Dyssynchrony
Treatment

“Ad
integrum”
Recovery

Fig. 8.7  The concomitant chronodisruption occurring in most acute or chronic diseases must be
adequately treated to obtain full recovery of health. Reproduced with permission from Cardinali [75]
328 8  Clinical Implications of the Timed Autonomic Nervous System

Morning light
Bases for the
adjuvant
chronobiologic
treatment
SCN

SLEEP

WAKEFULNESS
O
NH
SCN
CH3 MT1

H3C O

N
H
WAKEFULNESS

Melatonin at bed time SLEEP

Fig. 8.8  Exposure to light in the morning and the administration of melatonin in the evening
provide a strong synchronization signal and an increase in the amplitude of sleep/wake cycle. From
clinical standpoints, the changes in amplitude (“poor sleep together with poor vigilance”) is a para-
mount sign of the disease and its correction increases substantially the patient’s quality of life. The
morning light and melatonin in evening hours are the natural resources to restore proper rhythmic-
ity of sleep/wake rhythm. Reproduced with permission from Cardinali [75]

Chronobiological Treatment of a Desynchronized ANS

The jet lag and the alterations generated by shift work are two situations that reflect
the functioning of a circadian system of a normal subject that has not adjusted to a
schedule change of the sleep/wake cycle. Consequently, the treatment of these prob-
lems consists in the use of alternative zeitgebers that allow the adjustment of the
circadian clock to the new schedule, or to avoid the adjustment in situations where
it is unnecessary, as in rapid rotation shifts, or when the return flight takes place as
soon as within 1 or 2 days. On the other hand, the treatment of clinical problems
associated with circadian rhythm alterations needs both specific therapies and an
adjuvant chronobiological treatment to obtain an optimal result (Figs. 8.7 and 8.8).
To successfully overcome the effects of jet-lag disorder, adjustment to the new
time zone can be encouraged by adopting the social timing of life in the new time
zone as soon as possible. In field and simulation studies suitably timed melatonin
administration has been shown to accelerate phase shifts and to significantly
improve self-rated jet-lag disorder symptoms in large numbers of time zone travel-
ers. In addition, exposure to light has been shown to accelerate phase shifts. The
combination of melatonin and light exposure, one in the evening and the other in the
morning, is more effective than either treatment alone (Fig. 8.7) [20].
The effect of light or melatonin on the circadian system can be measured by a
phase response curve in which the minimum in core body temperature is used as an
estimate for the crossover point of the curve (Fig. 8.9). Light pulses administered
before this point delay the circadian clock, whereas light pulses after it phase
Chronobiological Treatment of a Desynchronized ANS 329

Light

Advance
Phase change
0

Melatonin
Delay

08:00 h 16:00 h 24:00 h 08:00 h

Fig. 8.9  The phase response curves to light and melatonin are opposite, but not symmetrical.
Reproduced with permission from Cardinali [75]

advance the clock. Light exposure close to the minimum core body temperature
produces the greatest phase shifts [34]. Phase delays of approximately 2.5–3 h per
day and phase advances of 1.5–2 h per day have been observed following carefully
timed exposure to bright light [35]. The phase response curve for melatonin is the
opposite to that of the light, although not symmetrical (Fig. 8.9).
Melatonin is the prototype of a chronobiotic drug [36]. It is produced in most
organisms from algae to mammals, but its role varies considerably across the phy-
logenetic spectra [37]. In humans, melatonin plays a major function in the coordina-
tion of circadian rhythmicity, remarkably the sleep–wake cycle (Chap. 2). Melatonin
secretion is an “arm” of the biological clock in the sense that it responds to signals
from the SCN. More particularly, the timing of the melatonin rhythm indicates the
status of the clock, both in terms of phase (i.e., internal clock time relative to exter-
nal clock time) and amplitude. From another point of view, melatonin is a chemical
code of the night: the longer the night, the longer the duration of its secretion. In
many species, this pattern of secretion serves as a time cue for seasonal rhythms
(Chap. 2).
The usefulness of melatonin for ameliorating the symptoms of jet lag has been
compellingly demonstrated in numerous investigations. A meta-analysis (Cochrane
database) concluded that melatonin taken at bedtime in the place of destination
(2200 h to midnight) was effective for decreasing the jet lag symptoms in air travel-
ers who crossed five or more time zones [38].
There is considerable evidence that a light stimulus of sufficient intensity applied
at a critical circadian phase can essentially stop the human circadian clock by reset-
ting the circadian oscillator close to a phaseless position at which the amplitude of
circadian oscillation is zero, i.e., type 0 resynchronization (Fig. 8.10) [39]. Indeed,
exposure of humans to cycles of bright light, centered on the time at which the
human circadian pacemaker is most sensitive to light-induced phase shifts, can
markedly attenuate or reduce endogenous circadian amplitude.
Timed light and melatonin administration allowed an almost immediate resyn-
chronization of circadian rhythms in a group of jet air travelers who had made a
330 8  Clinical Implications of the Timed Autonomic Nervous System

Fig. 8.10  To suppress the Two symmetrical


circadian clock oscillation, light pulses "knock
the application of out" circadian
symmetrical light pulses in oscillator
the first and second part of
the night is needed. In this
“knock out” of the circadian

Body temperature
clock, melatonin
administration at the local
time has an immediate effect
of synchronization to the
new time schedule.
Reproduced with permission
from Cardinali [75]

00:00 05:00 12:00 18:00 00:00

3 mg Melatonin
p.o. at local time

transmeridian flight over 12 time zones. Under the conditions of transfer of 12 time
zones over a period of hours, a fully inverted (180°) relationship between the sub-
jective day and the geophysical day occurs. Thus, a patterned exposure to natural
light covering portions that symmetrically delay and phase advance the circadian
rhythms resulted in suppression of the circadian pacemaker function. This allowed
the use of melatonin at local night to resynchronize the circadian oscillator to the
Tokyo time. Additionally, we administered a nonphotic stimulus (exercise) in a
schedule to coincide with exposure to natural light to mask the circadian oscillator.
The observed rate of resynchronization was about 2 days, significantly different
from a minimum resynchronization of up to 8–10 days expected after a flight
through 12 time zones [40, 41].
A positive correlation was found between the pre-flight melatonin production
rates, evaluated by measuring urinary 6-sulphatoxymelatonin excretion, and sleep
quality and morning alertness after a flight [40]. It is known that individuals who
possess a weak circadian time structure, as revealed by the low amplitude of body
temperature rhythm, are more prone to developing biological intolerance to shift
work [42] and presumably to jet lag. In Table 8.3 the tentative recommendations for
flights through >8 time zones are summarized.
Among the guidelines for the effective management of shift-work disorder, orga-
nizational level changes are important. Three types of intervention have been rec-
ommended: (a) switching from slow to fast rotation; (b) changing from backward to
forward rotation; (c) self-scheduling of shifts [6]. There is evidence that a rapidly
rotating schedule is less detrimental as it minimizes the time spent in a desynchro-
nized state. Clockwise rotation, rather than counterclockwise rotation, was reported
to be preferred by workers, probably because the body clock period is somewhat
Chronobiological Treatment of a Desynchronized ANS 331

Table 8.3  Estimated time of day for exposure to natural or artificial bright light (>1000 lux,
30 min) to blunt the circadian system of an air traveler in long (>8-h) time shifts
Number of time zones
crossed Eastbound Westbound
8 0600–0900 h and 1100–1400 h 1400–1700 h and 1900–2200 h
9 0700–1000 h and 1200–1500 h 1300–1600 h and 1800–2100 h
10 0800–1100 h and 1300–1600 h 1200–1500 h and 1700–2000 h
11 0900–1200 h and 1400–1700 h 1100–1400 h and 1600–1900 h
12–13 1000–1300 h and 1500–1800 h 1000–1300 h and 1500–1800 h

18:00h 00:00h 06:00h 08:00h 18:00h

Sleep
a S
Work Sleep

b L Sleep Minimum in
central body
L Sleep
temperature
L Sleep
L Sleep
Melatonin
L Sleep (3-6 mg)

Fig. 8.11  The common situation of a night worker working in a 5-day night shift is depicted. The
minimum central body temperature must be shifted toward daytime sleep to ensure a recovery
sleep at home. (a) Light and activity at night cause a phase delay that is counteracted by sunlight
on return home. Although light and physical activity during the night work shift cause some phase
delay, light on the commute home opposes this process (causing phase advance). (b) Periods of
bright light progressively longer in the first part of the night shift give a phase delay that is sus-
tained by using dark glasses on the return home and by taking melatonin before sleep at home
(melatonin treatment before daytime sleep is useful for causing sleepiness and inducing phase
shifting). Planned 30-min napping just before or on the job combined with caffeinated drinks
reduce sleepiness and improve alertness while working [43]

longer than 24 h. Longer duty shifts allow more time off work. It is possible that a
flexible approach is best because of major individual differences in workers. The
goal is to achieve at least 7 h of sleep per 24 h. Melatonin treatment before daytime
sleep is used to promote sleepiness and to induce phase shifting.
Figure 8.11 summarizes the common situation of a night worker working a 5-day
night shift [43]. The minimal central body temperature must be shifted toward day-
time sleep by using bright light during the first part of the work period and wearing
dark goggles on the commute home. Melatonin (3–6 mg) before daytime sleep
causes sleepiness and phase shifting. Planned 30-min napping just before or on the
job combined with caffeinated drinks reduce sleepiness and improve alertness while
working. The wakefulness-promoting agents armodafinil and modafinil have been
approved by the Food and Drugs Administration (FDA) for the treatment of exces-
sive sleepiness in patients with shift-work disorder.
332 8  Clinical Implications of the Timed Autonomic Nervous System

In conclusion, both jet-lag disorder and shift-work disorder share a similar cause
and management has major similarities. In both cases, three factors are important:
(a) sleep scheduling; (b) resetting the body clock with light and/or chronobiotics;
(c) use of drugs to promote wakefulness if needed.
For the treatment of clinical chronodisruption it is important to note that the pri-
mary alteration may have originated elsewhere, and its effects on the circadian sys-
tem are one of the many changes it causes. In other words, a chronobiological
treatment could be effective because it fights some of the symptoms of the disorder
rather than the disorder itself. These warnings are of little importance to those who
are limited to treating patients, but they pose a problem of interpretation for those
who wish to understand the substance of the problem and try to develop a more
rational treatment [14–17]. The combination of the specific and chronobiological
treatment is needed for full recovery (Figs. 8.7 and 8.8).
Subjects suffering from insomnia of various causes usually take melatonin, admin-
istered 1 or 2 h before the time when it is desired to sleep. In addition to insomnia
associated with jet lag and shift work, melatonin is also effective in insomnia of the
(otherwise healthy) elderly, in patients with senile dementia, in blind subjects, in sub-
jects who can see but present free-course rhythms, and in patients with delayed-phase
sleep syndrome. Its efficacy is demonstrated in both objective and subjective sleep
time estimates and in the objective measurement of actimetry. In a meta-analysis
including 19 studies and involving 1683 subjects, melatonin showed significant effi-
cacy in reducing sleep latency and increased total sleep time [44]. Trials of longer
duration and the use of higher doses of melatonin demonstrated greater effects.
Several consensus statements encourage the use of melatonin to treat insomnia [45].
On average, humans cannot do well without sleep for more than a few days
(about 2 or 3 days). With only 24–48 h of sleep deprivation, failure of short-term
memory appears, there is an increase in the feeling of fatigue, sleepiness, and
aggression, and a depressed mood. After 72–98 h without sleep, fatigue is severe
and episodes of mental confusion and distortion may occur; in certain individuals,
sleep deprivation may produce persecutory delusions.
To solve sleep deprivation, there is no choice but to sleep. In general, after sleep
deprivation, we recover one third of the total lost sleep time, 100% of slow-wave sleep
and 30–50% of REM sleep [46]. Therefore, the estimated 10 h of sleep deficit accumu-
lated during the week if we sleep 2 h less than needed daily can be recovered by sleeping
1.5 h more on Saturday and Sunday. Hence, the extraordinary importance of not using
strict schedules for sleep at the weekend and let it flow without using an alarm clock.

 ome Clinical Autonomic Entities Associated


S
with a Desynchronized ANS

Metabolic Syndrome

In the present world, food has become abundant and simultaneously, the need for
physical effort has been greatly reduced. From an evolutionary perspective, this is
another “environmental mutation,” that, together which the advent of artificial
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 333

lighting, has greatly contributed to the loss of the experience of a differentiated day
and night. Chronodisruption, with a disturbed balance among the three different
configurations of organ and system regulation in a 24-h cycle, is a direct conse-
quence of this (Fig. 8.2).
Obesity has a profound impact on health, such as T2DM and hypercholesterol-
emia, mainly through its influence on secretion and insulin sensitivity. The metabolic
syndrome comprises a group of metabolic abnormalities (hyperinsulinemia, insulin
resistance, hypertension, obesity, hyperlipoproteinemia, hypertriglyceridemia) that
increase the risk of cardiovascular disease and T2DM. The metabolic syndrome is
also associated with an increased risk of nonalcoholic fatty liver disease and renal
dysfunction. Similarly, there is evidence for the correlation of metabolic syndrome
with dementia (“type 3 diabetes mellitus”) and with cancers, mainly of the breast,
pancreas, and bladder [47].
There is impressive information indicating that the obesity is associated with
low-grade inflammation of the white adipose tissue, which can subsequently lead to

NORMAL ADIPOSE TISSIE

Excessive food intake


Expansion of adipose tissue
Macrophage infiltration

OBESE ADIPOSE TISSUE


 Food intake
Energy expenditure

 Hepatic glucose production
 Inflammation
 Oxidation of lipids
Altered secretion of adipocytokines from:  Steatosis
ADIPOCYTES MACROPHAGES  Insulin resistance

Adiponectin  Resistin  IL-6  Glycogenolysis
 Leptin  FFA  TNFa  Gluconeogenesis
 Glucose output
 Endothelial dysfunction  VLDL production
Endothelium-dependent vasodilation

 Vascular adhesiveness
 Intimal hyperplasia
 Atherosclerosis
 Glycemia Dyslipemia
 Insulin secretion (early stage)
  Inflammation
b cell function

Oxidation
 b-cell failure  Insulin resistance
Insulin secretion Glucose uptake
 

INSULIN TYPE 2 CARDIOVASCULAR TYPE 3


RESISTANCE DIABETES DISORDERS DIABETES

Fig. 8.12  The metabolic syndrome is the consequence of obesity-induced changes in adipokine
secretion that lead to the development of systemic insulin resistance, T2DM, type 3 diabetes, and
cardiovascular disorders. Overnutrition that results from a combination of increased food intake
and reduced energy expenditure leads to adipose tissue expansion, increased adipocyte size and
number, and increased macrophage infiltration that, together, lead to increased free fatty acid
release, dysregulated secretion from adipocytes of a variety of adipocytokines, including adiponec-
tin, leptin, and resistin, and increased release from resident macrophages of the inflammatory
cytokines (TNF-α, IL-6). Dysregulated secretion of these adipokines elicits a variety of adverse
effects on numerous tissues and leads to the development of systemic insulin resistance that
increases the risk for development of the metabolic syndrome, a variety of cardiovascular disor-
ders, and T2DM and type 3 diabetes (when combined with dementia). Reproduced with permis-
sion from Cardinali [75]
334 8  Clinical Implications of the Timed Autonomic Nervous System

insulin resistance, impaired glucose tolerance, T2DM, and type 3 diabetes [48, 49].
Adipocytes actively secrete proinflammatory cytokines such as TNF-α, IL-1β, and
IL-6 and trigger a vicious circle that leads to additional weight gain, largely as fat
(Fig. 8.12). Increased circulating levels of C-reactive protein and other inflamma-
tory biomarkers also support the occurrence of inflammation in obesity.
The altered production of proinflammatory cytokines modulate adipocyte size
and number through paracrine mechanisms that exert an important role in the regu-
lation of fat mass (Fig. 8.12). The amounts of proinflammatory molecules derived
from adipose tissue in obese patients diminishes after weight loss [50]. Therefore,
the fat cells are both a source and a target for TNF-α, IL-1β, and IL-6 (Fig. 8.12).
A number of factors play an unequivocal role in increasing the risk of developing
T2DM [51]. These factors include dysfunction of pancreatic β cells, abnormal adi-
pogenesis and absence of adequate insulin responsiveness in the liver, genetic sus-
ceptibility, physical inactivity, excessive food consumption and/or high calorie food
intake, and a sedentary life-style.
In addition to these predisposing factors, there is now an increasing amount of
evidence that disruption of circadian timing mechanisms is a major contributor to
the development of T2DM [49]. Various types of circadian disturbances have been
correlated with T2DM, including disruption of the timing of bodily functions that
are normally synchronized, improper timing of food intake, dampened clock gene
expression and polymorphisms, sleep loss/disturbance and impairments of the mel-
atonin signaling pathway [52].
In a study on 593 patients with a recent diagnosis of T2DM, sleep debt resulted
in long-term metabolic disruption, which may promote the progression of the dis-
ease [53]. Sleep quality rather than sleep duration played an important role in insu-
lin resistance in these newly diagnosed T2DM patients [54]. The increased obesity
rate recorded in 100,000 women of the Breakthrough Generations Study, associated
with increased levels of light at night exposure, supports this assumption [55].
Glucose metabolism is among the numerous physiological functions that are
governed by the circadian apparatus [56]. It is known that the distribution of glucose
to all parts of the body is organized by the molecular clock present in liver. Circadian
regulatory mechanisms utilize both neural and humoral communication to exert
close control over insulin, leptin, and plasma glucose levels (Chap. 2).
Direct evidence for the association between circadian clock disruptions and
T2DM has been provided by studies in mice in which linkages between clock gene
mutations and diabetes states were examined [57]. Per2 mutant mice show several
abnormal profiles, including the absence of rhythmicity in plasma glucocorticoid
levels, obesity, and low levels of neuropeptides involved in appetite regulation.
Clock gene mutations promote delays in pancreatic gene expression, which in turn
affects the regulation of the growth, development, and survival of insulin cells and
thus of the glucose signaling pathway [58]. Clock-mutant mice showed a lack of
rhythmicity in the action of insulin, a state that was reversible once the CLOCK
protein had been reintroduced [59].
Additional confirmatory evidence for circadian control over metabolic activity
was provided by a study of mice in which the Bmal1 gene was specifically knocked
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 335

in the pancreas. This produced an animal model that replicated T2DM in that blood
glucose levels were found to be elevated throughout the 24-h cycle [58].
In diabetes-prone, genetically engineered rats that were maintained in continu-
ous light and jet-lag-like conditions, circadian disruption ensues [60]. Similar find-
ings were obtained in a study of humans subjected to a forced desynchronization
protocol for a period of 28 h. The treatment produced circadian misalignment in all
subjects, with 30% of individuals exhibiting a disturbed glucose metabolism that
resembled diabetes [61]. Also, all subjects showed a decreased concentration of
leptin, a reversal of cortisol rhythm, and high amounts of glucose (postprandial)
even in the presence of increased insulin. Additionally, the increased levels of cor-
tisol late in the day, i.e., during the end of wakefulness, were a potentiating factor
for insulin resistance and hyperglycemia [59]. These findings suggest that the mis-
alignment of clock functions accelerates the development of T2DM (Fig. 8.13).
Light exposure at night, even at low levels, has been reported to alter food timing
and body mass accumulation, thus suggesting that artificial lighting is an important
contributing factor to the increased prevalence of metabolic disorders [56]. As dis-
cussed in Chap. 5, circadian rhythms could be entrained by manipulating the timing

