You are on page 1of 191

BRAIN BIOCHEMISTRY

And DISORDERS

Author
Dr.Dheaa Shamikh Zageer

Dr.Sundus Fadhil Hantoosh


AUTHOR
Dr.Dheaa Shamikh Zageer
Dr.Sundus Fadhil Hantoosh
ISBN: 978-81-939449-9-8
First Edition: January 2019

This book is sold subject to the condition that it shall not, by way of
trade or otherwise, be lent, resold, hired out, or otherwise circulated
without the publisher's prior written consent in any form of binding or
cover other than that in which it is published and without a similar
condition including this condition being imposed on the subsequent
purchaser and without limiting the rights under copyright reserved
above, no part of this publication may be reproduced, stored in or
introduced into a retrieval system, or transmitted in any form or by any
means (electronic, mechanical, photocopying, recording or
otherwise), without the prior written permission of both the copyright
owner and the above-mentioned publisher of this book.

350

PUBLISHER
MAHI PUBLICATION
Office No.1, Krishnasagar Society, Nr. Shivsagar sharda Mandir Road,
Ahmedabad-380007
mahibookpublication@gmail.com
+(91) 798 422 6340
www.mahipublication.com

Copyright © 2018\ MAHI PUBLICATION


Prelude

This book talks about the biochemistry of this magic organ, the brain. The
book shows briefly the anatomy of human brain and the blood supply and
drainage. The book deals satisfactorily with the saccharides and
derivatives and lipids and derivatives of the brain. It speaks
comprehensively about various diseases results from saccharides and/or
lipids disorders in the brain. This book is for every one interested in the
field of biochemistry of brain and also for every person wants to explore
the mystery of biochemistry of this precious jewel, i.e., the brain, which
God granted us.
CONTENTS
Introduction
Chapter One
Anatomy of Brain
1. Structure of Brain 16
1.1 Cerebrum 18
1.2 Cerebellum 22
1.2.1 Cognition 23
1.3 Brainstem 24
2. Blood Supply and Drainage 25
2.1 Blood Supply 25
2.2 Blood Drainage 25
2.3 The Blood-Brain Barrier 26
Chapter Two
Saccharides and Brain
1. Carbohydrates 27
2. Glucose 27
2.1 Homeostatic Regulation Glucose Metabolism 29
2.2 Glucose Metabolism 32
2.3 Fate of Lactate Generated in Brain 33
2.3.1 The Hypothesis of Lactate Transfer Between Astrocyte and Neuron 35
2.4 Alternative of Glucose as Energy Substrate for The Brain 36
3. Glycogen 37
3.1 Glycogen Metabolism 37
3.1.1 Glycogen Metabolism in Astrocytes 38
4. Glycoproteins 40
4.1 Glycoproteins in Brain 40
4.2 Synthesis of Glycoproteins 41
4.2.1 Carbohydrate Components of Glycoproteins 41
4.2.2 Synthesis of The N-Linked Glycosides 41
4.2.3 Synthesis of The O-Linked Glycosides 42
4.3 Lysosomal Degradation of Glycoproteins 44
Chapter Three
Saccharides Disorders and Brain Diseases
1. Pancreatic β Cells 46
2. Insulin Resistance 47
3. Islet Amyloid Polypeptide 50
4. Diabetes Mellitus 51
4.1 Type 1 Diabetes Mellitus 54
4.2 Type 2 Diabetes Mellitus 54
4.3 Type 3 Diabetes Mellitus 55
4.4 Gestational Diabetes Mellitus 55
5. Hypoglycemia 55
5.1 Hypoglycemia and Brain Energy Metabolism 58
5.2 Hypoglycemia and Cerebral Blood Flow 60
6. Brain Disorders and Diabetes Mellitus 60
6.1 Alzheimer's Disease 60
6.1.1 Amyloid-Beta Protein and Alzheimer's Disease 63
6.1.2 Oxidative Stress and Alzheimer's Disease 64
6.1.3 Advanced Glycation End Products and Alzheimer's Disease 65
6.2 Parkinson's Disease 66
6.3 Autism Spectrum Disorders 70
6.4 Mood and Psychotic Disorders and Type 2 Diabetes Mellitus 71
6.4.1 Defects in Bioenergetics Coupling in Schizophrenia 72
6.5 Effects of Diabetes on Cognitive Function and Brain Structure 73
6.5.1 Cognitive Dysfunction in Type 1 Diabetes Mellitus 73
6.5.2 Cognitive Dysfunction in Type 2 Diabetes Mellitus 74
6.5.3 Impact of Type 1 Diabetes Mellitus on Brain Structure 75
6.5.4 Impact of Type 2 Diabetes Mellitus on Brain Structure 76
7. Hypoglycemia and Brain Function Failure 76
8. Hypometabolism of Glucose in Anorexia Nervosa 78
9. Carbohydrate-Deficient Glycoprotein Syndromes 79
10. Glycoproteinoses 82
10.1 Mannosidosis 83
10.1.1 α-Mannosidosis 83
10.1.2 β-Mannosidosis 85
10.2 Fucosidosis 87
10.3 α-N-Acetylgalactosaminidase Deficiency (Schindler Disease and 89
Kanzaki Diseases)
Chapter Four
Lipids and Human Brain
1. Preview 94
2. Lipid Digestion, Absorption, Metabolism 96
3. Lipid Entrance to Brain 98
3.1 The Membrane Localized Fatty Acid Transporter Proteins 98
4. Fatty Acids 98
4.1 Fatty Acid Synthesis 101
4.2 Oxidation of Fatty Acids 102
4.3 Neuronal Uptake of Fatty Acids 104
4.4 Hypothalamic Fatty Acid Metabolism 105
4.5 Essential Fatty Acids in Brain 107
4.6 Brain Membrane Structure and Function Related to Fatty Acid Profile 108
4.7 Essential Fatty Acids Precursors of Eicosanoids 109
5. Cholesterol 110
5.1 Cholesterol in Brain 112
5.2 Synthesis of Cholesterol in Peripheral Tissues 112
5.3 Reverse Cholesterol Transport 114
5.4 Bile Acids Synthesis 115
5.5 Synthesis of Cholesterol in The Brain 115
6. Triglycerides 118
6.1 Lipogenesis 118
6.2 Lipolysis 120
6.3 Beta-Oxidation The Catabolic Pathway for Triacylglycerol 122
7. Lipoproteins 122
7.1 Chylomicrons 124
7.2 Very-Low-Density Lipoproteins 125
7.3 Intermediate-Density Lipoproteins 125
7.4 Low-Density Lipoproteins 125
7.5 High-Density Lipoproteins 126
7.6 Lipoprotein (a) 127
7.7 Cerebrospinal Fluid Lipoproteins 127
7.8 Apolipoproteins within The Central Nervous System 127
7.8.1 Amyloid Beta 128
8. Hypothalamic Fat Sensing 129
9. Phospholipids 130
9.1 Synthesis of Glycerophospholipids 132
9.2 Phosphatidylserine 133
9.2.1 Synthesis of Phosphatidylserine and Incorporation into Membranes 133
9.2.2 Phosphatidylserine and Neurotransmission 135
9.3 Degradation of Phospholipids 135
10. Glycolipids 136
10.1 Glycosphingolipid Metabolism Pathways 139
10.2 Distribution of Glycolipids in Cell 140
Chapter Five
Lipid Disorders and Brain Diseases
1. Lysosomal Lipid Storage Diseases 143
1.1 Gaucher Disease 144
1.2 Niemann-Pick Disease 148
1.3 Fabry Disease 155
1.4 Krabbe Disease 157
1.5 Farber Disease 159
1.6 Metachromatic Leukodystrophy 163
1.7 Tay-Sachs Disease 164
1.8 Sandhoff Disease 166
2. Hyperlipidemia 167
2.1 Classification of Hyperlipidemia 167
2.2 Ischemic Stroke 169
2.3 Impact of Hyperlipidemia on Cerebral Lipids 169
2.4 Mild Cognitive Impairment in Familial Hypercholesterolemia 170
2.4.1 Lipid Metabolism and Alzheimer's Disease 171
2.4.2 Effect of Amyloid-β on Lipid Metabolism 173
3. Parkinson's Disease 174
4. Prion Diseases 175
5. Smith-Lemli-Opitz Syndrome 178
6. Hereditary Sensory and Autonomic Neuropathy Type 1 182
7. Huntington Disease 183
8. Phosphatidylserine in Deteriorating Brain 186
INTRODUCTION