Disruption of Sleep /Wake Cycle

Sleep / wake Sleep / wake


cycle cycle

12 12
1 11 1
11
2 10 2
10
9 3 9 3

4 8 4
8
5 7 5
7 6 6

12 12 12 12
1 11 1 1 11 1
11 11
2 10 2 2 10 2
10 10
9 3 9 3 9 3 9 3

4 8 4 4 8 4
8 8
5 7 5 5 7 5
7 6 6 7 6 6

Feeding / fast Metabolic Feeding / fast Metabolic


rhythms rhythms rhythms rhythms

Metabolic Dysfunction, Obesity

Fig. 8.13  As a disruption of the timing of bodily functions that are normally synchronized such as
the sleep/wake cycle, improper timing of food intake, metabolism, and dampened clock gene
expression occur. Metabolic dysfunction leads to the metabolic syndrome. Reproduced with per-
mission from Cardinali [75]
336 8  Clinical Implications of the Timed Autonomic Nervous System

of food administration, and that this could be accomplished without the participa-
tion of the SCN (Fig. 5.33). Taken together, the data indicate that the administration
of food at inappropriate times may disrupt the metabolic profile, thus creating a
desynchronized physiological state that is causally linked to the development of
T2DM (Fig. 8.13).
The possible association between clock gene polymorphisms and T2DM has
been explored in humans. In one study a polymorphic allele was identified in Cry2
that correlated with T2DM [62]. Two Bmal1 variants have been found to be associ-
ated with diabetes and hypertension in a British population [63]. In a study to exam-
ine the existence of Per3 variants in patients with T2DM we reported that, compared
with the group without diabetes, the frequency of the occurrence of the five repeat
alleles of Per3 among affected patients was greater, and that of the four repeat
alleles was less [64]. Circadian clock variants have also been found to correlate with
body mass index [65], and with weight loss, sleep duration, and total plasma choles-
terol in obese Caucasian individuals [66]. Clock and Cry1 polymorphisms are
involved in individual susceptibility to abdominal obesity in a Chinese Han popula-
tion [67]. Single nucleotide polymorphisms in Bmal2 gene have been associated
with a high risk of developing T2DM in obese patients [68]. Cross-sectional studies
have reported associations between the clock gene polymorphisms and the preva-
lence of obesity, plasma glucose levels, hypertension, and T2DM. The association
of the Clock polymorphism and stroke in T2DM was reported, indicating that core
clock genes significantly contribute to increased cardiovascular risk in T2DM [69].
A high-fat diet that contributes to insulin resistance, impaired glucose metabo-
lism, and obesity, can feedback to influence the biological clock. Rats on a 35% fat
diet exhibited a disrupted 24-h rhythmicity of Per1, Per2, Cry1, and Cry2 expres-
sion, the Per2 expression profile being almost inverted by the high-fat diet (Fig. 8.14)
[70]. These results indicate that the inherent transcription, translation, and post-­
translational modifications that give the clock its own natural rhythmicity are dis-
rupted in obese rats. The current evidence thus suggests that these processes might
have a reciprocal relationship, as an abnormal functioning of metabolism promotes
altered expression of the clock genes, whereas impaired clock functioning can dis-
rupt metabolic activity. Indeed, in T2DM patients, clock gene expression was
directly associated with fasting glucose levels and with insulin mRNA and protein
concentration [71].
In a recent survey, samples of subcutaneous adipose tissue from 50 overweight
subjects were collected before and after an 8-week administration of a hypocaloric
diet. The expression of core clock genes, Per2 and NR1D1, increased after the
weight loss, and their levels correlated with the expression of several genes involved
in fat metabolism [72]. Another study demonstrated that a weight loss intervention
based on an energy-controlled Mediterranean diet influences the methylation levels
of three clock genes, Bmal1, Clock, and NR1D1, being an association between the
methylation levels and the diet-induced serum lipid profile [73]. Thus, an abnormal
functioning of metabolism promotes the low expression of clock genes, whereas
impaired clock functioning can disrupt metabolic activity. Figure 8.15 summarizes
the factors and their output that may be responsible for T2DM. The findings thus
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 337

4% fat diet
3,5 5 6
35% fat diet

Relative expression of Bmal1

Relative expression of Pert1


Relative expression of Clock

3,0 4 5

2,5
3 4
**
2,0

1,5
2 3
**
1 2
1,0

0,5 0 1 *
0,0 0
9 13 17 21 1 5 9 13 17 21 1 5 9 13 17 21 1 5
2,0 3,5 3,0
Relative expression of Cry1
Relative expression of Per2

Relative expression of Cry2


3,0 2,5
**
1,5
2,5
2,0 **
2,0
1,0 1,5
1,5
1,0
1,0
0,5
*
** ** * 0,5 0,5

0,0 0,0
** 0,0
9 13 17 21 1 5 9 13 17 21 1 5 9 13 17 21 1 5
Colck time

Fig. 8.14  Circadian oscillating expression of clock genes is sensitive to a high fat diet. In this
experiment, rats were fed a 4% (control) or a 35% fat diet for 11 weeks. Wistar male rats (n = 6–8
per group) were killed by decapitation at six different time intervals throughout a 24-h cycle.
Adenohypophysis was collected and RNA extracted. The means ± SEM of mRNA expression by
real-time PCR (**p < 0.01, *p < 0.05 compared with control rats in a Student’s t test). Results from
Cardinali et al. [70]. Reproduced with permission from Cardinali [75]

Impaired glucose Shift work/ Obesity Food


metabolism Jet lag
Impaired clock gene Hyperglycemia Clock gene
expression variation SCN lesion

Sleep disturbance Low leptin


level Clock regulation of
Rest phase neuropeptides
Regulation of insulin, leptin & Abnormal feeding regulating appetite
glucose level appetite
High fat diet leads to
High Food timing
Timing of food impaired locomotor
cortisol activity
T2DM
Insufficient
Noctumal light suppresses Altered Body mass & composition
evironmental
melatonin
Adipogenesis, adipocytes cues
Impaired melatonin differentiation
pathway Glucose
Pancreatic cell growth, development, metabolism
Low melatonin proliferation & survival
concerntration
Irregular sleep/wake
Circadian variation in Absence of feeding rhythm schedule
insulin response
Glucocorticoids levels

Insulin sensitivity Clock gene mutation Sleep/wake disruption

Fig. 8.15  Adapted fishbone diagram including the factors and their output responsible for type 2
diabetes mellitus
338 8  Clinical Implications of the Timed Autonomic Nervous System

Dyssynchrony of hypothalamic clocks: ALTERED APPETITE

Dyss
Dyssynchrony of clock
genes in immune cells:
Dyssynchrony of cardiac INF
INFLAMMAGING
clocks:
CARDIOVASCULAR Dyssynchrony of clock genes
Dyssync
DISEASES in p
pancreatic islets:
INSUL DEFFICIENCY
INSULIN

Dyssynchrony
Dyss of clock
genes
gen in adipocytes:
Dyssynchrony of clock genes
es OBESITY,
OBES ENDOCRINE
in hepatocytes: DISEASES
INSULIN RESISTANCE,,
STEATOSIS

Dyssyn
Dyssynchrony of clock
genes in muscle cells:
INSULIN RESISTANCE

Fig. 8.16  A hypothetical distribution of circadian alterations in organs and tissues in the meta-
bolic syndrome. Data are extrapolated from studies in transgenic mice that have provided evidence
of the association between circadian clock disruptions and metabolic and behavioral events in
metabolic syndrome. Reproduced with permission from Cardinali [75]

underscore the importance of circadian regulation for normal metabolic function-


ing, and further, that clock gene expression appears to be disturbed in T2DM
(Fig. 8.16).
The possibility that a relationship might exist between melatonin and T2DM is
supported by findings that insulin secretion is inversely proportional to plasma mel-
atonin concentration. It is known that melatonin can alter insulin function and that,
compared with normal healthy individuals, T2DM patients have lower circulating
levels of melatonin. In vitro co-incubation of pancreatic cells with melatonin has
been shown to inhibit the glucose-mediated release of insulin, additionally support-
ing the conclusion that melatonin activity plays a role in the function of insulin.
Further, circadian disruption has been shown to alter melatonin secretion and to
cause dysfunctions in pancreatic β cells. Genetic association studies that have shown
that mutations in the melatonin receptor gene correlate with an increased suscepti-
bility to T2DM [74].
It can be postulated that the alteration of phase, amplitude, and synthesis of mel-
atonin could lead to impaired glucose metabolism in circadian disrupted individu-
als. Suppression of melatonin secretion by nocturnal light exposure could be a
crucial factor for T2DM development [56]. In this respect, the successful manage-
ment of T2DM may require an ideal drug that besides antagonizing the triggers of
T2DM, also corrects the disturbed sleep–wake rhythm.
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 339

O (normalizes)
NH
Melatonin Hypertension
CH3 (augments)

H3C O HDL cholesterol


FFA SNS
IL-6
N TG (improves)
H Small dense LDL Insulin

VLDL (decreases) Glucose


(decreases)
TNF-α
Insulin (improves)
CRP IL-6
Glycogen

(restores)

11 12 1 FFA CO2
10 2 FFA
9 3
8 4 Fibrinogen
7 6 5
Adiponectin
PAI-1 IL4
Prothrombotic Triglyceride
IL10
state (increases) (intramuscular droplet)

Fig. 8.17  Effects of melatonin in the metabolic syndrome. Melatonin normalizes high blood pres-
sure and circulating indexes of inflammation. It improves insulin sensitivity and restores disrupted
circadian rhythms. Reproduced with permission from Cardinali [75]

Melatonin, together with morning light, is an interesting chronotherapeutic option


that can reset the phase and amplitude of circadian rhythms in T2DM patients. The
combination of the specific and chronobiological treatment is needed for full recov-
ery (Figs. 8.7 and 8.8). As has been shown in animal models of diabetes and obesity,
melatonin also possesses cytoprotective properties that may prevent a number of
unwanted effects in T2DM [75]. The efficacy of melatonin treatment for treating
metabolic and cardiovascular comorbidities in T2DM has been reported (Fig. 8.17)
[76–81]. However, melatonin treatment was reported to impair insulin release in
humans who carry MTNR1B, the risk allele of the MT2 receptor gene [82]. To what
extent insulin resistance is ameliorated by melatonin in these patients remains to be
defined. At an early stage of T2DM treatment, nonpharmacological approaches such
as lifestyle modification, low-fat diet, and exercise are recommended [52].

Mental Illnesses

The human species is very vulnerable to psychiatric diseases: 1% of the population


suffers from schizophrenia, 4% from severe uni- or bipolar depression, and 15% of
some form of reactive emotional illness in isolated episodes. There are epidemio-
logical elements to suspect a genetic predisposition in the appearance of several of
these diseases.
Emil Kraepelin, at the beginning of the twentieth century, was the first to
develop a modern classification of mental illness, differentiating “precocious
340 8  Clinical Implications of the Timed Autonomic Nervous System

dementia” (schizophrenia) from the more benign and recurrent forms of illness
(“manic disease”). Following Kraepelin, cognitive alterations (“disorders of
thought” such as schizophrenia) and affective disorders (“emotional disorders” or
“affective diseases” such as depression, manic depression, and anxiety) have been
identified [83].
Emotional illnesses are characterized by an abnormality of emotional and expres-
sive experiences that constitute affectivity. Normally, a person’s affectivity is in
neutrality, with mild episodes of euphoria (called “happiness”) and mild episodes of
sadness (called “unhappiness”). When one trespasses these limits, one enters emo-
tional pathology. The pathological end of euphoria is called “mania”; The patho-
logical end of unhappiness is called “depression.” When individuals express both
abnormal ends in a single episode or at different times throughout their life, they are
diagnosed with bipolar disease type 1 or 2 (depending on the intensity of the manic
phase).
Mania is defined by the presence of an abnormally expansive and irritable
affectivity, coupled with several related symptoms. They include decreased need
for sleep, excessive loquacity, changing thoughts, grandiosity, easy distraction,
and increased activities with pleasurable objectives. This has important conse-
quences: non-normal sexual behavior, risky economic ventures, etc. A maniac
patient is often enthusiastic and positive, but tolerates little frustration. Because
mania is not an unpleasant experience, patients rarely admit to being treated.
However, relatives or close friends recognize this situation as dangerous for the
patient. Ultimately, the persistence of the patient in the manic state may be accom-
panied by serious affective or economic breakdown for the patients and their
families.
At the opposite end of the mania is depression. The depressed patient presents a
variety of disorders, including sleep disturbances, greatly diminished interest or
pleasure, changes in appetite and body weight, psychomotor agitation or inhibition,
decreased energy, guilt, impairment, and recurrent thoughts of death or suicide.
Approximately 15% of depressed patients complete a suicide attempt. It is of inter-
est that these alterations are linked to a high degree of creativity. For example, Van
Gogh or Tolstoy were famous depressive patients.
There is a strong familial component in these diseases. The risk of becoming ill
is twice that of the general population in individuals with a sick first-degree relative.
However, this evidence does not necessarily indicate a genetic transmission, as edu-
cational and social factors may also explain the picture.
The link between psychiatric pathological conditions and the ANS and with
brainstem monoaminergic neurons discussed in Chap. 4 is indicated by the thera-
peutic efficacy of various drugs that interfere with these neuronal mechanisms. It is
not known whether the modification of these diffuse monoaminergic systems is the
cause or effect of the psychiatric alteration that originates in other zones of the ANS
(e.g., the limbic system). The introduction of methodologies such as fMRI, PET or
magnetoencephalography has allowed monitoring of alterations in neuronal activ-
ity, whose correction is accompanied by improvement of psychiatric symptoms, to
be systematized in different brain areas. The symptoms of mental illness, such as
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 341

hearing voices in solitude (hallucination) or feeling persecuted (delirium) are the


result of alterations in brain circuits.
Mental illnesses range from relatively mild to severe forms. The mildest of these
are personality alterations (e.g., obsessive–compulsive disorders, psychopathic per-
sonality). Although seen as primarily “psychological,” these alterations bring about
changes in brain activity, as revealed by PET or fMRI, such as increased blood
perfusion of the prefrontal cortex and basal ganglia. The psychopathic personality
with antisocial behavior has a genetic component.
There is little controversy about the organic substrate of more severe mental ill-
nesses, such as schizophrenia, manic–depressive (bipolar) illness, major depression,
or anxiety disorders such as panic attacks or post-traumatic stress disorder. Circadian
rhythm abnormalities, as shown by sleep/wake cycle disturbances, constitute one
the most prevalent signs of mental illness. A substantial proportion of patients with
a sleep complaint have a psychiatric illness and a significant number of patients
with a psychiatric illness have a sleep complaint [83].

Mood Disorders
Mood disorders constitute a family of complex multifactorial illnesses that are
characterized by disruptions of several physiological, neuroendocrine, and behav-
ioral processes. According to the World Health Organization (WHO) reports, these
disorders are the fourth leading cause of global burden of disease and by the year
2020, they are expected to be the second highest cause of morbidity. An interpreta-
tion of the public health impact by prevalence statistics, however, must take the
following two factors into consideration. First, the use of antidepressants has
increased considerably over the last 15 years, and concurrently the prevalence of
mood disorders is also increasing. Second, despite this fact, only one third of
patients with mood disorders are treated effectively with medication. This would
suggest that either some unrecognized subtypes of mood disorders may exist that
are resistant to current pharmacological treatments, or that present conceptualiza-
tions of the underlying causes of mood disorders need to be reconsidered. There is
a growing amount of evidence that supports the latter suggestion and the accumu-
lating evidence points to the possibility that mood disorders may be an overt symp-
tom of what is basically a circadian rhythm disorder (Fig. 8.18). An internal
desynchronization of circadian oscillators with a strong oscillator being linked to
phase advances was postulated as a hypothesis for various subtypes of affective
disorders.
Circadian rhythm abnormalities, as shown by sleep/wake cycle disturbances,
constitute one the most prevalent signs of mood disorders, advances or delays in the
circadian phase being documented in patients with major depressive disorder, bipo-
lar disorder or seasonal affective disorder [84]. Changes in the sleep–wake cycle
structure in mood disorders often precede changes in a patient’s ongoing clinical
state. During a depressive episode, approximately 80% of patients complain of
symptoms of insomnia (frequent awakenings, early morning awakening) and 20%
complain of hypersomnia. Changes in sleep during a depressive episode are a long
sleep latency, reduced sleep efficiency, reduced stage N3, a short REM latency, a
342 8  Clinical Implications of the Timed Autonomic Nervous System

Genetic abnormality

Circadian rhythm abnormality Disrupted sleep

Altered neurotransmission

Mood disorder

Fig. 8.18  Bidirectional relationships among circadian rhythms, sleep, and mood. It is proposed
that disturbed circadian rhythm regulation has an impact on sleep rhythms to produce changes in
the monoamine regulation of mood. The altered mood can then influence sleep. Reproduced with
permission from Cardinali [75]

longer first REM period, and a higher REM density in the early part of the night
(Chap. 2). Early morning awakening may also be present.
Insomnia can precede a major depressive episode and is often the last symptom
of depression to resolve. The persistence of changes in sleep–wake cycle is a risk
factor for a relapse of depression. During manic episodes, the patient reports a
decreased need for sleep (feeling rested on a few hours of sleep). In turn, sleep loss
can precipitate a manic episode.
Individuals suffering from mood disorders often have circadian misalignment
of many physiological phenomena in addition to the sleep–wake cycle, e.g., BP,
neurotransmitters, mood states, body temperature, energy balance, appetite regula-
tion, melatonin secretion, and the levels of cortisol [85]. Variations in the sleep–
wake schedule preferences and poor activity patterns have also been observed to
occur more frequently than in healthy control subjects. Reduced sleep disturbances
have been reported in morningness phenotypes, thus suggesting the existence of a
correlation between activity pattern preference and mood regulation. Favoring a
circadian rhythm hypothesis of the disease, sleep deprivation and light therapy
have clinically relevant antidepressant effects in patients [86]. The combination of
the specific and chronobiological treatment is usually needed for full recovery
(Figs. 8.7 and 8.8).
Changes in the sleep–wake cycle structure in mood disorders often precede
changes in a patient’s ongoing clinical state and can even signal a relapse or predict
the occurrence of suicidal behavior. In addition to an altered sleep–wake cycle,
daytime mood variation and periodic recurrences are clinical findings that relate
depressive states to the circadian system [87]. A significant proportion of patients
have regular changes in the intensity of their depressive mood during the day, with
parallel changes in anxiety symptoms, attention capacity, and psychomotor symp-
toms that frequently accompany depression. Depressive patients with melancholic
characteristics typically have an early morning awakening, worsening their mood
state, which additionally correlates with elevations in cortisol secretion. Both
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 343

symptoms are part of the clinical diagnostic criteria of the melancholic depressive
subtype [87].
We have discussed in Chap. 2 how circadian processes interact with homeostatic
mechanisms to regulate 24-h oscillations in sleep propensity. Intensity of light and
its duration ensure the proper functioning of the circadian clock. Hence, the strength
of the entrained circadian pacemaker greatly influences the sleep, cognitive, and
emotional functioning. Additionally, the interactive dialogue between circadian sys-
tem and homeostatic mechanism affects the timing of sleep onset and wake onset in
individuals (Chap. 2).
It has been proposed that sleep and mood have a bi-directional and interactive
relationship (“vicious circle”) in which disturbances in sleep perpetuate distur-
bances in mood throughout the day (Fig. 8.19) [88]. Various studies have shown that
prolonged sleep disruption results in abnormal mood states. Conversely, alterations
in mood can affect sleep: it has been shown that when patients with bipolar disease
switch from depression to mania, they experience one or more successive rest–
activity cycles [89]. Sleep profiles of both manic and depressive patients are similar
in showing continuous sleep disturbance and more time spent in N1 sleep. Reduction
in the tendency to sleep and the quality of sleep, disturbed and increased REM
sleep, early morning awakening, and short sleep duration are observed in patients
with bipolar disease. Supersensitivity to light was proposed as a trait marker in

MOOD DISORDERS
Clock gene polymorphisms
Daily rhythm in depression
Depressive
patients have in depression Mutants of clock
sleep Phototherapy in genes exhibit
disturbances Changes in depression
seasonal affective
circadian
disorder and major
rhythms in
depression
Changes in depression
sleep predict
depression Mutants of clock genes
show sleep disturbances
CIRCADIAN
SLEEP GENES
CIRCADIAN RHTYHMS
Clock genes
Sleep deprivation has controlling
antidepressant effects circadian
rhythms
Some
antidepressants Some
Most
have circadian antidepressants
antidepressants
effects have effects on
affect sleep
circadian genes