Introduction
The brain is the most complex part of the human body. This three-pound organ is the seat
of intelligence, interpreter of the senses, initiator of body movement, and controller of
behavior. Lying in its bony shell and washed by protective fluid, the brain is the source of
all the qualities that define our humanity. The brain is the crown jewel of the human body.

The neural developmental processes can be summarized by that the neural tissue,
including the brain and spinal cord, arises from the ectodermal layer, which becomes
identified in the embryonic disk during the second week after fertilization. The
ectodermal tissue soon thickens to form the symmetrical neural plate, which is the
forerunner of the brain and the spinal cord. By day 18 [Carnegie stage (CS) 8], a groove
appears in the midline of the embryonic disk, which becomes deeper and longer due to
the development of the bilateral neural folds along the lateral ends of the neural groove.
By around day 20 [Carnegie stage (CS) 9], the primordia of the three primary brain
vesicles (the forebrain of prosencephalon, the midbrain or mesencephalon and the
rhombencephalon) are discernible as thickenings in the neural plate. The bilateral neural
folds begin to fuse with each other at about 22 days [Carnegie stage (CS) 10]. This fusion
initiates at the level of the future cervical region and extends to both cranial and caudal
directions, to form a neural tube. The mode of fusion of neural folds is not the same in
different species, and at least two initiation sites are recognized in human embryos. The
openings at the rostral and caudal ends of the neural tube, i.e. the anterior and posterior
neuropores, close at 24 days [Carnegie stage (CS) 11] and 28 days [Carnegie stage (CS)
12], respectively, and complete the formation of the neural tube. If the closure of the
neuropores are disturbed, various forms of neural tube defects (NTD) can occur. Neural
tube formation described above, which is accomplished by the fusion of neural folds, is
called primary neurulation. On the other hand, the caudal end of the spinal cord is formed
in a quite different fashion caudally to the posterior neuropore. This caudal part of the
neural tube is induced in the mesenchyme (neural cord) of the tail bud or caudal
eminence, and this developmental event is called secondary neurulation. The three
divisions of the embryonic brain (the prosencephalon, mesencephalon and
rhombencephalon) can be recognized before neural tube closure begins. These three
parts of the brain are called primary brain vesicles. Soon after the closure of the neural
tube, it becomes bent by three flexures: (i) the mesencephalic flexure at the midbrain
level; (ii) the cervical flexure at the junction between the rhombencephalon and the
spinal cord, and (iii) the pontine flexure in the hindbrain. The forebrain soon divides into
an end portion, the telencephalon and the diencephalon from which optic vesicles arise
bilaterally. The hindbrain (rhombencephalon) also divides into a rostral part, the
metencephalon, and a caudal part, the medulla oblongata. The junction between the
midbrain and hindbrain is narrow and is known as the isthmus rhombencephali. By