ANTIDEPRESSANTS

Fig. 8.19  The circadian clock influences multiple systems and pathways that are thought to
underlie mood disorders. In most cases, there are reciprocal interactions that, in turn, regulate cir-
cadian rhythms. Circadian gene mutations may make an individual more vulnerable to mood
changes exacerbated by environmental deviations in the daily schedule. Reproduced with permis-
sion from Cardinali [75]
344 8  Clinical Implications of the Timed Autonomic Nervous System

bipolar disease patients after showing that melatonin levels in these patients fell
twice as much as the levels of normal subjects following exposure to light during the
night [90]. Phase advances in the rhythm of melatonin secretion have been docu-
mented in numerous studies of patients with major depressive disorder [91].
Bipolar patients suffer from abnormal circadian rhythms in their sleep–wake
cycles, with regard to BP, hormone levels, neurotransmitters, mood states, body
temperature, energy balance, appetite regulation, melatonin secretion, and the levels
of cortisol [92]. Variations in sleep–wake schedule preferences and poor activity
patterns have also been observed more frequently in bipolar disease patients than in
healthy control subjects. Reduced sleep disturbances were reported in morningness
phenotypes and this suggests a correlation between activity pattern preference and
mood regulation.
Globally, places receiving a reduced amount of sunlight over the year have a
higher incidence of depression. This is of importance considering that bipolar dis-
ease patients are more sensitive to light. Indeed, winter increases the chance of
depressive episodes, whereas the summer period promotes manic states in bipolar
individuals [92]. The loss of strength in coupling the circadian oscillator and envi-
ronmental cues contributes to the deterioration of mood. Hence, perturbations in the
environmental cues such as irregular sleep–wake cycle and social schedule, receiv-
ing a limited amount of sunlight and exposure to artificial light at unusual times,
travelling between two different photoperiods or shift work are the factors respon-
sible for the misalignment of the clock that may contribute to mood disorders.
A hypothesis has been postulated for bipolar disease, interlinking sleep and
wakefulness in such a way that disturbed mood states during the daytime can affect
the sleep process. Similarly, sleep deprivation during nighttime influences daytime
mood [93]. Dysregulation of neurotransmitters is one of the major manifestations of
mood disorders. An interesting hypothesis is that a high level of DA induces mania
and that low levels are associated with a depressive episode of bipolar disease.
Many medications prescribed to treat patients with mood disorder influence 5-HT,
DA, and NE levels and abnormal emotional states could be averted by modulating
the neural communication of monoamines.
The social zeitgeber theory is one of the theories proposed to explain the origin
of bipolar disorder and is based on the principles of biological rhythms. It states that
stressful life events influence the regular sleep–wake cycle, resulting in the disrup-
tion of the circadian clock, inducing mood disorder in vulnerable people. In addi-
tion, social zeitgebers possess a critical function in the regulation of normal mood
states, and social rhythm therapy is effective for mood regulation in bipolar disease
patients.
The most commonly used drug to treat maniac episodes in bipolar illness is
lithium. It acts on numerous steps in intracellular signaling and generally length-
ens free running rhythms [94]. Present-day psychiatric treatment of chronobio-
logical aspects of mood disorder comprises medication and nonpharmacological
treatments, including phase advance, light therapy, and sleep deprivation. The
combination of specific and chronobiological treatments is often employed
(Figs. 8.7 and 8.8).
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 345

The timing of melatonin secretion is important for the regulation of mood. Phase
advances of at least 1 h were observed in the nocturnal melatonin peak during the
manic phase of bipolar disease, whereas phase delays in circulating melatonin
occurred in bipolar disease type 1 patients [92]. Patients suffering from seasonal
affective disorder exhibit delayed circadian rhythms, with a delayed offset of mela-
tonin secretion of about 2 h. Presumably, the symptoms of hypersomnia and late
awakening seen in seasonal affective disorder are due to the delayed phase and long
duration of melatonin secretion [92].
Elucidating how biological clocks regulate the timing of physiological processes
could ultimately be helpful for understanding and treating mood disorders, which
may be the result of dampened circadian oscillators. Some initial efforts have
already been made to use sleep scheduling as one component of psychiatric therapy.
The benefit of this approach has been demonstrated in studies showing that the close
maintenance of sleep–wake timing in patients with mood disorders can produce
enhancements in patients’ mood profiles. Additionally, extended bedrest and opti-
mal sleep duration have been shown to substantially reduce manic episodes in bipo-
lar disease patients. Psychiatric treatment programs are increasingly incorporating
recommendations that chronotherapies be adopted as adjunctive strategies for mood
disorders (Figs. 8.7 and 8.8) [95, 96].
Clinical studies involving chronobiological manipulations, such as exposure to
bright light in the morning and melatonin administration in the evening, have been
found to be useful for reducing phase abnormalities and depressive symptomatol-
ogy [97]. Indeed, this is the basis for postulating the association of an adjuvant
chronobiological treatment to the specific treatment of the psychiatric disease to
obtain full recovery of patients (Figs. 8.7 and 8.8). It must be noted that from clini-
cal standpoints the abnormalities in amplitude of the sleep–wake cycle (“poor sleep
together with poor vigilance”) is a paramount sign of the disease and that its correc-
tion increases substantially the quality of life of the patient suffering an affective
disorder, regardless of the uncontrolled influence of external (light/dark cycle) or
internal (sleep–wake cycle) masking phenomena.
We discussed in Chap. 2 the intricacies of the cellular mechanisms of circadian
oscillation. These clock genes influence mood disorders. Genes that can tolerate
environmental disruption are not associated with abnormal mood swings, whereas
mutations in circadian clock genes are related to them. Moreover, clock gene allelic
variations can worsen mood symptoms and contribute to differential effects of psy-
chiatric medications. Mutations of circadian genes in animal models resemble mood
alterations.
In humans, the relationship between mood regulation and clock genes was first
identified in seasonal affective disorder [98]. A positive association for Npas2 and
Per3 and sleep–wake time preference was found in patients. Genetic variants of
Bmal1, Npas2, and Per2 have been found to increase susceptibility toward the
development of seasonal affective disorder. Cry2 alleles and their mRNA levels
have been directly associated with a depressive mood profile and Clock alleles were
associated with hyperactivity in bipolar disease patients, who exhibited significant
delayed sleep phase and reduction in sleep duration [99]. Significant reductions in
346 8  Clinical Implications of the Timed Autonomic Nervous System

insomnia and antidepressant treatment have correlated with Clock variants in bipo-
lar disease and major depressive disorder patients. The post-mortem analysis of
major depressive disorder patients indicated a dampened clock gene expression in
several brain regions compared with healthy controls [100]. In a collaborative study
with Indian and Canadian colleagues [101], we found in a South Indian population
an increased prevalence of five repeat homozygotes of Per3 in bipolar disease
patients, which was particularly notable among female patients. No significant asso-
ciation was observed in the allele frequencies of four and five repeat alleles in
schizophrenia patients. Therefore, the occurrence of the five-repeat allele of Per3
can be a risk factor for type 1 bipolar disease onset.
Panic disorder is a complex anxiety disorder characterized by recurrent panic
attacks. It is a poorly understood psychiatric condition that is associated with sig-
nificant morbidity and an increased risk of suicide attempts and completed suicide.
Accumulating evidence suggests that ANS mechanisms controlling acidosis consti-
tute a contributing factor in the induction of panic [102]. Challenge studies in
patients reveal that panic attacks are provoked by agents that lead to acid–base
imbalance, such as CO2 inhalation or sodium lactate infusion. Chemosensory mech-
anisms that translate pH into panic-relevant fear, autonomic, and respiratory
responses include regions such as the subfornical organ and medullary raphe that
can directly detect pH fluctuations in the internal milieu. The hypothalamus, amyg-
dala, and periaqueductal gray, in addition to their chemosensory potential, also rep-
resent key nodes in the processing of external threats and sensory stimuli (Chaps. 4
and 5). Acidosis sensed by chemosensory mechanisms is translated to autonomic,
behavioral, and respiratory symptoms of a panic attack. The amygdala, periaque-
ductal grey and the hypothalamus regulate behavioral and autonomic symptoms of
panic, whereas respiratory symptoms are regulated by brainstem regions such as the
medullary raphe and the PBN via inputs from the hypothalamus and indirectly from
the subfornical organ through the OVLT [102]. Many of these structures connect via
thalamic nuclei with the insula, a region relevant for interoceptive sensing and
shown to be dysfunctional in panic disorder patients.
Patients with panic disorder are at an increased risk of myocardial infarction and
sudden death, which is evident in epidemiological surveys. Cardiac arrhythmias and
coronary artery spasm have been demonstrated during panic attacks. A constellation
of brain and sympathetic nervous system abnormalities, both in the quiescent phase
and during panic attacks, have been documented in panic disorder sufferers; these
are believed to underlie the increased cardiovascular risk. Abnormalities demon-
strable at rest are very high brain turnover of 5-HT, perhaps the CNS substrate for
panic disorder, E cotransmission in sympathetic nerves, and impairment of neuronal
reuptake of the NE released by sympathetic nerves. During panic attacks, there is a
surge of E secretion, accompanied by high-level activation of central sympathetic
outflow, including to the heart, and release of neuropeptide Y by cardiac sympa-
thetic nerves.
Melatonin has shown efficacy for treating mood disorders and the data indirectly
suggest that disruption in melatonin pathways could alter circadian clock mecha-
nisms, leading to the disruption of physiological processes, including sleep and
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 347

mood behavior. In addition to sleep promotion, MT1 and MT2 receptors appear to be
involved in sedative and antiexcitatory effects of melatonergic drugs. This has been
mainly studied in relation to anticonvulsant actions that have been linked to a facili-
tatory role of melatonin on GABA neurotransmission. In mammals, the antiexcit-
atory actions of melatonin may be also related to additional anxiolytic,
antihyperalgesic, and antinociceptive effects [75]. Because melatonin impairs con-
textual fear conditioning, a hippocampus-dependent task, and facilitates the extinc-
tion of conditional cued fear without affecting its acquisition or expression [103], it
may serve as an agent for the treatment of posttraumatic stress disorder. Clinical
evidence supports such a possibility [14].
To improve the efficacy of the sleep-promoting effects of melatonin, several ana-
logs of melatonin have been developed for treating circadian rhythm sleep disorders
or insomnia. Among these, agomelatine has been licensed by the European
Medicines Agency (EMA) for the treatment of major depressive disorder in adults.
Agomelatine has a unique pharmacological profile, as it is both a MT1/MT2 melato-
nin receptor agonist and an antagonist of 5-HT2C receptors [104].
For decades, the treatment of depression had revolved around drugs that increase
synaptic amounts of monoamine neurotransmitters (5-HT, NE, etc.). As the first
melatonergic antidepressant on the market, agomelatine displays a nonmonoami-
nergic mechanism that addresses sleep disturbances and depressive symptoms
together. Agomelatine has an early onset of action even in a severely depressed
population and may stand unique among antidepressants for effective management
of a major depressive disorder. In several studies, agomelatine has not only pro-
duced a cure of depressive symptoms, but also the patients return to a normal social
and occupational functioning [104, 105].
Increases in alanine aminotransferase and/or aspartate aminotransferase (three
times the upper limit of normal) have been noted in patients treated with agomela-
tine and this has become of great concern. In early 2015, the European health
authorities chose to keep agomelatine on the market despite its serious adverse
hepatic effects. It is recommended that liver function tests should be performed in
all patients: at initiation of treatment and then periodically after around 6 weeks,
after around 12 and 24 weeks, and thereafter when clinically indicated.

Schizophrenia
Schizophrenia is a complex psychiatric disorder comprising both positive and nega-
tive symptoms, including hallucinations, delusions, poor social behavior, and low
motivation [106]. The diagnosis of schizophrenia is currently based on the presence
of specific symptoms in affected individuals. About 1% of the global population
suffer from schizophrenia. Many genes have been identified to play a potential role
in the pathogenesis of schizophrenia, but none are established as causative. To date,
there has been no biological marker for schizophrenia.
Disruption of rest/activity cycles have been observed in schizophrenia patients
and may be generated by abnormal functioning of the circadian clock [106].
Schizophrenic, schizoaffective and bipolar patients have greater eveningness scores
than control subjects. Hence, there may be vulnerability for evening chronotype
348 8  Clinical Implications of the Timed Autonomic Nervous System

individuals to develop these disorders [107]. A recent study of 100,420 individuals


from the UK Biobank confirmed this relationship between schizophrenia and a eve-
ning chronotype [108].
Schizophrenic patients exhibit altered circadian rhythms of serum concentrations
of tryptophan, serotonin, prolactin, and cortisol. Aberrant rest/activity behaviors
have been documented in several studies of schizophrenic patients, including phase
shifts, abrupt and segmented sleep cycles, and waking during sleep [106]. The most
common sleep disturbances in schizophrenia are prolonged sleep onset latency and
problems of sleep maintenance [109]. Sleep disruption has been found to occur in
schizophrenic patients during NREM and REM sleep stages, although no signifi-
cant changes have been observed in REM sleep duration. As REM sleep is regulated
by the cholinergic system, an altered cholinergic activity may have a connection to
the development of hallucinations.
Proper entrainment by zeitgebers has been shown to stabilize affected schizo-
phrenic individuals [86]. Because of an abnormal activity pattern, schizophrenia
patients often have an unusual length of light exposure and this in turn, has an effect
on the patient’s sleep. Insufficient exposure and inappropriate timing of environ-
mental cues can lead to phase misalignments between internal and external timing
factors [110]. These in turn may have an impact on mental and neurological func-
tioning. A combination of weak zeitgebers, delayed sleep schedule, absence of
morning light, and receiving more light during the evening can generate a differen-
tial phase between internal and external time keepers. Thus, the circadian system
may play a significant role in schizophrenic symptoms. A combination of the spe-
cific and chronobiological treatment may be needed for full recovery from the crisis
(Figs. 8.7 and 8.8).
DA brain systems play a crucial role in schizophrenia pathology. Both the first-
and second-generation antipsychotic medications, the prime drugs used to treat
schizophrenia, block D2 receptors. They are particularly effective against the posi-
tive symptoms of the disorder. The term second generation is usually reserved for
agents modified to reduce side effects or to decrease negative symptoms [106].
Disruption of glutamate, GABA and orexin/hypocretin systems are also thought to
be involved in schizophrenic pathology [111].
Disruption of the melatonin profile in schizophrenia has been described in sev-
eral studies, often correlating with sleep–wake disruption [106, 112]. In some
instances, the circadian period of melatonin in schizophrenia patients was longer
than 24 h; thus, is phase shifted, in others it is phase advanced, and in others still
it may be arrhythmic. Schizophrenia patients often exhibit a unique difference in
the circadian phase angle, i.e., there exists a difference among the timing of mela-
tonin secretion, sleep–wake cycle, and the zeitgeber. Moreover, this phase angle
difference has been associated with the symptoms observed in schizophrenia.
Neurodegeneration is often found in schizophrenia [113]. As discussed later in this
chapter, melatonin acts as a potent antioxidant and cytoprotective agent, which suggests
that the reduced melatonin levels seen in schizophrenia patients may lead to or aggravate
an excitatory hyperoxidative status [114]. Melatonin treatment ameliorated the metabolic
syndrome caused by second-generation antipsychotics in schizophrenic patients [80, 81].
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 349

Postmortem analysis of the temporal cortex region in schizophrenia patients


revealed dampened expression of the clock gene Per1 [115]. Signaling associated
with the DA D2-receptor enhanced the transcription of clock components, whereas
DA D2-receptor mutant mice displayed impaired expression of Per1. In another
study, single nucleotide polymorphisms identified on the locus of Per3 and Timeless
were associated with schizophrenia/schizoaffective disorder [107]. However, no
correlation for Per3 length variants was observed in other studies. Our report about
a South Indian population was compatible with the view that there is no correlation
between Per3 polymorphism and schizophrenia [101].
In neuropsychiatric disorders, clock gene polymorphisms may have a direct
impact on the regulation of neural communication. Secretion of cortisol is influ-
enced by the clock genes and increased cortisol concentration has been observed in
schizophrenia patients [116], thus suggesting that clock genes might be involved in
pathogenesis. Understanding the processes relating to the internal time-keeping sys-
tem and neurobehavioral processes may shed new light on and help to develop new
approaches to the treatment of schizophrenia.

Brain Aging

The term longevity includes two different concepts. Average longevity is defined
as the average life expectancy at birth for individuals of a given species. Maximal
longevity is the maximum age that an individual of a given species can reach. The
average longevity of the human species has increased considerably throughout
history owing to the decline in infant mortality, the discovery of antibiotics, vac-
cines and, more generally, improvement in the control of infectious diseases, in
addition to more balanced nutrition, better sanitary conditions, and advances in
the treatment of diseases such as cancer or diabetes. In contrast, maximal longev-
ity has remained unmodified. In ancient times, people reaching the ages of
80–90 years or more were also found, although the percentage was much lower
than today.
The increase in average longevity is seen in the growing segment of the popula-
tion between 60 and 100 years (Fig. 8.20). Those over 80 now constitute 1.6% of the
world population and by 2050 they will constitute 4.3% (about 400 million people).
During the period from 2008 to 2040, the population of individuals aged 65 years
and above is projected to increase by 160%, whereas that of individuals aged 80
years and older will be over 230% (Fig. 8.20) [117]. Considering that the estimation
for the number of patients with Alzheimer’s disease for 2050 is 150 million people,
the goal of “successful aging” has become very important in avoiding the conse-
quences of neurodegenerative diseases, cancer, or arteriosclerosis, i.e., those most
likely to affect this elderly group. These figures represent an increase of 56% in
high-income and of 239% in low-income countries. They underlie the extreme
importance of maintaining healthy aging for public health policies in years to come.
The desired goal is to minimize chronodisruption.
350 8  Clinical Implications of the Timed Autonomic Nervous System

PERCENT INCREASE IN WORLD POPULATION


(2008–2040)

250

200

150

100

50

0
ALL AGES
65 YEARS OR
MORE 80 YEARS OR
MORE

Fig. 8.20  Projected increase in the total world population contrasts with that of the elderly popu-
lation (>65 years or >80 years)

Virtually all physiological functions become less efficient with age (Fig. 8.21).
As discussed in Chap. 2, slow-wave sleep decreases exponentially with aging and
often disappears after 60 years of age. Many elderly individuals complain of inter-
rupted sleep and of daytime sleepiness [118]. Other common complaints are early
awakenings and a poor capacity to maintain alertness in the evenings. Both are
indexes of the aging of the circadian apparatus: a decrease in the amplitude and
phase advance of circadian rhythms [119]. Thus, the relative distribution of wake-
fulness, NREM, and REM body configurations changes significantly in aged indi-
viduals (Fig. 8.2).
Aging entails a homeostatic loss of the capacity to maintain the stability of
the internal environment of the individual to combat environmental disturbances.
An example of this is the reduced ability of the elderly to withstand extreme
temperatures, trauma, infections, and stress in general. With aging, most vital
organs suffer atrophy or degeneration phenomena. This is most noticeable in
differentiated cells such as neurons, myocardial cells, muscle cells or the renal
parenchyma.
It is generally accepted that cell disruption due to the oxidative stress is a major
physiopathological event in aging. In addition, other current hypotheses suggest a
direct relationship among aging, genetic programs, and telomere loss that occurs
after each cell division, leading to apoptotic cell death. Although all these processes
and mechanisms are probably involved in diseases associated with aging, their role
in normal aging has not yet been clarified.
Studies aimed at evaluating the impact of modifiable risk factors on pathological
brain aging are justified by their potential for prevention at a population level. The
risk factors with strongest evidence for possible causal associations with brain
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 351

Chaperone
decrease

Inflammation Oxidative stress

Proteotoxicity
Genome and
inestability AGING macromolecular
aggregation

Impairment of
mitocondrial Sarcopenia
function
Decrease of
melatonin, GH,
DHEA, IGF, sex
steroids

Fig. 8.21  Some consequences of aging

pathology, e.g., dementia, are poor sleep, T2DM throughout life, hypertension in
midlife, smoking, and a low level of education in early life.
Concerning sleep loss in aging, it typically altered the three body configurations
toward a sympathetic predominance in the face of poor NREM sleep (Fig. 8.2).
Indeed, the prevalence of insomnia is up to 25–30% in the elderly, which leads to a
concomitant increase in the use of hypnotics. The benzodiazepines (BZDs) and
BZD receptor agonists (Z drugs: zolpidem, zaleplon, zopiclone) are the most com-
monly prescribed drugs for the treatment of insomnia in the elderly. The BZDs are
a group of compounds that exert their therapeutic effect on sleep through the allo-
steric modulation of GABAA receptor complex. BZDs have broad inhibitory effects
on brain functions, including promoting sleep, anxiolysis, anticonvulsant effects,
cognitive and motor impairment, and reinforcing effects. In addition, significant
adverse effects, such as cognitive and psychomotor impairment, next day hangover,
rebound insomnia, anterograde amnesia, and dependence have been documented,
352 8  Clinical Implications of the Timed Autonomic Nervous System

thus rendering the use of BZDs for the prolonged treatment of insomnia highly
controversial. “Z drugs” are a group of agents that are not part of the chemical class
BZD, but act through the same mechanism, the allosteric modulation of GABAA
receptor. Generally, Z drug hypnotics, although effective at reducing sleep latency,
are only moderately effective at increasing sleep efficiency.
Advice against long-acting hypnotic BZD and Z drug use and recommendations
to employ them for the shortest time possible in older patients (no more than 2–3
weeks of treatment) are common nowadays. For example, the American Geriatrics
Society has recently updated its list of inappropriate medications for older patients
and advised physicians to “avoid benzodiazepines (any type) for treatment of
insomnia, agitation, or delirium” [120]. Z drugs are used, unlike the BZDs, exclu-
sively for the treatment of insomnia and are supposed to have a lower tendency to
induce physical dependence and addiction than BZDs. However, adverse effects
have been reported in more than 40% of users of both types of drugs, with no differ-
ence between BZDs and Z drugs.
In Europe, health authorities are increasingly initiating policies and recommenda-
tions to reduce the consumption of BZDs and Z drugs. However, the campaigns have
not generally been successful, despite national guidelines and recommendations, and
the use of these drugs has continued to increase. The clearer strategy to reduce chronic
BZD use is to reduce medication; abrupt cessation can only be justified if a serious
adverse effect occurs during treatment. There is no clear evidence for the optimal rate
of tapering, and recommended times vary from 4 weeks to several months.
In 2007, a sustained release form for 2 mg of melatonin was approved by the
EMA for the treatment of insomnia in elderly people. The fact that melatonin does
not show evidence for dependency, isolation, rebound insomnia or a negative influ-
ence on alertness during the day was emphasized by the EMA for melatonin and by
the FDA for the melatonin analogs ramelteon and tasimelteon.
Melatonin competes with BZPs and Z drugs at their site of action. A facilitatory
role of melatonin on GABA neurotransmission was first demonstrated in the
author’s laboratory and explains the anxiolytic, antihyperalgesic, and antinocicep-
tive effects of melatonergic agents [75]. Several clinical studies now support the
efficacy of melatonin in reducing BZP use in chronically treated patients [121, 122].
In a pharmacoepidemiological study aimed at evaluating the impact of anti-BZD/Z
drugs campaigns and the availability of alternative pharmacotherapy (melatonin) in
the consumption of BZDs and Z drugs in several European countries, the results
indicated that the campaigns failed unless they were associated with the availability
of melatonin in the market [123]. Several consensus statements support the role of
melatonin in the elderly. For example, the BritishAssociation for Psychopharmacology
consensus statement on evidence-based treatment of insomnia, parasomnias, and
circadian rhythm disorders concluded that “…melatonin should be the choice hyp-
notic for insomniacs over 55 years of age” [45].
To maintain normal sleep is crucial in the elderly. As discussed in Chap. 2, the
repair function of sleep is, among others, the consequence of the increased activ-
ity of the glymphatic system with elimination of potentially neurotoxic waste
products accumulated in the CNS during wakefulness. Age-associated glymphatic
dysfunction has been reported, with decreased and delayed CSF penetration along
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 353

paravascular pathways and the pial surface [124]. The mechanics and importance
of the glymphatic system in several cerebrovascular disorders are still being
unraveled and investigations of therapeutic strategies that can protect or restore its
integrity are warranted [125].
Glymphatic dysfunction has been reported in neurological disease states such as
stroke, traumatic brain injury, and AD. The pathological signature of AD includes
extracellular senile plaques, formed mainly by amyloid β (Aβ) deposits, and intracel-
lular neurofibrillary tangles, resulting mainly from abnormally hyperphosphorylated
microtubule-associated tau protein (Fig. 8.22). Aβ is composed of 39–43 amino acid

INSULIN/IGF1

Free radicals
Mitochondrial damage Anti-
IRS-1
oxidants TNF-a
IL-6
ROS IL-1b
Insulin
Ca2+dysfunction signaling
Oxidative stress NFκB
MELATONIN IN AGING
Tryptophan
Nitrosative Neuro-
stress inflammation Autophagy
5-Hydroxytryptophan
AANAT
Serotonin

N-acetylserotonin Normal / Pathologic CNS aging LONGEVITY Sirtuin 1 FOXO1


HIOMT

Melatonin

6-Sulphatoxy- AFMK Stem cell aging


melatonin AMK
BDNF Reduced auto renewal.
Melatonin receptors
Protein de- Increased quiescence.
aggregation
Protein Stem cell senescence.
aggregation
Telomere disfunction.
Tau dysfunction Tau-P

Protein dysfunction
S-Appα

sAppβ
P3

β-secretase

γ-secretase α--secretase
γ-secretase
APP
C83
AICD

AICD

Differentiation Proliferation Neural stem


C99

APP dysfunction cell


Ubiquitin/proteosome dysfunction

Fig. 8.22  Mechanisms promoting normal (blue arrows) and pathological (red arrows) central
nervous system aging. Free radicals and mitochondrial damage promote oxidative stress and ner-
vous system aging when the level of reactive oxygen species (ROS) production is higher than that
of antioxidants. Oxidative stress leads to cell damage and calcium dysfunction. Melatonin, its
metabolites, and receptors decrease in aging. Insulin, insulin-like growth factor 1 (IGF1), and
insulin-like growth factor 2 (IGF2) act via insulin receptor substrate 1 (IRS-1) to trigger the insulin
signaling pathway stimulating, for example, autophagy, a recycling pathway that maintains protein
and organelle quality control. Inflammaging and the low-degree inflammation seen in obesity-­
related metabolic disorders lead to overproduction of proinflammatory cytokines and activate
microglia. This interferes with the ability of IRS-1 to engage in insulin signaling and blocks the
intracellular actions of insulin. Protein misfolding, aggregation, and degradation impairment are
the main characteristic features of age-related neurodegenerative diseases. In the stem cell niche,
depletion of the neural stem cell pool or decreased potential to produce progenitor cells leads to
neural stem cell aging via impairments in self-renewal, stem cell senescence, increased quies-
cence, and neural stem cell fate changes
354 8  Clinical Implications of the Timed Autonomic Nervous System

residues derived from its precursor, the amyloid precursor protein (APP). APP is
proteolytically processed by α- or β-secretases in different pathways. The α-non-
amyloidogenic pathway involves cleavage of APP by α-secretase to release a frag-
ment of APP N-terminal, which, after cleavage by γ-secretase, precludes the
formation of Aβ. The β-amyloidogenic pathway includes β-secretase, which results
in the formation of intact Aβ peptide and is mediated by the sequential cleavage of
β-secretase and γ-secretase at the N- and C-terminal of the Aβ molecule [126].
In AD, glymphatic impairment has emerged as a piece of the disease pathology
puzzle. The Aβ peptide, which typically accumulates for years preceding AD demen-
tia, is also produced by the normal brain and is present in the circulating blood and
CSF. However, unlike the healthy brain, which is able to clear Aβ via glymphatic
drainage, in AD there is a gradual build-up of Aβ in the brain parenchyma and vas-
cular structures, leading to neurovascular uncoupling, including cerebral blood flow
decrease, blood–brain barrier disruption, and impairment of vasculature [125]. The
release of Aβ from neurons is dependent on synaptic activity and before the accumu-
lation of Aβ plaques, brain Aβ levels fluctuate with the sleep–wake cycle in a pattern
in which the Aβ concentration of interstitial fluid is higher during wakefulness and
lower during sleep. Chronic sleep deprivation increases Aβ deposition. Recent results
indicate that blood–brain barrier disruption induced by chronic sleep loss is due to
the low-grade inflammation that the sleep deprivation entails [127].
A decline in cognitive capacities, including reasoning, memory, and semantic flu-
ency, characterizes normal aging and is detectable as early as the 5th decade of life.
There is evidence for a preclinical stage in dementia in which cognitive performance
is borderline compared with normal aging. In community-based studies, up to 30%
of a sample of healthy, community-dwelling, elderly individuals show deficits in
performance that were not explained by age-related changes, education levels, mood,
or health status. This strongly suggests the existence of early pathological changes,
which is a transitional state taking place between normal aging and early AD [128].
Cross-sectional studies reveal that sleep disturbances are associated with memory
and cognitive impairment. A severe disruption of the circadian timing system occurs
in AD, as indicated by alterations in numerous overt rhythms such as body tempera-
ture, glucocorticoids, and/or plasma melatonin. The internal desynchronization of
rhythms is significant in AD patients. One emerging symptom is “sundowning,” a
chronobiological phenomenon observed in AD patients in conjunction with sleep–
wake disturbances [129]. Sundowning includes symptoms such as disorganized
thinking, a reduced ability to maintain attention to external stimuli, agitation, wan-
dering, and perceptual and emotional disturbances, all appearing in late afternoon or
early evening. Chronotherapeutic interventions such as exposure to bright light and
timed administration of melatonin in selected circadian phases alleviated sundown-
ing symptoms and improved the sleep–wake patterns of AD patients (Fig. 8.7) [130].
Levels of melatonin in the CSF decrease, even in preclinical stages of AD when
the patients do not manifest any cognitive impairment, suggesting that the reduction
in CSF melatonin might be an early trigger and marker for AD. Although it is not
known whether the relative melatonin deficiency is either a consequence or a cause
of neurodegeneration, it seems clear that the loss in melatonin aggravates the
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 355

disease and that early circadian disruption can be an important deficit to be consid-
ered [75]. Significant differences were observed in melatonin levels between mild
cognitive impairment (MCI) and AD patients, with a negative correlation between
the neuropsychological assessment of dementia and melatonin levels [131].
Mild cognitive impairment is diagnosed in those who have an objective and mea-
surable deficit in cognitive functions, but with a preservation of daily activities. The
estimates of annual conversion rates to dementia vary across studies, but may be as
high 10–15%, as MCI represents a clinically important stage for identifying and
treating individuals at risk [132]. Indeed, the degenerative process in AD brain starts
20–30 years before the clinical onset of the disease. During this phase, plaque and
tangle loads increase and at a certain threshold, the first symptom appears.
An analysis of published data with melatonin in the early stages of cognitive decline
consistently showed that administration of melatonin every night improves the quality
of sleep and cognitive performance in this phase of the disease. Therefore, melatonin
treatment can be effective in the early stages of neurodegenerative disease [75].
The mechanisms accounting for the therapeutic effect of melatonin in MCI
patients remain to be defined. Melatonin treatment mainly promotes slow-wave
sleep in the elderly and can be beneficial in MCI as it augments the restorative
phases of sleep, including the secretion of GH and neurotrophins and the function-
ing of the glymphatic system. The antioxidant, mitochondrial, and anti-amyloido-
genic effects of melatonin may possibly interfere with the onset of the disease
(Fig. 8.23). Therefore, the point at which melatonin treatment begins can be deci-
sive for the final response.

AD PROGRESSION HALTED?

MAINTENANCE OF
NORMAL
MITOCHONDRIAL
FUCTION
CHRONOBIOTIC CYTOSKELETAL
circadian rhythmicity hyperphosphorylation
preserved suppressed

ANTIOXIDANT ANTIFIBRILLOGENIC
prevention of prevention of amyloid-β
amyloid-β protein protein formation
effect

Fig. 8.23  Melatonin and Alzheimer’s disease. The multiple effects of melatonin and the different
degree of overlap (interrelations and mutual influences) are indicated by their respective intersec-
tions. Reproduced with permission from Cardinali [75]
356 8  Clinical Implications of the Timed Autonomic Nervous System

Double-blind multicenter studies are needed to further explore and investigate


the potential and usefulness of melatonin as an antidementia drug in the early stages
of the disease. Acetylcholinesterase inhibitors, the first-line drugs used today do not
prevent and treat AD. So far, over 90 phase 3 trials of AD have been unsuccessful,
with a 99.0% failure rate [133]. Owing to the multifactorial nature of AD pathogen-
esis, polypharmacy with drugs that target heterogeneous pathophysiological path-
ways, must be considered. A novel pharmacological treatment paradigm (the “M”
drugs) was proposed, involving the use of melatonin, minocycline, modafinil, and
memantine [133]. Minocycline is neuroprotective, it reduces neuroinflammation
and CNS pathology, and prevents cell death. Modafinil, a wake-promoting agent,
improves global mental status, hippocampal neurogenesis, attention, and cognition.
Memantine is a NMDA receptor antagonist and is approved for the management of
moderate-to-severe AD. This sort of strategy could provide a comprehensive and
pragmatic means of combating multiple pathological targets and ameliorating cog-
nitive dysfunction in AD.
Other major risk factors for dementia are hypertension and T2DM. They occur
in middle-aged and older adults and are strongly influenced by the dysregulation
of insulin signaling, which starts with insulin resistance and is followed by hyper-
insulinemia, metabolic syndrome and finally T2DM. Consumption of energy-
dense diets, high in saturated fat and sugar, is associated not only with weight gain
and metabolic syndrome, but also with impaired hippocampal-dependent memory
and the emergence of pathological hippocampal conditions [134]. The hippocam-
pus is one of the first brain regions in AD to show Aβ deposition, possibly associ-
ated with cognitive impairment. It has been demonstrated that greater cognitive
and affective decline occurs in AD patients with metabolic syndrome than in those
without, suggesting that insulin resistance and vascular endothelial dysfunction
might be strongly correlated with AD before pathological changes of the brain
can be observed. In this context, the hypothesis that metabolic syndrome might
operate as a “second hit” is suggestive as a potential trigger of AD progression
(Fig. 8.24) [134].
Inflammaging as a central phenomenon in senescence and its role in the develop-
ment of age-associated brain diseases are now recognized. Genetic predispositions
such as an immune risk profile, which comprises an increased tendency toward
inflammatory responses, may set limits to health and lifespan, whereas an “inverted
immune risk profile” found in centenarians may be the basis of successful aging.
This is indicated by enhanced inflammation in neurodegenerative diseases and the
effects of inflammatory signals and free radicals released by immune cells on neu-
rons and astrocytes [10].
The immune system is remodeled with aging. One of the contributing factors is
a progressive thymic involution, which leads to losses in the number of both CD4+
and CD8+ T lymphocytes, in addition to a disturbed balance among naïve, memory,
and effector T cells. For example, the subset of CD8+ cells is largely replaced by the
clonal expansion of cells of lower proliferative activity. Type 1 cytokines are ele-
vated, such as IL-2, IFN-γ, and TNF-α, and, to a lesser extent, type 2 cytokines,
such as IL-4, IL-6, and IL-10. This contributes to a higher inflammatory state;
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 357

PERIPHERY Blood Brain Barrier

OBESITY Microglial  IL-6, IL-


 TNFa activation 1β, TNFα

¯ Insulin/ ¯ Insulin -R activity / ¯ MAPK


IGF-1 IGF-1-R

¯ IRS-1, ¯ Growth,
IRS-2 proliferation,
differentiation
¯
Synaptic Cell
plasticity
¯ PI3K death
Neuroinflammation
¯
β-amiloide
yTAU
¯ Learning, memory
CENTRAL
ALZHEIMER’S NERVOUS
DISEASE SYSTEM

Fig. 8.24  Possible links between obesity and neurodegeneration. Obesity promotes chronic low-­
grade peripheral inflammation and insulin and IGF-1 resistance. Cerebrovascular dysfunction,
together with increased blood–brain barrier permeability, allows macrophage/cytokine entry and
reduced transport of trophic factors. Chronic inflammation, coupled with insulin and IGF-1 resis-
tance, promotes neurodegenerative pathological conditions

however, with large interindividual differences. Age-dependent changes occur for


CD4+ T lymphocytes, mostly in the activated/memory cells, among which TNFα-
positive cells are decreased, but IL-4-positive cells increased, reflecting a shift from
type 1 to type 2 cytokines. B cell immunosenescence results in decreases of IgM
and IgD production, whereas increased levels of IgG1 have been reported.
Concerning the innate immune system, centenarians have a relatively higher num-
ber of NK cells, especially NKT cells bearing the γδ-T cell receptor, in contrast to
less successfully aging subjects. Moreover, these cells had a stronger response to
activation in centenarians by releasing IFN-γ, underlying the importance of a robust
innate immune system in longevity [135].
In addition to the gradual functional losses, immunosenescence has two other
undesired consequences, a higher incidence of autoimmune responses and increased
levels of inflammatory mediators. This latter observation is also made in centenari-
ans. However, in these successfully aged subjects, the higher quantities of proin-
flammatory factors are associated with augmentations of anti-inflammatory
cytokines and a protective genotype [136].
Inflammaging differs from acute or chronic inflammation elsewhere in the body.
It is of a slowly progressing, lingering type, with moderate microglia activation that
358 8  Clinical Implications of the Timed Autonomic Nervous System

is sustained by the degenerative processes and oxidative stress resulting from the
release of ROS and RNS by immune cells, astrocytes, and neurons that is further
enhanced by damage to mitochondria (Fig. 8.25). Inflammaging, especially in neu-
rodegenerative diseases, is intertwined with other potentially deteriorating pro-
cesses, among which mitochondrial dysfunction is of prime importance [10].
Inflammaging entails a senescence-associated secretory phenotype, which
depends on DNA damage, and becomes more likely with increasing age, accumu-
lating over time. Senescent cells carrying damaged DNA are usually mitotically
arrested, a mechanism that keeps them alive and metabolically active. However,
these cells display a chronic DNA damage response leading to the release of proin-
flammatory cytokines, the hallmark of the senescence-associated secretory pheno-
type (Fig. 8.25). Aging astrocytes also express senescence-associated secretory
phenotype characteristics [10].
Neuronal overexcitation and inflammatory responses are intimately associated.
Although glutamate excitotoxicity can lead to microglia activation, primary immune
responses involving the microglia may, in turn, lead to excitotoxicity. Astrocytes are
frequently coactivated with microglial cells. They also contribute to both

Apoptosis vía cytochrome C Reduction of mitochondrial


release and mtPTP mass and peripheral
breakdown mitochondria

Mitophagy Insufficient energy supply

ETC blockade and


damage Impaired cell functions
and viability

Neuronal overexcitation Damage to


with calcium overload endothelia
Damage to
ROS
proteins and
Activation of microglia &
lipids
NOS
Cancer stem
Damage to cells
INFLAMMAGING DNA Reduction of stem
cells with high
Nox isoforms proliferative capacity
Radiation and Increased telomere
DDR with attrition
other external proliferative arrest Reduced tissue repair and
factors reduction of immune
progenitor cells
Proinflammatory SASP
cytokines
Autoimmune
responses
Long-term exposure to
foreign antigens Reduced germ and
IMMUNOSENESCENCE virus resistance
Thymic involution

Fig. 8.25  The imbalance between inflammatory and anti-inflammatory signals is a hallmark of
aging and contributes to its progression. An age-related proinflammatory tendency (inflammaging)
is mostly unavoidable because of thymic involution and extended germ exposure. Mitochondrial
dysfunction ensues with disruption of the electron transport chain and ROS and reactive nitrogen
species generation. Reproduced with permission from Cardinali [75]
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 359

excitotoxicity and microglia activation, by mechanisms that involve impaired gluta-


mate uptake, inflammatory signals, and/or oxidative/nitrosative stress.
Recent results indicate that blood–brain barrier disruption induced by chronic
sleep loss is due to the low-grade inflammation that the sleep deprivation entails
[127]. Figure 8.24 summarizes the possible links between obesity and neurodegen-
eration in AD. Obesity promotes chronic low-grade peripheral inflammation and
insulin and insulin growth factor (IGF)-1 resistance. Cerebrovascular dysfunction,
together with increased blood–brain barrier permeability, allows macrophage/cyto-
kine entry and reduced transport of trophic factors. Chronic inflammation, coupled
with insulin and IGF-1 resistance, promotes neurodegenerative pathological condi-
tions. Potential signaling pathways further lead to AD-like molecular and cognitive
changes via increased Aβ and tau phosphorylation (Fig. 8.24).
Neurodegeneration is facilitated by stimulation of proinflammatory processes in
different cell types, that implies the possibility of positive feedback loops between
neurons, astrocytes, and microglia, and that expand the grade and area of inflamma-
tion recruiting other immune cells. Activation of microglia can lead to different
phenotypes, from neurodestructive and phagocytic cells to others that are primarily
neuroprotective or promote growth.
Inflammaging is specifically accelerated and aggravated under neuropathologi-
cal conditions. Sustained, progressive, primarily low-grade inflammation is
observed in presumably all neurodegenerative disorders, with differences in details,
pathways, and affected cells. Neuroinflammation has been demonstrated in
Huntington’s disease, amyotrophic lateral sclerosis, Friedreich’s ataxia, Parkinson’s
disease, frontotemporal lobe degeneration, and AD [10].
As the early phases of neurodegenerative disorders are poorly distinguishable
from normal inflammaging or other forms of low-grade inflammation, these patho-
logical conditions display important mechanisms of aggravation, in terms of vicious
cycles. These comprise the accumulation of toxic products, such as Aβ peptides,
oligomers, and plaques in AD, microglial proliferation and phagocytosis after their
activation, and the progressive formation of proinflammatory signal molecules.
Elevated levels of IL-6 and TNFα are demonstrable in the CNS and in the circula-
tion, reflecting the dysregulation of inflammatory pathways, and are meanwhile
regarded as markers of frailty. Interrupting the proinflammatory vicious cycles rep-
resents a main challenge in combating neurodegenerative diseases [135, 136].
A particular aspect of brain inflammaging concerns the impaired function of
SCN, which is observed during normal aging and, even more frequently, in neuro-
degenerative disorders. This change may be of high relevance, because it systemi-
cally affects a host of other functions in the body (Chap. 2). Several alterations can
be involved in SCN insufficiency, from reduction of blue light perception to losses
in signal transmission to the SCN, but a major factor is that of SCN neural degenera-
tion. Because of the crucial role of the SCN in the control of the mammalian pineal
gland, dysfunction of the hypothalamic master clock strongly contributes to aging-
or disease-related reductions of nocturnal melatonin secretion and changes in the
secretory patterns and phasing of melatonin [131]. Although SCN deterioration is
more strongly pronounced in AD and other forms of dementia, it seems that the
360 8  Clinical Implications of the Timed Autonomic Nervous System

relevance of this phenomenon is already high in normal aging. In part, this may be
related to changes in melatonin, but additional impairments of other rhythmic func-
tions likely arise when a master clock decays. The replacement of the poorly func-
tional SCN of a senescent hamster by transplantation of a juvenile SCN not only
restored the previously decomposed circadian rhythmicity, but caused a rejuvena-
tion in terms of physical appearance and extended the lifespan of the recipient [137].
These findings impressively show how important the SCN and a well-operating
circadian system are for preventing aging-related impairments. Because of the
reduction in neurons in the SCN the sequence and spacing of the maximum values
of daily rhythms are progressively decreased. Correction of chronodisruption by
chronobiological treatment is thus needed for maintaining health in the elderly
(Fig. 8.26).
A role of melatonin in attenuating inflammaging and its progression has been
especially discussed with regard to treatment options under conditions of reduced
endogenous melatonin levels [10]. Melatonin declines during aging and, even more
so in several age-related diseases, changes that have been documented in humans.
Interindividual variations observed among elderly persons may be explained, to a
certain extent, by differences in the acquisition of melatonin-depressing diseases
and disorders.
Reversion of inflammaging by melatonin occurs at various levels (Fig. 8.27).
One level concerns the correction of metabolic dysregulation, including the preven-
tion of insulin resistance, an inflammation-promoting change and hallmark of meta-
bolic syndrome. Notably, and as already mentioned, melatonin was effective at
suppressing insulin resistance. Melatonin reversed the blockade of a key step in