9
Carnegie stage (CS)15, the future cerebral hemischeres can be recognized in the
trelencephalon. The cerebral hemispheres enlarge rapidly and completely cover the
diencephalon by the end of the embryonic period. During the feral period (nine weeks
and later), the cerebral hemispheres continue to expand, due to the active formation and
differentiation of neurons and glias (described in the next section), and form brain lobes,
sulci and gyri. In addition, the formation of commissural connections, the corpus
callosum in particular, and the development of the cerebellum are important
developmental events that occur during the feral period. Up to 6 weeks of development,
the neuroepithelium from which the brain and spinal cord arise appears as a
homogeneous tissue comprising of neuroepithelial cells. During the following weeks,
some zones become recognizable histologically: (i) the ventricular zone, which is
composed of ventricular cells and dividing neural precursor cells; (ii) the subventricular
zone; (iii) the intermediate zone, through which neurous migrate radially along radial
processes; (iv) the subplate; (v) the cortical plate, where postmitotic neurous condense
and will become layers II-VI of the mature cortex; and (vi) the marginal zone, the
superficial layer important for establishing the laminar organization of the cortex.
Cortical neurous are generated in the ventricular zone, migrate radially and reach their
destination. The radial migration is the mechanism by which developing neurous move
along associated radial glial cells. For proper and effective function of the nervous
system, the axons of fiber tracts need to undergo myelination. Myelination in the central
nervous system (CNS) is undertaken by oligodendrocytes, whereas that in the peripheral
nervous system (PNS) depends on neurilemmal cells derived from the neural crest. The
process is very slow, and the myelin begins to be present in the spinal cord at the end of
the first trimester. In the brain stem, myelination starts at eight weeks of embryonic
development. The vestibulospinal tract becomes myelinated at the end of the second
trimester, and the pyramidal tracts begin later at the end of the third trimester.
Myelination in these tracts completes after birth. The rate of myelin deposition is
greatest during the first two years after birth, and cortical association fibers are the last to
be myelinated. Therefore, the human brain is rather immature at birth with regard to the
degree of myelination.

The brain can be divided into three basic units: the forebrain, the midbrain, and the
hindbrain. The hindbrain includes the upper part of the spinal cord, the brain stem, and a
wrinkled ball of tissue called the cerebellum. The hindbrain controls the body's vital
functions such as respiration and heart rate. The cerebellum coordinates movement and
is involved in learned rote movements. The uppermost part of the brainstem is the
midbrain, which controls some reflex actions and is part of the circuit involved in the
control of eye movements and other voluntary movements. The forebrain is the largest
and most highly developed part of the human brain: it consists primarily of the cerebrum
and the structures hidden beneath it. The cerebrum sits at the topmost part of the brain
and is the source of intellectual activities. The cerebrum is split into two halves
(hemispheres) by a deep fissure. Despite the split, the two cerebral hemispheres
communicate with each other through a thick tract of nerve fibers that lies at the base of
this fissure. Although the two hemispheres seem to be mirror images of each other, they
are different. For instance, the ability to form words seems to lie primarily in the left
hemisphere, while the right hemisphere seems to control many abstract reasoning skills.
For some as-yet-unknown reason, nearly all of the signals from the brain to the body and
10
vice-versa cross over on their way to and from the brain. This means that the right
cerebral hemisphere primarily controls the left side of the body and the left hemisphere
primarily controls the right side.

Each cerebral hemisphere can be divided into sections, or lobes, each of which
specializes in different functions and involve frontal, parietal, occipital, and temporal
lobes. Coating the surface of the cerebrum and the cerebellum is a vital layer of tissue the
thickness of a stack of two or three dimes. It is called the cortex, from the Latin word for
bark. Most of the actual information processing in the brain takes place in the cerebral
cortex. When people talk about "gray matter" in the brain they are talking about this thin
rind. The cortex is gray because nerves in this area lack the insulation that makes most
other parts of the brain appear to be white. The folds in the brain add to its surface area
and therefore increase the amount of gray matter and the quantity of information that can
be processed.

Deep within the brain, hidden from view, lie structures that are the gatekeepers between
the spinal cord and the cerebral hemispheres. These structures not only determine our
emotional state, they also modify our perceptions and responses depending on that state,
and allow us to initiate movements that one make without thinking about them. Like the
lobes in the cerebral hemispheres, the structures described below come in pairs: each is
duplicated in the opposite half of the brain.