By correcting
AGING chrono-
disruption:
• Sleep-wake
cycle is
improved
• CNS trophic
YOUNG OLD functions
are reset
Amplitude

Amplitude

• Better
immuno-
endocrine
homeostasis

Time Time

Fig. 8.26  The aging of the circadian apparatus results in a decrease in amplitude and phase-­
advance of circadian rhythms. By restoring chronodisruption, the sleep–wake cycle improves, cen-
tral nervous system trophic functions are reset and better immunoendocrine homeostasis occurs.
Reproduced with permission from Cardinali [75]
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 361

Actions on central and peripheral


circadian oscillators; effects on clock
proteins with tumor suppressor and
aging suppressor properties

Inhibition of Modulation of
neuronal
metabolic sensing
overexcitation

MELATONIN
Improved insulin
Support of mitochondrial sensitivity; counteraction
electron flux, reduction of of metabolic syndrome
electron overflow
Reduction of ROS & RNS by
lmproved upregulation of antioxidant
enzymes, inhibition of prooxidant Mitochondrial proliferation
respiratory
enzymes, increase of GSH, radical
Prevention efficiency
scavenging and lower radical
/inhibition formation Improved energy supply
of
apoptosis

Less mitophagy Reduced damage to proteins, lipids Anti-


and DNA, to mitochondria and other inflammatory
organelles actions
Avoidance of
Reduced number of DNA-damaged,
oxidant-induced Modulation of natural and
mitotically arrested cells with SASP
telomere attrition adaptive immunity:
improved
Greater number of cells with high immunosenescence, less
proliferative capacity, including inflammaging
leukocytes, stem and progenitor
cells

Fig. 8.27  Overview of the multiple actions of melatonin that antagonize brain inflammaging.
Reproduced with permission from Cardinali [75]

insulin signal transduction, i.e., the reduced phosphorylation of IRS-1 (insulin


receptor substrate 1), which is generally accompanied by an upregulation of IRS-1
expression [75].
A further level of action of melatonin concerns the suppression of processes that
favor or lead to inflammation. This comprises calcium overload, excessive NO
release that results in the formation of peroxynitrite, peroxynitrite-derived free radi-
cals and tyrosine nitration, and mitochondrial dysfunction (Fig. 8.27).
The immunological effects of melatonin represent a third area relevant to inflam-
maging. The role of melatonin as an immunemodulatory agent comprises both pro-
inflammatory and anti-inflammatory actions, which, consequently, leads to an either
pro-oxidant or antioxidant balance. In several conditions concerning senescence,
the anti-inflammatory side of melatonin seems to prevail [138].

Cancer

Cancer is responsible for about 25% of all deaths in the Western world. Most diag-
noses of cancer occur in people over 55 years of age, with breast and prostate cancer
being the leading types of cancer followed by lung cancer. More than half of new
cases of cancer are breast, prostate, lung or gastrointestinal cancer [139]. Many etio-
logical factors including genetic, environmental, dietary, hormonal and aging, the
362 8  Clinical Implications of the Timed Autonomic Nervous System

immune status or presence of medical or psychiatric illnesses have all been sug-
gested as predisposing factors for the development of cancer in humans.
Chronodisruption is linked to cancer. On the one hand, cancer itself (e.g., tumor
invasion symptoms, pain), chemotherapy, corticosteroid treatment, environmental
factors, or psychological distress are among the factors that contribute to chronodis-
ruption, immunosuppression, and disruption of sleep. On the other hand, chronodis-
ruption given by exposure to light at night is considered a major cause of the
increased risk of cancer [140].
Working in nondaylight hours are associated with an increased risk of cancer,
and the International Agency for Research on Cancer, WHO, classified shift work
with chronodisruption as a probable human carcinogen (group 2A carcinogen)
[141]. The reduction of melatonin levels following repeated nocturnal exposure to
light, as this occurs in women engaged in night-shift work, can result in higher rates
of breast cancer development and proliferation [142]. Breast cancer accounts for
25% of all cancers and caused 522,000 deaths worldwide in 2012. A meta-analysis
of 16 investigations, involving 2,020,641 participants, 10,004 incident breast cancer
cases, 7185 cancer-related deaths, 4820 cardiovascular endpoints, and 2480 all-
cause mortalities indicated that night work increased the risk of breast cancer mor-
bidity by: 1.9% for 5 years, 2.5% for 5–10 years, 7.4% for 10–20 years, and 8.8%
for >20 years of work [143]. Additionally, night work enhanced the morbidity of
breast cancer by 8.9% and was associated with a 2.7% increase in cardiovascular
death.
Circadian genes are essential in the regulation of the cell cycle, a finely regulated
process from a cell that can generate multiple cells through a series of cell divisions,
including four critical and successive steps named G1 phase (growth phase 1), S
phase (synthesis), G2 (growth phase 2), and M phase (mitosis). Knock-down of the
circadian gene Bmal1 in the carcinoma cells of the colon, fibroblast cells, and intes-
tine epithelial cells produces cellular proliferation in vitro and increments the size
of tumor cells injected subcutaneously, via, among others, the inhibition of apopto-
sis and the reduction of the time transition between G2/M. Moreover, the knock-out
for clock genes Bmal1 and Per2 in mice previously exposed to radiation caused
hyperplastic growth and development of lymphoma, hepatic carcinomas, ovary
tumors, and osteosarcomas [144].
In humans, the association between clock gene polymorphisms and cancer is
common. An incremented risk of breast cancer, colorectal carcinoma, prostate can-
cer, lung cancer, non-Hodgkin’s lymphoma, glioma, and primary hepatocellular
carcinoma were associated with alleles of circadian genes. Therefore, clock genes
may be modulators of the cellular cycle that modulates the risk of cancer [144].
In addition, the oncogenic alteration of circadian rhythms has also been reported
to occur. Mutations in circadian clock genes, including promoter methylation, cod-
ing region mutation, deletion, or rare amplification, have been documented across
many different types of cancer [145]. Given that these mutations disrupt normal
oscillation, it has been suggested that the clock might be tumor-suppressive. Many
proto-oncogenes and tumor suppressors are normally under circadian control; thus,
disruption of oscillation could potentially release these proteins to be constitutively
Some Clinical Autonomic Entities Associated with a Desynchronized ANS 363

overexpressed or suppressed. Given the frequency of mutations, it can be speculated


that many cancers with oncogenic mutations have altered or disrupted circadian
rhythm and altered oscillation of gene expression and metabolism [145].
With few exceptions, melatonin levels were found to be low in cancer patients
[146]. This may be an indication for the ongoing chronodisruption, as circulating
melatonin levels are an index of SCN function (Chap. 2). In addition, another preva-
lent idea is that the reduced melatonin secretion plays a role in the occurrence of
cancer in chronodisruption.
The light–melatonin–cancer hypothesis has been supported by experimental
studies conducted on athymic rats in which human breast or prostate cancer tissue
was transplanted. Rats that had been exposed to a constant light environment exhib-
ited a sevenfold increase in tumor growth compared with rats that had remained in
normal light/dark cycles [147]. There was an augmented uptake of linoleic acid and
of its metabolism in rats kept under constant illumination. This accelerated rate of
linoleic acid metabolism was attributed to the suppression of the circadian melato-
nin signal, which normally inhibits linoleic acid uptake at nighttime. Compared
with tumors perfused with melatonin-deficient human blood collected during the
daytime, human breast cancer xenografts and rat hepatomas perfused in situ with
nocturnal, physiologically melatonin-rich blood collected during the night, exhib-
ited markedly suppressed proliferative activity and linoleic acid uptake/metabolism
[148]. In the case of prostate cancer xenografts, the amplification of nighttime mela-
tonin levels by exposing nude rats to blue light during the daytime significantly
reduces the human prostate cancer metabolic, signaling, and proliferative activities
[149]. These studies are relevant for explaining the risk of increased breast and
prostate cancer risk in night-shift workers [143] and that of colon cancer [150].
Additionally, melatonin disruption increases the risk of lung cancer [151]. The
oncostatic properties of melatonin are supported by several experimental and clini-
cal studies [146].
Sleep disturbance is a major outcome of chronodisruption among cancer patients
[152]: 30–60% of cancer patients report insomnia symptoms, whereas approxi-
mately 20% of them meet the diagnostic criteria for an insomnia disorder, which is
more than twice as frequent as in the general population. However, they generally
remained underdiagnosed and poorly treated [153–156].
Prevalence rates of sleep disturbances have also been found to be up to three
times higher in patients undergoing chemotherapy compared with the general popu-
lation. Sleep difficulties may occur before, during, and may persist long after the
end of cancer treatment [157]. An 18-month longitudinal study revealed that such
persistence was even more likely to occur in patients with an insomnia syndrome
[158]. Moreover, about 20% of patients who experienced an insomnia remission
had a relapse later during the study.
Sleep–wake disturbance is among the most severe and common symptoms
reported by primary brain tumor patients, particularly those undergoing radiation
therapy. As with other cancers and neurological illness, sleep–wake disturbance
may also be clustered or related to additional symptoms such as fatigue, depression,
and cognitive impairment [159].
364 8  Clinical Implications of the Timed Autonomic Nervous System

It must be noted that an imprecise conceptualization of chronodisruption in can-


cer has led to narrowly focused interventions being diffusely targeted to symptoms,
rather than focused and specific to one or more chronobiological disorders underly-
ing those symptoms. The consequences of chronodisruption in cancer are numerous
and can negatively affect both psychological and physical functioning. Compared
with the consequences of cancer itself, those related to sleep disruption are often
overlooked both by the patients and by the health care providers.
The most common consequences reported by the patients are symptoms of
fatigue, psychological distress, impaired daytime functioning, and disrupted cogni-
tive functioning. Chronodisruption is also associated with an increased risk of sub-
sequently developing a psychiatric disorder (e.g., anxiety and depressive disorder),
exacerbation of pain, impaired immune functioning, and increased risk of infec-
tions. As stated in Figs. 8.7 and 8.8, the concomitant chronodisruption must be
adequately treated to obtain full recovery, i.e., the specific treatments must be com-
bined with a chronobiological adjuvant treatment.
Pharmacotherapy is currently the most common treatment employed to counter-
act sleep difficulties in the general population, and in cancer patients. Surveys that
have documented the use of sedative and/or hypnotic medications among cancer
patients reached utilization rates close to 25%. Among the factors associated with
the increased use of hypnotic medication, being older, experiencing more stressful
life events during the past 6 months, suffering higher levels of anxiety or past or
current chemotherapy treatment were quoted [160]. Almost 80% of participants
who were taking drugs were prescribed BZDs (mostly lorazepam and oxazepam),
followed by zopiclone (9%). As discussed earlier in this chapter, the side effects and
risks associated with the usage of BZPs and Z drugs include drowsiness, dizziness,
headache, cognitive impairments, loss of motor coordination, and a risk of tolerance
and dependence when the medication is used daily to treat chronic difficulties.
Depression is a frequent and serious comorbid condition affecting the quality of
life. Such comorbidity reduces compliance with treatment and aggravates the physi-
cal consequences of the disease. Although there are studies showing that about 40%
of tumor patients need professional psycho-oncological support, less than 10% of
patients are referred for psychosocial intervention in daily clinical practice. Studies
of effective pharmacotherapy are relatively scarce in cancer patients with depres-
sion and they are biased by a high number of dropouts because of the side effects
relating to the use of antidepressants compared with placebo [161]. It is therefore
difficult to determine with clarity what is the best pharmacological treatment for
major depression in cancer patients.
Circadian rhythm abnormalities, as shown by the sleep–wake cycle disturbances,
constitute one the most prevalent signs of depression. The disturbances in the ampli-
tude and rhythm of melatonin secretion that occur in patients with depression
resemble those seen in subjects with chronobiological disorders, thus suggesting a
link between disturbance of melatonin secretion and depressed mood. As melatonin
is involved in the regulation of both circadian rhythms and sleep, any antidepressant
drug with effects on melatonin receptors could be an advantage in treatment.
Melatonin has been found to be effective at treating circadian rhythm disorders and
References 365

has antidepressant activity [75]. Among the analogs developed to improve the effi-
cacy of the effects of melatonin, agomelatine has been licensed by the EMA for the
treatment of major depression disorder in adults (Chap. 6).
Melatonergic receptors, particularly MT1, are also involved in the sedatory and
anxiolytic effects of melatonin that have been linked to a facilitatory role in GABA
transmission. This anti-excitatory action of melatonin underlies the anxiolytic, anti-
hyperalgesic, and antinociceptive effects of melatonergic agents, all of them having
potential application in cancer patients [75].
Increasing importance is given to chronotherapy, or the timed administration of
treatment to patients, based on circadian rhythm, to increase the efficacy and reduce
the toxicity of drugs or radiation. Several traditional cancer therapeutics, including
the anti-metabolite folate pathway antagonist methotrexate, have known circadian-­
dependent toxicity. In addition, new research indicates that several targeted thera-
pies currently in clinical use have strongly circadian-dependent efficacy depending
on the time of day given, including erlotinib (inhibits epidermal growth factor
receptor, EGFR, used in lung cancer), lapatinib (inhibits HER/Neu and EGFR, used
in breast cancer), and evirolimus (inhibits mTOR, used in some breast cancers and
pancreatic neuroendocrine tumors). In fact, there are several chronotherapy dosing
schedules under clinical trial [162]. An appropriate adjuvant chronobiological treat-
ment (Fig. 8.8) contributes to keeping the circadian system normal, with obvious
advantages for chronotherapy.

References
1. Chang AM, Aeschbach D, Duffy JF, Czeisler CA. Evening use of light-emitting eReaders
negatively affects sleep, circadian timing, and next-morning alertness. Proc Natl Acad Sci U
S A. 2015;112:1232–7.
2. Lavie P. Sleep-wake as a biological rhythm. Annu Rev Psychol. 2001;52:277–303.
3. Gringras P, Middleton B, Skene DJ, Revell VL. Bigger, brighter, bluer-better? Current light-­
emitting devices – adverse sleep properties and preventative strategies. Front Public Health.
2015;3:233.
4. Covassin N, Singh P. Sleep duration and cardiovascular disease risk: epidemiologic and
experimental evidence. Sleep Med Clin. 2016;11:81–9.
5. Blanco M, Kriguer N, Pérez Lloret S, Cardinali DP. Attitudes towards treatment among
patients suffering from sleep disorders. A Latin American survey. BMC Fam Pract. 2003;4:17.
6. Wright KP Jr, Bogan RK, Wyatt JK. Shift work and the assessment and management of shift
work disorder (SWD). Sleep Med Rev. 2013;17:41–54.
7. Singh A, Yeh CJ, Verma N, Das AK. Overview of attention deficit hyperactivity disorder in
young children. Health Psychol Res. 2015;3:2115.
8. Smolensky MH, Hermida RC, Portaluppi F. Circadian mechanisms of 24-hour blood pres-
sure regulation and patterning. Sleep Med Rev. 2017;33:4–16.
9. Cardinali DP. Qué es el Sueño. Paidós: Buenos Aires; 2014.
10. Hardeland R, Cardinali DP, Brown GM, Pandi-Perumal SR. Melatonin and brain inflammag-
ing. Prog Neurobiol. 2015;127-128:46–63.
11. Anothaisintawee T, Reutrakul S, Van CE, Thakkinstian A. Sleep disturbances compared to
traditional risk factors for diabetes development: systematic review and meta-analysis. Sleep
Med Rev. 2016;30:11–24.
12. Kaur J. A comprehensive review on metabolic syndrome. Cardiol Res Pract. 2014;2014:943162.
366 8  Clinical Implications of the Timed Autonomic Nervous System

13. Reid KJ, Abbott SM. Jet Lag and shift work disorder. Sleep Med Clin. 2015;10:523–35.
14. Agorastos A, Linthorst AC. Potential pleiotropic beneficial effects of adjuvant melatonergic
treatment in posttraumatic stress disorder. J Pineal Res. 2016;61:3–26.
15. Laermans J, Depoortere I. Chronobesity: role of the circadian system in the obesity epidemic.
Obes Rev. 2016;17:108–25.
16. Oldham MA, Lee HB, Desan PH. Circadian rhythm disruption in the critically ill: an oppor-
tunity for improving outcomes. Crit Care Med. 2016;44:207–17.
17. Madrid-Navarro CJ, Sanchez-Galvez R, Martinez-Nicolas A, Marina R, Garcia JA, Madrid
JA, Rol MA. Disruption of circadian rhythms and delirium, sleep impairment and sepsis in
critically ill patients. potential therapeutic implications for increased light-dark contrast and
melatonin therapy in an ICU environment. Curr Pharm Des. 2015;21:3453–68.
18. Waterhouse J, Reilly T, Atkinson G, Edwards B. Jet lag: trends and coping strategies. Lancet.
2007;369:1117–29.
19. Cajochen C, Munch M, Knoblauch V, Blatter K, Wirz-Justice A. Age-related changes in the
circadian and homeostatic regulation of human sleep. Chronobiol Int. 2006;23:461–74.
20. Srinivasan V, Singh J, Pandi-Perumal SR, Brown GM, Spence DW, Cardinali DP. Jet lag,
circadian rhythm sleep disturbances, and depression: the role of melatonin and its analogs.
Adv Ther. 2010;27:796–813.
21. Golombek DA, Casiraghi LP, Agostino PV, Paladino N, Duhart JM, Plano SA, Chiesa JJ. The
times they’re a-changing: effects of circadian desynchronization on physiology and disease.
J Physiol Paris. 2013;107:310–22.
22. Ariznavarreta C, Cardinali DP, Villanua M, Granados B, Martin M, Chiesa JJ, Golombek DA,
Tresguerres JAF. Circadian rhythms in airline pilots submitted to long-haul transmeridian
flights. Aviat Space Environ Med. 2002;73:445–55.
23. Tresguerres J, Ariznavarreta C, Granados B, Martín M, Villanua M, Golombek D, Cardinali
DP. Circadian urinary 6-sulphatoxymelatonin, cortisol excretion and locomotor activity in
airline pilots during transmeridian flights. J Pineal Res. 2001;31:16–22.
24. van DA, Boot CR, Hlobil H, Twisk JW, Smid T, van der Beek AJ. Evaluation of an mHealth
intervention aiming to improve health-related behavior and sleep and reduce fatigue among
airline pilots. Scand J Work Environ Health. 2014;40:557–68.
25. McEwen BS. Sleep deprivation as a neurobiologic and physiologic stressor: allostasis and
allostatic load. Metabolism. 2006;55:S20–3.
26. Coutinho JF, Goncalves OF, Maia L, Fernandes VC, Perrone-McGovern K, Simon-Dack
S, Hernandez K, Oliveira-Silva P, Mesquita AR, Sampaio A. Differential activation of the
default mode network in jet lagged individuals. Chronobiol Int. 2015;32:143–9.
27. Pinheiro SP, Schernhammer ES, Tworoger SS, Michels KB. A prospective study on habitual
duration of sleep and incidence of breast cancer in a large cohort of women. Cancer Res.
2006;66:5521–5.
28. Megdal SP, Kroenke CH, Laden F, Pukkala E, Schernhammer ES. Night work and breast
cancer risk: a systematic review and meta-analysis. Eur J Cancer. 2005;41:2023–32.
29. Dawson D, Reid K. Fatigue, alcohol and performance impairment. Nature. 1997;388:235.
30. Diez JJ, Vigo DE, Pérez Lloret S, Ritgers S, Rolé N, Cardinali DP, Pérez-Chada D. Sleep
habits, alertness, cortisol levels and cardiac autonomic activity in short distance bus drivers.
Differences between morning and afternoon shifts. J Occup Environ Med. 2011;36:806–11.
31. Diez JJ, Vigo DE, Cardinali DP, Pérez-Chada D. Sleep habits, daytime sleepiness and work-
ing conditions in short-distance bus drivers. Int J Workplace Health Manag. 2014;7:202–12.
32. Eckel RH, Depner CM, Perreault L, Markwald RR, Smith MR, McHill AW, Higgins J,
Melanson EL, Wright KP Jr. Morning circadian misalignment during short sleep duration
impacts insulin sensitivity. Curr Biol. 2015;25:3004–10.
33. Hardeland R. Melatonin and circadian oscillators in aging--a dynamic approach to the multi-
ply connected players. Interdiscip Top Gerontol. 2015;40:128–40.
34. Reid KJ, Burgess HJ. Circadian rhythm sleep disorders. Prim Care. 2005;32:449–73.
35. Shanahan TL, Czeisler CA. Physiological effects of light on the human circadian pacemaker.
Semin Perinatol. 2000;24:299–320.
References 367

36. Dawson D, Armstrong SM. Chronobiotics--drugs that shift rhythms. Pharmacol Ther.


1996;69:15–36.
37. Pandi-Perumal SR, Srinivasan V, Maestroni GJM, Cardinali DP, Poeggeler B, Hardeland
R. Melatonin: nature’s most versatile biological signal? FEBS J. 2006;273(13):2813–38.
38. Herxheimer A, Petrie KJ. Melatonin for the prevention and treatment of jet lag. Cochrane
Database Syst Rev. 2002;CD001520.
39. Czeisler CA, Kronauer RE, Allan JS, Duffy JF, Jewett ME, Brown EN, Ronda JM. Bright
light induction of strong (type 0) resetting of the human circadian pacemaker. Science.
1989;244:1328–33.
40. Cardinali DP, Bortman GP, Liotta G, Perez LS, Albornoz LE, Cutrera RA, Batista J, Ortega
GP. A multifactorial approach employing melatonin to accelerate resynchronization of sleep-­
wake cycle after a 12 time-zone westerly transmeridian flight in elite soccer athletes. J Pineal
Res. 2002;32:41–6.
41. Cardinali DP, Furio AM, Reyes MP, Brusco LI. The use of chronobiotics in the resynchroni-
zation of the sleep-wake cycle. Cancer Causes Control. 2006;17:601–9.
42. Smolensky MH. Circadian rhythms in medicine. CNS Spectr. 2001;6:467–82.
43. Crowley SJ, Eastman CI. Phase advancing human circadian rhythms with morning bright
light, afternoon melatonin, and gradually shifted sleep: can we reduce morning bright-light
duration? Sleep Med. 2015;16:288–97.
44. Ferracioli-Oda E, Qawasmi A, Bloch MH. Meta-analysis: melatonin for the treatment of
primary sleep disorders. PLoS One. 2013;8:e63773.
45. Wilson SJ, Nutt DJ, Alford C, Argyropoulos SV, Baldwin DS, Bateson AN, Britton TC,
Crowe C, Dijk DJ, Espie CA, Gringras P, Hajak G, Idzikowski C, Krystal AD, Nash JR,
Selsick H, Sharpley AL, Wade AG. British Association for Psychopharmacology consensus
statement on evidence-based treatment of insomnia, parasomnias and circadian rhythm dis-
orders. J Psychopharmacol. 2010;24:1577–601.
46. Dinges DF. The state of sleep deprivation: from functional biology to functional conse-
quences. Sleep Med Rev. 2006;10:303–5.
47. O’Neill S, O’Driscoll L. Metabolic syndrome: a closer look at the growing epidemic and its
associated pathologies. Obes Rev. 2015;16:1–12.
48. Makki K, Froguel P, Wolowczuk I. Adipose tissue in obesity-related inflammation and insulin
resistance: cells, cytokines, and chemokines. ISRN Inflamm. 2013:139239.
49. Cardinali DP, Hardeland R. Inflammaging, metabolic syndrome and melatonin: a call for
treatment studies. Neuroendocrinology. 2017;104:382–97.
50. Tchernof A, Despres JP. Pathophysiology of human visceral obesity: an update. Physiol Rev.
2013;93:359–404.
51. Lemche E, Chaban OS, Lemche AV. Neuroendocrine and epigetic mechanisms subserving auto-
nomic imbalance and HPA dysfunction in the metabolic syndrome. Front Neurosci. 2016;10:142.
52. Pagano ES, Spinedi E, Gagliardino JJ. White adipose tissue and circadian rhythm dysfunc-
tions in obesity pathogenesis and available therapies. Neuroendocrinology. 2017;104:347–63.
53. Arora T, Chen MZ, Cooper AR, Andrews RC, Taheri S. The impact of sleep debt on excess
adiposity and insulin sensitivity in patients with early type 2 diabetes mellitus. J Clin Sleep
Med. 2016;12:673–80.
54. Arora T, Chen MZ, Omar OM, Cooper AR, Andrews RC, Taheri S. An investigation of the
associations among sleep duration and quality, body mass index and insulin resistance in
newly diagnosed type 2 diabetes mellitus patients. Ther Adv Endocrinol Metab. 2016;7:3–11.
55. McFadden E, Jones ME, Schoemaker MJ, Ashworth A, Swerdlow AJ. The relationship
between obesity and exposure to light at night: cross-sectional analyses of over 100,000
women in the Breakthrough Generations Study. Am J Epidemiol. 2014;180:245–50.
56. Fonken LK, Nelson RJ. The effects of light at night on circadian clocks and metabolism.
Endocr Rev. 2014;35:648–670.
57. Karthikeyan R, Marimuthu G, Spence DW, Pandi-Perumal SR, Bahammam AS, Brown GM,
Cardinali DP. Should we listen to our clock to prevent type 2 diabetes mellitus? Diabetes Res
Clin Pract. 2014;106:182–90.
368 8  Clinical Implications of the Timed Autonomic Nervous System

58. Marcheva B, Ramsey KM, Buhr ED, Kobayashi Y, Su H, Ko CH, Ivanova G, Omura C, Mo
S, Vitaterna MH, Lopez JP, Philipson LH, Bradfield CA, Crosby SD, JeBailey L, Wang X,
Takahashi JS, Bass J. Disruption of the clock components CLOCK and BMAL1 leads to
hypoinsulinaemia and diabetes. Nature. 2010;466:627–31.
59. Shi SQ, Ansari TS, McGuinness OP, Wasserman DH, Johnson CH. Circadian disruption
leads to insulin resistance and obesity. Curr Biol. 2013;23:372–81.
60. Gale JE, Cox HI, Qian J, Block GD, Colwell CS, Matveyenko AV. Disruption of circadian
rhythms accelerates development of diabetes through pancreatic beta-cell loss and dysfunc-
tion. J Biol Rhythms. 2011;26:423–33.
61. Scheer FA, Hilton MF, Mantzoros CS, Shea SA. Adverse metabolic and cardiovascular con-
sequences of circadian misalignment. Proc Natl Acad Sci U S A. 2009;106:4453–8.
62. Kelly MA, Rees SD, Hydrie MZ, Shera AS, Bellary S, O’Hare JP, Kumar S, Taheri S, Basit
A, Barnett AH. Circadian gene variants and susceptibility to type 2 diabetes: a pilot study.
PLoS One. 2012;7:e32670.
63. Woon PY, Kaisaki PJ, Braganca J, Bihoreau MT, Levy JC, Farrall M, Gauguier D. Aryl
hydrocarbon receptor nuclear translocator-like (BMAL1) is associated with susceptibility to
hypertension and type 2 diabetes. Proc Natl Acad Sci U S A. 2007;104:14412–7.
64. Karthikeyan R, Marimuthu G, Sooriyakumar M, BaHammam AS, Spence DW, Pandi-­
Perumal SR, Brown GM, Cardinali DP. Per3 length polymorphism in patients with type 2
diabetes mellitus. Horm Mol Biol Clin Invest. 2014;18:145–9.
65. Monteleone P, Tortorella A, Docimo L, Maldonato MN, Canestrelli B, De LL, Maj
M. Investigation of 3111T/C polymorphism of the CLOCK gene in obese individuals with
or without binge eating disorder: association with higher body mass index. Neurosci Lett.
2008;435:30–3.
66. Garaulet M, Corbalan MD, Madrid JA, Morales E, Baraza JC, Lee YC, Ordovas JM. CLOCK
gene is implicated in weight reduction in obese patients participating in a dietary programme
based on the Mediterranean diet. Int J Obes (Lond). 2010;34:516–23.
67. Ye D, Cai S, Jiang X, Ding Y, Chen K, Fan C, Jin M. Associations of polymorphisms in
circadian genes with abdominal obesity in Chinese adult population. Obes Res Clin Pract.
2016;10(Suppl 1):S133–41.
68. Yamaguchi M, Uemura H, Arisawa K, Katsuura-Kamano S, Hamajima N, Hishida A, Suma
S, Oze I, Nakamura K, Takashima N, Suzuki S, Ibusuki R, Mikami H, Ohnaka K, Kuriyama
N, Kubo M, Tanaka H. Association between brain-muscle-ARNT-like protein-2 (BMAL2)
gene polymorphism and type 2 diabetes mellitus in obese Japanese individuals: a cross-­
sectional analysis of the Japan Multi-institutional Collaborative Cohort Study. Diabetes Res
Clin Pract. 2015;110:301–8.
69. Corella D, Asensio EM, Coltell O, Sorli JV, Estruch R, Martinez-Gonzalez MA, Salas-­
Salvado J, Castaner O, Aros F, Lapetra J, Serra-Majem L, Gomez-Gracia E, Ortega-Azorin
C, Fiol M, Espino JD, Diaz-Lopez A, Fito M, Ros E, Ordovas JM. CLOCK gene variation
is associated with incidence of type-2 diabetes and cardiovascular diseases in type-2 dia-
betic subjects: dietary modulation in the PREDIMED randomized trial. Cardiovasc Diabetol.
2016;15:4.
70. Cardinali DP, Cano P, Jimenez-Ortega V, Esquifino AI. Melatonin and the metabolic syndrome:
physiopathologic and therapeutical implications. Neuroendocrinology. 2011;93:133–42.
71. Stamenkovic JA, Olsson AH, Nagorny CL, Malmgren S, Dekker-Nitert M, Ling C, Mulder
H. Regulation of core clock genes in human islets. Metabolism. 2012;61:978–85.
72. Pivovarova O, Gogebakan O, Sucher S, Groth J, Murahovschi V, Kessler K, Osterhoff M,
Rudovich N, Kramer A, Pfeiffer AF. Regulation of the clock gene expression in human adi-
pose tissue by weight loss. Int J Obes (Lond). 2016;40:899–906.
73. Samblas M, Milagro FI, Gomez-Abellan P, Martinez JA, Garaulet M. Methylation on the
circadian gene BMAL1 is associated with the effects of a weight loss intervention on serum
lipid levels. J Biol Rhythms. 2016;31:308–17.
74. Hardeland R. Melatonin and the pathologies of weakened or dysregulated circadian oscilla-
tors. J Pineal Res. 2017;62.
References 369

75. Cardinali DP. Ma Vie en Noir. Fifty years with melatonin and the stone of madness.
Switzerland: Springer; 2016.
76. Kozirog M, Poliwczak AR, Duchnowicz P, Koter-Michalak M, Sikora J, Broncel M. Melatonin
treatment improves blood pressure, lipid profile, and parameters of oxidative stress in patients
with metabolic syndrome. J Pineal Res. 2011;50:261–6.
77. Gonciarz M, Bielanski W, Partyka R, Brzozowski T, Konturek PC, Eszyk J, Celinski K,
Reiter RJ, Konturek SJ. Plasma insulin, leptin, adiponectin, resistin, ghrelin, and melatonin in
nonalcoholic steatohepatitis patients treated with melatonin. J Pineal Res. 2013;54:154–61.
78. Gonciarz M, Gonciarz Z, Bielanski W, Mularczyk A, Konturek PC, Brzozowski T, Konturek
SJ. The effects of long-term melatonin treatment on plasma liver enzymes levels and plasma
concentrations of lipids and melatonin in patients with nonalcoholic steatohepatitis: a pilot
study. J Physiol Pharmacol. 2012;63:35–40.
79. Gonciarz M, Gonciarz Z, Bielanski W, Mularczyk A, Konturek PC, Brzozowski T, Konturek
SJ. The pilot study of 3-month course of melatonin treatment of patients with nonalcoholic
steatohepatitis: effect on plasma levels of liver enzymes, lipids and melatonin. J Physiol
Pharmacol. 2010;61:705–10.
80. Romo-Nava F, Alvarez-Icaza GD, Fresan-Orellana A, Saracco AR, Becerra-Palars C, Moreno
J, Ontiveros Uribe MP, Berlanga C, Heinze G, Buijs RM. Melatonin attenuates antipsy-
chotic metabolic effects: an eight-week randomized, double-blind, parallel-group, placebo-­
controlled clinical trial. Bipolar Disord. 2014;16:410–21.
81. Modabbernia A, Heidari P, Soleimani R, Sobhani A, Roshan ZA, Taslimi S, Ashrafi M,
Modabbernia MJ. Melatonin for prevention of metabolic side-effects of olanzapine in
patients with first-episode schizophrenia: randomized double-blind placebo-controlled study.
J Psychiatr Res. 2014;53:133–40.
82. Tuomi T, Nagorny CL, Singh P, Bennet H, Yu Q, Alenkvist I, Isomaa B, Ostman B,
Soderstrom J, Pesonen AK, Martikainen S, Raikkonen K, Forsen T, Hakaste L, Almgren
P, Storm P, Asplund O, Shcherbina L, Fex M, Fadista J, Tengholm A, Wierup N, Groop L,
Mulder H. Increased melatonin signaling is a risk factor for type 2 diabetes. Cell Metab.
2016;23:1067–77.
83. American Psychiatric Association. Diagnostic and statistical manual of mental disorders. 5th
ed. 2015.
84. Campos CI, Nogueira CH, Fernandes L. Aging, circadian rhythms and depressive disorders:
a review. Am J Neurodegener Dis. 2013;2:228–46.
85. Milhiet V, Boudebesse C, Bellivier F, Drouot X, Henry C, Leboyer M, Etain B. Circadian
abnormalities as markers of susceptibility in bipolar disorders. Front Biosci (Schol Ed).
2014;6:120–37.
86. Wirz-Justice A. Diurnal variation of depressive symptoms. Dialogues Clin Neurosci.
2008;10:337–43.
87. Jagannath A, Peirson SN, Foster RG. Sleep and circadian rhythm disruption in neuropsychi-
atric illness. Curr Opin Neurobiol. 2013;23:888–94.
88. Bersani FS, Iannitelli A, Pacitti F, Bersani G. Sleep and biorythm disturbances in schizophre-
nia, mood and anxiety disorders: a review. Riv Psichiatr. 2012;47:365–75.
89. Lewy AJ, Kern HA, Rosenthal NE, Wehr TA. Bright artificial light treatment of a manic-­
depressive patient with a seasonal mood cycle. Am J Psychiatry. 1982;139:1496–8.
90. Lewy AJ, Nurnberger JI Jr, Wehr TA, Pack D, Becker LE, Powell RL, Newsome
DA. Supersensitivity to light: possible trait marker for manic-depressive illness. Am
J Psychiatry. 1985;142:725–7.
91. Cardinali DP, Pandi-Perumal SR, Brown GM. Sleep and circadian dysregulation in depres-
sive illness. Pharmacological implications. Clin Neuropsychiatry. 2011;8:321–38.
92. Soreca I. Circadian rhythms and sleep in bipolar disorder: implications for pathophysiology
and treatment. Curr Opin Psychiatry. 2014;27:467–71.
93. Ng TH, Chung KF, Ho FY, Yeung WF, Yung KP, Lam TH. Sleep-wake disturbance in interepi-
sode bipolar disorder and high-risk individuals: a systematic review and meta-analysis. Sleep
Med Rev. 2015;20:46–58.
370 8  Clinical Implications of the Timed Autonomic Nervous System

94. McCarthy MJ, Welsh DK. Cellular circadian clocks in mood disorders. J Biol Rhythms.
2012;27:339–52.
95. Dallaspezia S, Benedetti F. Chronobiological therapy for mood disorders. Expert Rev
Neurother. 2011;11:961–70.
96. Coogan AN, Thome J. Chronotherapeutics and psychiatry: setting the clock to relieve the
symptoms. World J Biol Psychiatry. 2011;12(Suppl 1):40–3.
97. Lewy AJ, Emens J, Jackman A, Yuhas K. Circadian uses of melatonin in humans. Chronobiol
Int. 2006;23:403–12.
98. Johansson C, Willeit M, Smedh C, Ekholm J, Paunio T, Kieseppa T, Lichtermann D, Praschak-­
Rieder N, Neumeister A, Nilsson LG, Kasper S, Peltonen L, Adolfsson R, Schalling M,
Partonen T. Circadian clock-related polymorphisms in seasonal affective disorder and their
relevance to diurnal preference. Neuropsychopharmacology. 2003;28:734–9.
99. Benedetti F, Dallaspezia S, Colombo C, Pirovano A, Marino E, Smeraldi E. A length poly-
morphism in the circadian clock gene Per3 influences age at onset of bipolar disorder.
Neurosci Lett. 2008;445:184–7.
100. Li SX, Liu LJ, LZ X, Gao L, Wang XF, Zhang JT, Lu L. Diurnal alterations in circadian
genes and peptides in major depressive disorder before and after escitalopram treatment.
Psychoneuroendocrinology. 2013;38:2789–99.
101. Karthikeyan R, Marimuthu G, Ramasubramanian C, Arunachal G, BaHammam AS, Spence
DW, Cardinali DP, Brown GM, Pandi-Perumal SR. Association of Per3 length polymorphism
with bipolar I disorder and schizophrenia. Neuropsychiatr Dis Treat. 2014;110:2025–30.
102. Vollmer LL, Strawn JR, Sah R. Acid-base dysregulation and chemosensory mechanisms in
panic disorder: a translational update. Transl Psychiatry. 2015;5:e572.
103. Huang F, Yang Z, Li CQ. The melatonergic system in anxiety disorders and the role of mela-
tonin in conditional fear. Vitam Horm. 2017;103:281–94.
104. Cardinali DP, Vidal MF, Vigo DE. Agomelatine: Its role in the management of major depres-
sive disorder. Clin Med Insights Psychiatry. 2012;4:1–23.
105. Liu J, Clough SJ, Hutchinson AJ, Adamah-Biassi EB, Popovska-Gorevski M, Dubocovich
ML. MT1 and MT2 melatonin receptors: a therapeutic perspective. Annu Rev Pharmacol
Toxicol. 2016;56:361–83.
106. Monti JM, BaHammam AS, Pandi-Perumal SR, Bromundt V, Spence DW, Cardinali
DP, Brown GM. Sleep and circadian rhythm dysregulation in schizophrenia. Prog
Neuropsychopharmacol Biol Psychiatry. 2013;43:209–16.
107. Mansour HA, Wood J, Logue T, Chowdari KV, Dayal M, Kupfer DJ, Monk TH, Devlin B,
Nimgaonkar VL. Association study of eight circadian genes with bipolar I disorder, schizoaf-
fective disorder and schizophrenia. Genes Brain Behav. 2006;5:150–7.
108. Lane JM, Vlasac I, Anderson SG, Kyle SD, Dixon WG, Bechtold DA, Gill S, Little MA, Luik
A, Loudon A, Emsley R, Scheer FA, Lawlor DA, Redline S, Ray DW, Rutter MK, Saxena
R. Genome-wide association analysis identifies novel loci for chronotype in 100,420 indi-
viduals from the UK Biobank. Nat Commun. 2016;7:10889.
109. Benson KL. Sleep in schizophrenia: pathology and treatment. Sleep Med Clin. 2015;10:49–55.
110. Gerbaldo H, Demisch L, Cardinali DP. Light exposure patterns in schizophrenia. Acta
Psychiatr Scand. 1992;85:94–5.
111. Ju P, Cui D. The involvement of N-methyl-D-aspartate receptor (NMDAR) subunit NR1 in the
pathophysiology of schizophrenia. Acta Biochim Biophys Sin (Shanghai). 2016;48:209–19.
112. Wulff K, Dijk DJ, Middleton B, Foster RG, Joyce EM. Sleep and circadian rhythm disruption
in schizophrenia. Br J Psychiatry. 2012;200:308–16.
113. Pasternak O, Westin CF, Dahlben B, Bouix S, Kubicki M. The extent of diffusion MRI mark-
ers of neuroinflammation and white matter deterioration in chronic schizophrenia. Schizophr
Res. 2015;161:113–8.
114. Morera-Fumero AL, Abreu-Gonzalez P. Role of melatonin in schizophrenia. Int J Mol Sci.
2013;14:9037–50.
115. Aston C, Jiang L, Sokolov BP. Microarray analysis of postmortem temporal cortex from
patients with schizophrenia. J Neurosci Res. 2004;77:858–66.
References 371