The hypothalamus, about the size of a pearl, directs a multitude of important functions. It
wakes one up in the morning, and gets the adrenaline flowing during a test or job
interview. The hypothalamus is also an important emotional center, controlling the
molecules that make one feel exhilarated, angry, or unhappy. Near the hypothalamus lies
the thalamus, a major clearinghouse for information going to and from the spinal cord
and the cerebrum.

An arching tract of nerve cells leads from the hypothalamus and the thalamus to the
hippocampus . This tiny nub acts as a memory indexer—sending memories out to the
appropriate part of the cerebral hemisphere for long-term storage and retrieving them
when necessary. The basal ganglia are clusters of nerve cells surrounding the thalamus.
They are responsible for initiating and integrating movements.

The core component of the nervous system in general, and the brain in particular, is the
neuron or nerve cell, the “brain cells” of popular language. A neuron is an electrically
excitable cell that processes and transmits information by electro-chemical signaling.
Unlike other cells, neurons never divide, and neither do they die off to be replaced by
new ones. By the same token, they usually cannot be replaced after being lost, although
there are a few exceptions.

The average human brain has about 100 billion neurons (or nerve cells) and many more
neuroglia (or glial cells) which serve to support and protect the neurons. Each neuron
may be connected to up to 10,000 other neurons, passing signals to each other via as
many as 1,000 trillion synaptic connections.

11
Neurons consist of three parts. The cell body contains the nucleus, where most of the
molecules that the neuron needs to survive and function are manufactured. Dendrites
extend out from the cell body like the branches of a tree and receive messages from other
nerve cells. Signals then pass from the dendrites through the cell body and may travel
away from the cell body down an axon to another neuron, a muscle cell, or cells in some
other organ. The neuron is usually surrounded by many support cells. Some types of cells
wrap around the axon to form an insulating sheath. This sheath can include a fatty
molecule called myelin, which provides insulation for the axon and helps nerve signals
travel faster and farther. Axons may be very short, such as those that carry signals from
one cell in the cortex to another cell less than a hair's width away. Or axons may be very
long, such as those that carry messages from the brain all the way down the spinal cord.

Information transmission within the brain, such as takes place during the processes of
memory encoding and retrieval, is achieved using a combination of chemicals and
electricity. It is a very complex process involving a variety of interrelated steps.

A typical neuron possesses a soma (the bulbous cell body which contains the cell
nucleus), dendrites (long, feathery filaments attached to the cell body in a complex
branching “dendritic tree”) and a single axon (a special, extra-long, branched cellular
filament, which may be thousands of times the length of the soma).

Every neuron maintains a voltage gradient across its membrane, due to metabolically-
driven differences in ions of sodium, potassium, chloride and calcium within the cell,
each of which has a different charge. If the voltage changes significantly, an
electrochemical pulse called an action potential (or nerve impulse) is generated. This
electrical activity can be measured and displayed as a wave form called brain wave or
brain rhythm.

This pulse travels rapidly along the cell's axon, and is transferred across a specialized
connection known as a synapse to a neighbouring neuron, which receives it through its
feathery dendrites. A synapse is a complex membrane junction or gap (the actual gap,
also known as the synaptic cleft, is of the order of 20 nanometres, or 20 millionths of a
millimetre) used to transmit signals between cells, and this transfer is therefore known as
a synaptic connection. Although axon-dendrite synaptic connections are the norm, other
variations (e.g. dendrite-dendrite, axon-axon, dendrite-axon) are also possible. A typical
neuron fires 5 - 50 times every second.

Each individual neuron can form thousands of links with other neurons in this way,
giving a typical brain well over 100 trillion synapses (up to 1,000 trillion, by some
estimates). Functionally related neurons connect to each other to form neural networks
(also known as neural nets or assemblies). The connections between neurons are not
static, though, they change over time. The more signals sent between two neurons, the
stronger the connection grows (technically, the amplitude of the post-synaptic neuron's
response increases), and so, with each new experience and each remembered event or
fact, the brain slightly re-wires its physical structure.