116. Zorn JV, Schur RR, Boks MP, Kahn RS, Joels M, Vinkers CH. Cortisol stress reactivity across
psychiatric disorders: a systematic review and meta-analysis. Psychoneuroendocrinology.
2016;77:25–36.
117. Tang Q, Song P, Xu L. The government’s role in regulating, coordinating, and standardiz-
ing the response to Alzheimer’s disease: anticipated international cooperation in the area of
intractable and rare diseases. Intractable Rare Dis Res. 2016;5:238–43.
118. Zdanys KF, Steffens DC. Sleep disturbances in the elderly. Psychiatr Clin North Am.
2015;38:723–41.
119. Duffy JF, Zitting KM, Chinoy ED. Aging and circadian rhythms. Sleep Med Clin.
2015;10:423–34.
120. American Geriatrics Society 2015. Updated Beers criteria for potentially inappropriate medi-
cation use in older adults. J Am Geriatr Soc. 2015;63:2227–46.
121. Cardinali DP, Golombek DA, Rosenstein RE, Brusco LI, Vigo DE. Assessing the efficacy of
melatonin to curtail benzodiazepine/Z drug abuse. Pharmacol Res. 2016;109:12–23.
122. Baandrup L, Fasmer OB, Glenthoj BY, Jennum PJ. Circadian rest-activity rhythms during
benzodiazepine tapering covered by melatonin versus placebo add-on: data derived from a
randomized clinical trial. BMC Psychiatry. 2016;16:348.
123. Clay E, Falissard B, Moore N, Toumi M. Contribution of prolonged-release melatonin and
anti-benzodiazepine campaigns to the reduction of benzodiazepine and Z-drugs consumption
in nine European countries. Eur J Clin Pharmacol. 2013;69:1–10.
124. Zeppenfeld DM, Simon M, Haswell JD, D’Abreo D, Murchison C, Quinn JF, Grafe MR,
Woltjer RL, Kaye J, Iliff JJ. Association of perivascular localization of aquaporin-4 with
cognition and Alzheimer disease in aging brains. JAMA Neurol. 2017;74:91–9.
125. Venkat P, Chopp M, Chen J. New insights into coupling and uncoupling of cerebral blood
flow and metabolism in the brain. Croat Med J. 2016;57:223–8.
126. MacLeod R, Hillert EK, Cameron RT, Baillie GS. The role and therapeutic targeting of
alpha-, beta- and gamma-secretase in Alzheimer’s disease. Future Sci OA. 2015;1:FSO11.
127. Hurtado-Alvarado G, Dominguez-Salazar E, Pavon L, Velazquez-Moctezuma J, Gomez-­
Gonzalez B. Blood-brain barrier disruption induced by chronic sleep loss: low-grade inflam-
mation may be the link. J Immunol Res. 2016;2016:4576012.
128. Collins O, Dillon S, Finucane C, Lawlor B, Kenny RA. Parasympathetic autonomic dysfunc-
tion is common in mild cognitive impairment. Neurobiol Aging. 2012;33:2324–33.
129. Ooms S, Ju YE. Treatment of sleep disorders in dementia. Curr Treat Options Neurol.
2016;18:40.
130. Riemersma-van der Lek RF, Swaab DF, Twisk J, Hol EM, Hoogendijk WJ, Van Someren
EJ. Effect of bright light and melatonin on cognitive and noncognitive function in elderly
residents of group care facilities: a randomized controlled trial. JAMA. 2008;299:2642–55.
131. Sirin FB, Kumbul DD, Vural H, Eren I, Inanli I, Sutcu R, Delibas N. Plasma 8-isoPGF2alpha
and serum melatonin levels in patients with minimal cognitive impairment and Alzheimer
disease. Turk J Med Sci. 2015;45:1073–7.
132. Allan CL, Behrman S, Ebmeier KP, Valkanova V. Diagnosing early cognitive decline-When,
how and for whom? Maturitas. 2017;96:103–8.
133. Daulatzai MA. Pharmacotherpy and Alzheimer’s disease: the M-Drugs (melatonin, minocy-
cline, modafinil, and memantine) approach. Curr Pharm Des. 2016;22:2411–30.
134. Martino Adami PV, Galeano P, Wallinger ML, Quijano C, Rabossi A, Pagano ES, Olivar N,
Reyes Toso C, Cardinali DP, Brusco LI, DoCarmo S, Radi R, Gevorkian G, Castaño EM,
Cuello AC, Morelli L. Worsening of memory deficit induced by energy-dense diet in a rat
model of early-Alzheimer’s disease is associated to neurotoxic Aβ species and independent
of neuroinflammation. Biochim Biophys Acta. 2017;1863:731–43.
135. Monti D, Ostan R, Borelli V, Castellani G, Franceschi C. Inflammaging and omics in human
longevity. Mech Ageing Dev. 2016.
136. Minciullo PL, Catalano A, Mandraffino G, Casciaro M, Crucitti A, Maltese G, Morabito N,
Lasco A, Gangemi S, Basile G. Inflammaging and anti-inflammaging: the role of cytokines
in extreme longevity. Arch Immunol Ther Exp (Warsz). 2016;64:111–26.
372 8  Clinical Implications of the Timed Autonomic Nervous System

137. Ralph MR, Lehman MN. Transplantation: a new tool in the analysis of the mammalian hypo-
thalamic circadian pacemaker. Trends Neurosci. 1991;14:362–6.
138. Carrillo-Vico A, Lardone PJ, Alvarez-Sanchez N, Rodriguez-Rodriguez A, Guerrero
JM. Melatonin: buffering the immune system. Int J Mol Sci. 2013;14:8638–83.
139. Siegel R, Naishadham D, Jemal A. Cancer statistics, 2013. CA Cancer J Clin. 2013;63:11–30.
140. Stevens RG, Zhu Y. Electric light, particularly at night, disrupts human circadian rhythmicity:
is that a problem? Philos Trans R Soc Lond B Biol Sci. 2015;370.
141. Straif K, Baan R, Grosse Y, Secretan B, El GF, Bouvard V, Altieri A, Benbrahim-Tallaa
L, Cogliano V. Carcinogenicity of shift-work, painting, and fire-fighting. Lancet Oncol.
2007;8:1065–6.
142. Hill SM, Belancio VP, Dauchy RT, Xiang S, Brimer S, Mao L, Hauch A, Lundberg PW,
Summers W, Yuan L, Frasch T, Blask DE. Melatonin: an inhibitor of breast cancer. Endocr
Relat Cancer. 2015;22:R183–204.
143. Lin X, Chen W, Wei F, Ying M, Wei W, Xie X. Night-shift work increases morbidity of breast
cancer and all-cause mortality: a meta-analysis of 16 prospective cohort studies. Sleep Med.
2015;16:1381–7.
144. Valenzuela FJ, Vera J, Venegas C, Munoz S, Oyarce S, Munoz K, Lagunas C. Evidences
of polymorphism associated with circadian system and risk of pathologies: a review of the
literature. Int J Endocrinol. 2016;2016:2746909.
145. Altman BJ. Cancer clocks out for lunch: disruption of circadian rhythm and metabolic oscil-
lation in cancer. Front Cell Dev Biol. 2016;4:62.
146. Cardinali DP, Escames G, Acuña-Castroviejo D, Ortiz F, Fernández-Gil B, Guerra Librero A,
García-López S, Shen Y, Florido J. Melatonin-induced oncostasis. Mechanisms and clinical
relevance. Integr Oncol. 2016;S1:006. doi:10.4172/2329-6771.S1-006.
147. Blask DE, Dauchy RT, Sauer LA, Krause JA, Brainard GC. Growth and fatty acid metabolism
of human breast cancer (MCF-7) xenografts in nude rats: impact of constant light-induced
nocturnal melatonin suppression. Breast Cancer Res Treat. 2003;79:313–20.
148. Blask DE, Brainard GC, Dauchy RT, Hanifin JP, Davidson LK, Krause JA, Sauer LA, Rivera-­
Bermudez MA, Dubocovich ML, Jasser SA, Lynch DT, Rollag MD, Zalatan F. Melatonin-­
depleted blood from premenopausal women exposed to light at night stimulates growth of
human breast cancer xenografts in nude rats. Cancer Res. 2005;65:11174–84.
149. Dauchy RT, Hoffman AE, Wren-Dail MA, Hanifin JP, Warfield B, Brainard GC, Xiang S,
Yuan L, Hill SM, Belancio VP, Dauchy EM, Smith K, Blask DE. Daytime blue light enhances
the nighttime circadian melatonin inhibition of human prostate cancer growth. Comp Med.
2015;65:473–85.
150. Wang X, Ji A, Zhu Y, Liang Z, Wu J, Li S, Meng S, Zheng X, Xie L. A meta-analysis includ-
ing dose-response relationship between night shift work and the risk of colorectal cancer.
Oncotarget. 2015;6:25046–60.
151. Ma Z, Yang Y, Fan C, Han J, Wang D, Di S, Hu W, Liu D, Li X, Reiter RJ, Yan X. Melatonin
as a potential anticarcinogen for non-small-cell lung cancer. Oncotarget. 2016;7:46768–84.
152. Howell D, Oliver TK, Keller-Olaman S, Davidson JR, Garland S, Samuels C, Savard J, Harris
C, Aubin M, Olson K, Sussman J, MacFarlane J, Taylor C. Sleep disturbance in adults with
cancer: a systematic review of evidence for best practices in assessment and management for
clinical practice. Ann Oncol. 2014;25:791–800.
153. Dahiya S, Ahluwalia MS, Walia HK. Sleep disturbances in cancer patients: underrecognized
and undertreated. Cleve Clin J Med. 2013;80:722–32.
154. Chen WY, Giobbie-Hurder A, Gantman K, Savoie J, Scheib R, Parker LM, Schernhammer
ES. A randomized, placebo-controlled trial of melatonin on breast cancer survivors: impact
on sleep, mood, and hot flashes. Breast Cancer Res Treat. 2014;145:381–8.
155. Dickerson SS, Connors LM, Fayad A, Dean GE. Sleep-wake disturbances in cancer patients:
narrative review of literature focusing on improving quality of life outcomes. Nat Sci Sleep.
2014;6:85–100.
156. Harris B, Ross J, Sanchez-Reilly S. Sleeping in the arms of cancer: a review of sleeping dis-
orders among patients with cancer. Cancer J. 2014;20:299–305.
References 373

157. Palesh OG, Roscoe JA, Mustian KM, Roth T, Savard J, Ancoli-Israel S, Heckler C, Purnell
JQ, Janelsins MC, Morrow GR. Prevalence, demographics, and psychological associations of
sleep disruption in patients with cancer: University of Rochester Cancer Center-Community
Clinical Oncology Program. J Clin Oncol. 2010;28:292–8.
158. Jia Y, Lu Y, Wu K, Lin Q, Shen W, Zhu M, Huang S, Chen J. Does night work increase the risk
of breast cancer? A systematic review and meta-analysis of epidemiological studies. Cancer
Epidemiol. 2013;37:197–206.
159. Armstrong TS, Shade MY, Breton G, Gilbert MR, Mahajan A, Scheurer ME, Vera E, Berger
AM. Sleep-wake disturbance in patients with brain tumors. Neuro Oncol. 2017;19:323–35.
160. Casault L, Savard J, Ivers H, Savard MH, Simard S. Utilization of hypnotic medication in the
context of cancer: predictors and frequency of use. Support Care Cancer. 2012;20:1203–10.
161. Thekdi SM, Trinidad A, Roth A. Psychopharmacology in cancer. Curr Psychiatry Rep.
2015;17:529.
162. Dallmann R, Okyar A, Levi F. Dosing-time makes the poison: circadian regulation and phar-
macotherapy. Trends Mol Med. 2016;22:430–45.
Epilogue

The history of the ANS starts in the Western world with the ancient Greeks, who
coined the idea that the body was divided into two systems (somatic and autonomic)
[1]. It was Galen (second century AD) who followed the vagus into the chest and
abdominal cavities, documenting its communications with the viscera. Galen iden-
tified the sympathetic trunks as crossing the ribs, connecting with the thoracic and
lumbar spinal cord, and continuing further to communicate with organs within the
body cavity via the plexuses. Galen believed that the nerves acted as pipes, allowing
the flow of “animal spirits” to pass among the organs.
He conceived the notion of “sympathy,” i.e., when a change in the condition of
one organ or part of the body occurs, this in turn causes another organ or portion of
the body to react or alter its function. Sympathy remained a vague notion for centu-
ries. It explains the cooperation or coordination of organs, such as an irritation of the
stomach that produces syncope or seizures, as it is transmitted by the brain and
nerves to the heart. Galen also postulated a humoral sympathy through the blood
vessels, such as the relationship between the pregnant uterus and the mammary
glands.
Indeed, sympathy is the major inspiring idea for this book. The bio-psycho-­
social-ecological nature of the individual can only be embraced if the enlarged and
timed view of the ANS is followed, as outlined in this book.
It is futile to approach the disease from the statistical point of view of evidence-­
based medicine, as the popular paradigm claims nowadays. The physician must
know in depth the individual reality of the person he is attempting to cure and this
part of medicine is forgotten because of the inappropriate pressure of public health
administrators and pharmaceutical companies. The enlarged and timed view of the
ANS holds that all body systems are dependent and affected by the actions of others
in a multicellular organization.
In summary, the ANS innervates the entire human body, and is involved in the
regulation of every single organ in the body. Thus, perturbations in autonomic func-
tion account for everything from abnormalities in pupillary function to gastropare-
sis, intestinal dysmotility, diabetic diarrhea, genitourinary dysfunction, amongst
others. Know autonomic function and you will know the whole of medicine! [2].

© Springer International Publishing AG 2018 375


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1
376 Epilogue

References
1. Oakes PC, Fisahn C, Iwanaga J, DiLorenzo D, Oskouian RJ, Tubbs RS. A history of the auto-
nomic nervous system: part I: from Galen to Bichat. Childs Nerv Syst. 2016;32:2303–8.
2. Vinik AI, Erbas T.  22  – Diabetic autonomic neuropathy. In: Buijs RM, Swaab D, editors.
Handbook of clinical neurology. Autonomic nervous system. New  York: Elsevier; 2013,
p. 279–94.
Index

A Amygdala, 7, 208, 247, 249


Abdominal pain, somatosensory and limbic anatomical localization of, 252
pathways, 86 bilateral ablation of, 258
Acetylcholine (ACh), 3, 168 electrical stimulation of, 253, 254, 258
receptors, 74–75 limbic system and, 252–257
Acid neutralization, 163 Amyloid β (Aβ), 353–354
Acquired immunity, 147 Amyloidotic autonomic failure, 299
Acute bacterial infections, 100 ANS. See Autonomic nervous system (ANS)
Acute hepatic porphyrias, 306 Anterior cingulate cortex, 7
Adenohypophyseal hormone release, 180 Anteroventral periventricular nucleus
Adipocytes, 213 (AVPV), 182, 236
Adrenergic leukocytosis, 156 Antigens, 202
Adrenergic neurons, 58 Anti-inflammatory pathway, 151
Adrenergic neurotransmission, 61, 63 Aortic depressor nerve, 125–126
Adrenergic receptors, 75–76 Apneustic center, 133
Adrenocorticotropic hormone (ACTH) Arcuate nucleus (ARC), 182, 209, 211
adrenal sympathetic nerves, 192 Arginine vasopressin
circadian pattern of, 186 (AVP), 177, 198, 225, 226
CRH and, 198 Atrial natriuretic peptide (ANP), 226, 227
elevation of, 190 nocturnal levels of, 228
hypothalamic neurons, 219 nyctohemeral fluctuation of, 226
minimal secretion of, 189 secretion, 221
plasma cortisol, 186 sleep apnea, 225
Advanced sleep phase syndrome (APSP), 274 Auditory cortex, 141
Afferent autonomic pathways, 83 Autoimmune autonomic ganglionopathy, 299
Aging Automatic bladder, 91
autonomic dysfunction, 305–306 Autonomic behaviors, 8
brain, 353 Autonomic dysfunction
CNS, 353 with aging, 305
consequences of, 351 amyloidotic autonomic failure, 299–300
Aldosterone, constitutive secretion of, 227 autoimmune autonomic
Allodynia, 170 ganglionopathy, 299
Allostasis, 194, 195 cardiovascular disorders, 303–305
Alzheimer’s disease, 319 chronic fatigue syndrome, 307
brain aging, 349 chronic pain, 307
melatonin, 355 classifications, 292–295
Ambiguous nucleus (AN), 127, 133 clinical signs, 288
Ambiome, 13 diabetes mellitus, 298–299

© Springer International Publishing AG 2018 377


D.P. Cardinali, Autonomic Nervous System, DOI 10.1007/978-3-319-57571-1
378 Index

Autonomic dysfunction (cont.) Bötzinger complex, 134, 135


drug-induced, 307 Braak model of neurodegeneration, 298
fibromyalgia, 307–309 Bradycardia, 132
function tests, 290 Brain
Guillain–Barré syndrome, 306 aging, 349
hereditary autonomic neuropathies, leptin action on, 214
306–307 Brain natriuretic peptide (BNP), 226
hyperthermia, 309–310 Brainstem, 7, 144–171
hypothermia, 309 afferent control, 119
insomnia, 300–301 ANS and, 114–120
paraneoplastic, 299 cerebellum and autonomic posture,
pathophysiological classification, 292 138–144
periodic limb movement, 302 cholinergic projections, 124
peripheral neuropathies, 293–296 cranial nerves and column organization, 115
REM sleep, 302–303 digestive function, 161
restless legs syndrome, 302 histaminergic projections, 125
semiological aspects, 288–292 neurohumoral responses, 127
sleep-related breathing disorder, 301–302 respiratory centers, 134
spinal cord injuries, 307 reticular formation, 121–123
symptoms, 289 sleep deprivation, 157
α-synucleinopathies, 296–298 spider web neuron, 121
Autonomic nerve fibers, 24-h rhythms
neurochemical code, 70 in cardiovascular control, 125–132
Autonomic nervous system (ANS) in gastrointestinal function, 160–171
autonomic posture, 10 in immune response, 144–159
and Brainstem, 114 in physiological function, 120–125
functional neuroimaging, 268–271 in respiratory control, 133–138
hierarchical organization of, 6, 7 visceral and cutaneous afferents, 120
homeostasis control and, 2 visceral sensory input, 117
neuronal organization of, 14–17 Branchial motor neurons. See Visceral motor
noradrenergic vs adrenergic responses, 58 neurons
physiological programs, 316 Breathing, regulatory system of, 133
trigger area, 15 Brown adipose tissue (BAT), 231, 233
Autonomic posture, 9–13, 138 Bulbopontine, 9
Autonomic reflex, neurons, 62

C
B Cancer, 361
Baroreceptor, 290 Cardiac activity, modulation of, 130
Baroreflex sensitivity, 304 Cardiac- and respiratory-related rhythms, 131
Basal ganglia, limbic system and, 262–268 Cardiorespiratory homeostasis, 137
Basic rest–activity cycle (BRAC), 33, 161, Cardiovascular control, 24-h rhythms in, 125
233 Cardiovascular disorders, autonomic
Basolateral portion, 254 dysfunction, 303
Benzodiazepines (BZDs), 351–352 Cardiovascular homeostasis, 224
Bile secretion, cholesterol concentrations in, Cardiovascular risk factors, prevalence of, 326
169 Carotid body, 116, 135
Bmal1, 129 Caudal, 116
Body posture, 9 Central autonomic neural network, 114
Body temperature regulation, 228, 231 Central nervous system (CNS), aging, 353
Bone formation, ANS, 106–109 Central nucleus, 254
Bone resorption, 188 Cerebellar cortex, 142
Index 379

Cerebellar nuclei, 139 brain aging, 349–361


Cerebellar–hypothalamic projections, 138 cancer, 361–365
Cerebellum, 8 chronobiological treatment, 328–332
motor and autonomic projection, 139 chronodisruption, 319, 320, 327–328
posture, 138 jet lag, 321–324
Cerebrocerebellum, 142 mental illnesses, 339–341
Chemosensory mechanisms, 346 mood disorders, 341–347
Cholesterol concentrations, in bile secretion, 169 schizophrenia, 347–349
Cholinergic neurotransmission, 61, 63, 72 metabolic syndrome, 332–339
Cholinergic system, 123 shift-work disorder, 324–326
Chronic fatigue syndrome, 307 Diabetes mellitus, autonomic
Chronic pain, autonomic dysfunction, 307 dysfunction, 298–299
Chronobiology, development of, 2, 328 Dopamine (DA), 213, 228
Chronodisruption, 319, 320, 327 Dopaminergic system, 122
Chronotypes, 24-h rhythms and emotion, 271 Dorsal column, 118
Circadian clock, 2, 129, 218 Dorsal respiratory group (DRG), 133, 134
Circadian pacemaker, 161 Dorsomedial hypothalamus
Circadian rhythms, 157, 214, 273 (DMH), 209, 212, 231, 233
Climbing fibers, 143 Dreaming, 51–52
Clock genes, 21 Drug-induced autonomic dysfunction, 307
Cocaine- and amphetamine-regulated Dynorphin, 183
transcript (CART), 209, 211 Dysautonomia, peripheral neuropathies, 293
Cognitive circuit, 266
Cognitive memory, 255
Complex spikes, 144 E
Conscious perception, 84, 95 Effector organs, ANS effects, 64–65
Coping responses, during stress, 195 Electroencephalography (EEG), 31, 32, 278
Corticomedial amygdala, 254 Emotion
Cortico-striatal circuits, 267 external expression of, 258
Corticotropin-releasing hormone (CRH), 198 limbic system and, 257–262
hypersecretion, 190 Emotional circuit, 256
Cortisol, 182 Emotional illnesses, 339
constitutive secretion of, 186 Emotionality, limbic system in, 246–252
rhythmicity of, 186 Emotional memory, 255
Cytokines Endocrine, 153
with activity on sleep, 50 Engram, 276
secretion pattern, 150 Enkephalinergic interneurons, 77
Enteric ANS, 95–102
enteric glia, 96
D extrinsic innervation, 95
Declarative memory, 275 Entero-endocrine cells (EE), 165
Defecation, 89, 92 Epinephrine, 4
Defense behavior, 190–203 Estrogens, 213
Deiodinase type 2 enzyme, 31 Excitatory postsynaptic potential (EPSP), 15
Delayed orthostatic hypotension, 305 Extended amygdala, 254
Delayed sleep phase syndrome (DSPS), 274
Dementia with Lewy bodies, 297, 298
Demyelination process, neuropathies, 295 F
Dentate gyrus, 277 Familial dysautonomia, 306
Depression, 340, 347. See also Mental illness, Fastigial nucleus, 138
desynchronized ANS Feeding activity, 24-h rhythm in, 203, 204
Desynchronized ANS, 320–321, 341–349 Fibromyalgia, 307
380 Index

Fight-or-flight response, 192 Growth hormone (GH), 181


Flight dysrhythmia. See Jet lag Growth hormone-releasing
Flocculonodular lobe, 142 hormone (GHRH), 184
Follicle-stimulating hormone (FSH), 201 Guillain–Barré syndrome, 296, 306
Food, as circadian synchronizer, 216 Gut wall, 96
Food-entrained oscillators (FEO), 165, 217
Functional magnetic resonance imaging
(fMRI), 324 H
Hamster, locomotor activity rhythm, 26
Heart rate variability (HRV), 131, 291
G Helicobacter pylori, 164
GABAergic neurons, 128 Hereditary autonomic neuropathies, 306
GABAergic Purkinje cells, 143 Hippocampal neurogenesis, 200
Gallbladder, 169 Hippocampus, 198, 276
Ganglion cells, 26 Histaminergic system, 124
Ganglion neurons, 70 Homeostasis
Gastric acid secretion, 98 and ANS, 2–5
Gastric emptying, 164 biological rhythms, 20–31
Gastric motor function, 167 Homologies, 265
Gastroesophageal reflux, 161, 162 spinal autonomic reflexes, 87–89
Gastrointestinal dysmotility, 306 Humoral immunity, 147
Gastrointestinal function, 24-h rhythms in, 160 Hyperadrenergic postural tachycardia
Gastrointestinal tract motility, 166 syndrome, 304
Gastroparesis, prevalence of, 298 Hypertension, 303
General adaptation syndrome, 190 Hyperthermia, 309
General somatic afferent column, 116 Hypophysiotropic area, 177, 179
General somatic sensitivity, 115 Hypotension, orthostatic, 296–298
General visceral afferent columns, 116 Hypothalamic GnRH pulse generator, 182
General visceral motoneurons, 115 Hypothalamic heating, 232
General visceral motor column, 116 Hypothalamic nuclei, 176, 183, 231
General visceral sensitivity, 115 Hypothalamic–pituitary–thyroid axis, 187
Genetic predispositions, 339, 356 Hypothalamus, 7, 180–190, 203–235
GH release, 184, 185 autonomic components, 176–180
Ghrelin, 210 behavioral and biological aspects, 191
Globus pallidus, 262, 264 food behavior, 205
Glucocorticoids, 198 gonadotropin release, 183
neurogenesis, 200 hormonal rhythm, 182
secretion of, 219 lateral hypothalamic area, 177
Glutamate (Glu), 23 medial zone, 177
Glutamatergic neurons, 127 motivational components, 176
Glycemia, 209 neuroendocrine components, 176
Glymphatic dysfunction, 353 neuroendocrine connections, 179
Glymphatic system, 52–54 neuroendocrine profiles, 193
GnRH neurons, 235, 236 neuropeptide perfusion studies, 189
Golgi type I neurons, 14 nuclei, 177
Golgi type II neurons, 14, 177 orexigenic and anorexigenic circuits, 208
Gonadal steroids, 201 paraventricular nucleus of, 197
Gonadotropic axis rhythms, 235 periventricular area, 177
Gonadotropin-inhibiting hormone (GnIH), 236 plasma LH concentration, 237
neurons, 183 plasma testosterone levels, 238
pulses, 181–182 reactive homeostasis,
G proteins, 73 defense behavior, 190
Granule cells, 143 sagittal view of, 178
Index 381

sexual and maternal behavior, 235–244 Limbic system, 7


24-h rhythms and amygdala, 252
in body temperature control, 228–235 anatomical composition of, 249
in food intake, energy storage, and ANS, functional neuroimaging, 268
metabolism, 203–221 and basal ganglia, 262
in neuroendocrine function, 180–190 brain function, 246
in plasma osmolality and intravascular chronotypes, 24-h rhythms and emotion,
volume, 221–228 271–274
Hypothermia, 309 components, 248
Hypoxemia, 301 cortical portion, 247
and emotion, 257
in emotionality, motivation, learning, and
I memory, 246
IL-6, 213 structures of, 248
Immune cells, intercellular interactions of, 148 subcortical portion, 247
Immune response, 24-h rhythms in, 144 24-h rhythms and learning and memory,
Immunity, 139, 147 274–283
Inferior esophageal sphincter (IES), 163 Liver glycogenolysis, 191
Inflammatory bowel disease, 170 Long-term memory, 275
Inhibitory postsynaptic potential (IPSP), 15 Long-term potentiation (LTP), 278
Insomnia, 300, 302, 332, 342 Low-grade inflammation, 156
Insular cortex, 8, 137, 270 Luteinizing hormone (LH), 201
Insulin, 210, 215 Lymphocytes, 147, 148
International Agency for Research on Cancer,
325
Interoception, 84, 85 M
Intestinal disaccharidases, 169 Mania, 340
Intravascular volume, 24-h rhythms in, 221 Master of sleep, 48
Ionotropic transmission, 71–73 Maternal behavior, hypothalamus, 235
Irritable bowel syndrome, 170 MC-4 receptors, 210
Medial preoptic area (mPOA), 235
Median preoptic nucleus (MnPOn), 223
J Melanin-concentrating hormone
Jet lag, 321, 322 (MCH), 209, 212
Melatonin, 25, 27, 154, 164
Alzheimer’s disease, 355
K and brain inflammaging, 360, 361
Kisspeptin (Kiss1), 183 chronobiological treatment, 328
KNDy neurons, 183 circadian rhythm, 28
constitutive secretion of, 186
drinking water, 30
L effects in metabolic syndrome, 339
Lamina terminalis, 223 immunological effects of, 361
Lateral hypothalamic area (LHA), 177, 178, nocturnal release of, 314
220 nocturnal, secretion, 30
Lateral hypothalamus, 256 overnight pulse, 30
Lateral parabrachial nucleus (LPB), 232 phase-and amplitude-altering effect, 28
Learning plasma concentration of, 186
limbic system in, 246 secretion, 23, 28, 30, 31, 35
24-h rhythms, 274 sleep-promoting effects of, 347
Leptin, 210, 213–215 synthesis control, 28
Light therapy, 30 treatment with, 216
Limbic gyrus, 247 Memantine, 356
382 Index

Memory Neuropeptides, 69, 70


limbic system in, 246 Neuropeptide transmitters, 52
24-h rhythms, 274 Neuropeptide Y (NPY), 190, 199, 210
Menarche, 30 Neurotransmitter norepinephrine
Mental illnesses, desynchronized ANS, 339 (NE), 3
mood disorders, 341 Neurotransmitters, 264
schizophrenia, 347 Nicotinic acetylcholine receptors, 74
symptoms of, 341 NO synthase (NOS) isoforms, 67
Mesencephalon, 209 Nocturnal gastroesophageal reflux, 163
Mesocortical circuit, 267 Nocturnal sleep, 215
Mesolimbic circuit, 267 Non-REM (NREM)
Metabolic mechanisms, 136 sleep, 32, 40, 43, 49, 53, 130, 131
Metabolic syndrome, 332 BP decreases during, 129
circadian alterations in organs/tissues, 338 chemosensitivity, 136
melatonin effects in, 339 episodic memory, 280
obesity-induced changes, 333 gastric engine cycle, 167
Metabotropic transmission, 71–73 humoral biological marker, 228
Microbiome, 13 immune response, 147
Microneurography, 291 sleep pressure and, 158
Migrating motor complex (MMC), 166–169 sympathetic nerve activity, 132
Mild cognitive impairment (MCI), 355 24-h rhythms in, 160
Mitochondrial dysfunction, 358 Noradrenergic system, 120
MM3-LN tumors, 201 Norepinephrine (NE), 74, 303
Mononeuropathies, 295 catabolism, 75
Mood disorders, 347 metabolism of, 76
Mossy fibers, 142, 143 synthesis, 75
Motivation, limbic system in, 246 Nucleotides, purinergic, 69
Motor circuit, 266 Nucleus accumbens, 247, 249, 266–268
Motor neurons, 3 Nucleus of the solitary tract (NTS), 116–118,
Motor nuclei, 116 180, 209, 222
Müller’s maneuver, 290
Multiple system atrophy, 297
Muscarinic acetylcholine receptors, 74 O
Muscle tone, 262 Obesity, and neurodegeneration, 357
Obsessive–compulsive disorder, 266
Obstructive sleep apnea (OSA), 300–302
N Orbitofrontal cortex, 137
Nasogastric recording, 164 Orexin, 38–40, 47
Nerve fibers, 16 Orthostatic hypotension, 296–298
Neural circuitry mapping techniques, 117 OSA. See Obstructive sleep apnea (OSA)
Neural regulation of cardiovascular Osmoreceptors, 222
function, 125 Oxidative stress, 353
Neurobehavioral cascade, 193 Oxytocin-releasing neurons, 239
Neurodegeneration, 357, 359
Neuroendocrine communication, 102–106
Neuroendocrine function, 179, 180 P
Neuroendocrine–immune Pancreatic juice, 169
mechanisms, 12, 145 Panic disorder, 346
Neuroendocrine profiles, 193 Papez circuit, 249
Neurokinin B, 183 Parabrachial nucleus
Neuropathies (PBN), 118, 133
demyelination process, 295 Paradoxical fear, 192
hereditary autonomic, 306 Paraneoplastic autonomic dysfunction, 299
peripheral, 293 Parasympathetic ganglionic neurons, 3
Index 383

Parasympathetic nerve fibers, 5 R


Parasympathetic nervous system, 58, 63 Rapid eye movement (REM) sleep, 31, 32, 36,
Parasympathetic system, 5 40, 48, 52, 130, 131
Parasympatholytic drugs, 94 BP decreases during, 129
Parasympathomimetic drugs, 94 cholinergic activation, 281
Paraventricular nucleus episodic memory, 280
(PVN), 177, 179, 197, flow of urine, 225
198, 210, 211, 220 GABAergic control, 42
Paravertebral sympathetic gastric engine cycle, 167
ganglia, 59, 60, 104, 107 humoral biological marker, 228
Parietal–temporal–occipital hypothetical circuitry, 41
association cortex, 258 immune response, 147
Parkinson’s disease, 297 with irregular breathing patterns, 132
PBN. See Parabrachial nucleus (PBN) limbic activity during, 132
Penile erection, 94, 95 PGO (cholinergic) wave, 51
Pepsin, 165 predominance of, 137
Peptidergic co-transmission, ANS, 63 prevalence, 48
Periodic limb movement respiratory rhythm, 137
disorder, in sleep, 302 24-h rhythms in, 160
Peripheral neuropathies, wakefulness and, 43
dysautonomia, 293 Rapid eye movement (REM) sleep behavior
Peripheral sensors, 135 disorder, 302
Periphery of ANS, 58–63 Reactive homeostasis, 2
cholinergic/adrenergic Reflexive/procedural memory, 275
neurotransmission, 63–79 Relay nuclei, 116
peptidergic co-transmission, 63 Renal sympathetic nerves, 225
Peristaltic reflex, 101 Renin–angiotensin–aldosterone system, 224
Phagocytes, 147 Renin, constitutive secretion of, 227
Phase maps, 20, 21, 321, 322 Reproductive process, seasonality in, 29, 30
Physiological function, 24-h rhythms in, 120 Respiratory centers, 133
Plasma osmolality, 24-h rhythms in, 221 Respiratory control
Polysomnography (PSG), 32, 42, 170, 189, 24-h rhythms in, 133
302 wakefulness, 136
Postural reflexes, 10 Restless legs syndrome, 302
Postural tachycardia syndrome, 304 Reticular formation, 114, 118, 120–124
Predictive homeostasis, 2 Rhinencephalon, 247
Preeclampsia, 305 Rostral, 116
Prefrontal control systems, 257
Preganglionic fibers, 58–60
Preoptic area (POA), 198, 232, 233 S
Prevertebral ganglia, 58, 61 Saliva, 186
Prevertebral sympathetic ganglia, 104 Saliva melaton, 187
Projection circuits, 14 Salivary flow, 163
Prolactin (PRL), 239 Schizophrenia, 347
constitutive secretion of, 185 Second-order neurons, 126
nocturnal increase in, 185 Second-order sensory neurons, 114
plasma levels, 185 Secretomotor functions, 289
secretion, 122, 184 Sensory autonomic neurons, 79–87
Pro-opiomelanocortin Serotoninergic system, 123
(POMC), 197, 209, 210 Sexual arousal, physical expression of, 95
Pure autonomic failure, 297 Sexual behavior, hypothalamus, 235
Purkinje cell, 142, 143 Sexual responses, spinal autonomic
Pyrogen production, 234 reflexes, 89
PYY 3-36, 210 Shift-work disorder, 324
384 Index

Short-term memory, 279 core and shell regions, 22


Shy–Drager syndrome, 297 in daylight conditions, 314
Sinus nerve, 126 deterioration, 359
Sirtuin 1, 218 electrical activity, 23, 34
Sleep integrity, 21
cytokines with activity, 50 light response, 35
disturbances, 274 neural degeneration, 359
neurophysiology, 36–45 neuron, 23, 28, 34
and glymphatic system, 52 neuronal cell bodies, 22
physiological states, 46–47 synchronizer, 27
premortem statuses of, 37 Swallowing, 161–163
responsible for, 32 Sympathetic nerve activity, 132
and wake cycle, 31–35 Sympathetic nerve fibers, 5
Sleep apnea, 132 Sympathetic nervous system, 58
Sleep deprivation, 157 Sympathetic neuroimaging, 291, 292
Sleep disorders, autonomic dysfunction Sympathetic rhythms, 131
insomnia, 300 Sympathoexcitation, 192
periodic limb movement, 302 Sympathomimetic drugs, 94
REM sleep, 302 α-Synucleinopathies, 296
restless legs syndrome, 302
sleep-related breathing disorder, 301
Sleep homeostatic pressure, 35 T
Sleep-related breathing disorder, 301 T cells, phases of, 149
Slow-wave sleep, 36, 45, 154, 155 T helper (Th) lymphocytes, 149, 152
Somatic motor neurons, 3, 6–10, 90, 114, 116 T regulatory (Treg), 149
Somatostatin, 185 Temperature-sensing receptors, 231
Special somatic afferent column, 116 Th1/Th2 cytokine, 154
Special somatic sensitivity, 115 Thalamic stimulation, 36
Special visceral afferent columns, 116 Thalamus, 209
Special visceral motoneurons, 115, 116 Thermogenesis, 231
Special visceral sensitivity, 115 Thermoregulation, 229
Spider web circuits, 15 Thermosensory signals, 232
Spinal autonomic reflexes, 87, 89 Thoracolumbar, 59
Spinal cord, 7 Thyroid-stimulating hormone (TSH), 187
Spinal cord injuries, autonomic dysfunction, Thyrotropin-releasing hormone (TRH), 187
307 Tooth eruption, 108
Spinal motor neurons, 136 Tourette’s syndrome, 266
Spinal motor reflexes, 87 Transient receptor potential (TRP) channels,
Spinocerebellum, 142 231
Stomach, functions of, 164 Trefoil protein 1, 164
Stress Treg cells, 149
coping responses during, 195 Trine brain concept, 251
leucocytes, 156 Tuberoinfundibular system, 122
Striatum, 123 Tuberomammillary nucleus (TMN), 38, 124,
Substantia nigra, 262 125
Sudomotor functions, 289 Tumescence, 95
Superior cervical ganglion (SCG), 87, 89, 93 24/7 Society, 314–319
neuroendocrine relevance, 104 24-h rhythms
relevance of, 104 in body temperature control, 228
territory, 202 in cardiovascular control, 125
Suprachiasmatic nuclei (SCN), 22–24, 26, 29, and emotion, 271
34, 203, 273, 314, 315 in food intake, energy storage, and
circadian apparatus, 22 metabolism, 203
circadian effects, 22 in gastrointestinal function, 160
Index 385

in immune response, 144 Ventral striatum, 264


and learning and memory, 274 Ventrolateral preoptic area (VLPO), 35, 39,
melatonin secretion, 188 40, 49
in neuroendocrine function, 180 Ventromedial hypothalamus (VMH), 207, 208,
in physiological function, 120 212, 270
in plasma osmolality and intravascular Ventromedial nucleus (VMN), 118, 212
volume, 221 Vermis, 140
in respiratory control, 133 Vertebrate motor neurons, 3
Type 2 diabetes mellitus (T2DM), 337. See Vestibulocerebellum, 142
also Diabetes mellitus Visceral afferents, 62, 80–83
metabolic syndrome, 332 Visceral motor neurons, 3
Visceral preganglionic neurons, 3
Volume clearance, 163
U Voluntary control, 136
Ultradian rhythms, 169
Urination, spinal autonomic reflexes, 89–95
W
Wakefulness, 39, 45–51
V Wallerian degeneration, 295
Vagal nerve, 82, 97 White adipose tissue (WAT), 233
Valsalva maneuver, 269, 289 Working memory, 257, 266, 273, 275, 282
Varicose fibers, 144
Vasomotor changes, 231
Ventral respiratory group, 133 Z
Ventral respiratory group (VRG), 133, 134 Zeitgeber, 26, 344

You might also like