The interactions of neurons is not merely electrical, though, but electro-chemical. Each

12
axon terminal contains thousands of membrane-bound sacs called vesicles, which in
turn contain thousands of neurotransmitter molecules each. Neurotransmitters are
chemical messengers which relay, amplify and modulate signals between neurons and
other cells. The two most common neurotransmitters in the brain are the amino acids
glutamate and GABA; other important neurotransmitters include acetylcholine,
dopamine, adrenaline, histamine, serotonin and melatonin.

When stimulated by an electrical pulse, neurotransmitters of various types are released,


and they cross the cell membrane into the synaptic gap between neurons. These
chemicals then bind to chemical receptors in the dendrites of the receiving (post-
synaptic) neuron. In the process, they cause changes in the permeability of the cell
membrane to specific ions, opening up special gates or channels which let in a flood of
charged particles (ions of calcium, sodium, potassium and chloride). This affects the
potential charge of the receiving neuron, which then starts up a new electrical signal in
the receiving neuron. The whole process takes less than one five-hundredth of a second.
In this way, a message within the brain is converted, as it moves from one neuron to
another, from an electrical signal to a chemical signal and back again, in an ongoing
chain of events which is the basis of all brain activity.

The electro-chemical signal released by a particular neurotransmitter may be such as to


encourage to the receiving cell to also fire, or to inhibit or prevent it from firing. Different
neurotransmitters tend to act as excitatory (e.g. acetylcholine, glutamate, aspartate,
noradrenaline, histamine) or inhibitory (e.g. GABA, glycine, seratonin), while some
(e.g. dopamine) may be either. Subtle variations in the mechanisms of
neurotransmission allow the brain to respond to the various demands made on it,
including the encoding, consolidation, storage and retrieval of memories.

As has been mentioned, in addition to neurons, the brain contains about an equal mass of
glial cells (neuroglia or simply glia), the most common types being oligodendrocytes,
astrocytes and microglia. Because they are so much smaller than neurons, there are up to
10 times as many in number, and different areas of the brain have higher or lower
concentrations of glia. It used to be thought that the role of glial cells was limited to the
physical support, nutrition and repair of the neurons of the central nervous system.
However, more recent research suggests that glia, particularly astrocytes, actually
perform a much more active role in brain communication and neuroplasticity, although
the extent and mechanics of this role is still uncertain, and a substantial amount of
contemporary brain research is now focused on glials cells.

Saccharides are ubiquitous and form the intricate coat that surrounds the cells of every
organism. Glycoconjugates are found on the surface of all cells, cell membranes, and in
the extracellular space of tissues, and serve as recognition markers (identification tags)
and cell receptors crucial for transmitting biochemical signals into and between cells.
Thus, the body uses glycoconjugates to facilitate, guide, and allow the cell-to-cell
communication that is essential for normal physiological function. There is emerging
evidence that saccharides are important for the growth and development of the brain. A
further supporting role of saccharides is in relationship to myelin, an essential
component of the nervous system. Myelin serves as an insulator for neuronal

13
membranes and is needed for rapid and efficient nerve impulse conduction. Several
studies in animals have investigated the composition of myelin and have shown that
glycoconjugates, specifically glycolipids, and in particular those associated with
galactose (galactolipids), are essential for proper myelin formation. Furthermore,
galactolipids have been found to play an essential role in the interactions between
myelinating glial cells and neuronal axons important for the propagation of an electrical
impulse. In addition to being present in myelin, saccharides are also concentrated in
neuronal membranes and the synaptic junctions. Recent research has demonstrated that
saccharides are incorporated into the cell membrane of hippocampal neurons and in
isolated dendrites of rats. A recent review suggested that saccharides play an important
role in synapse formation and synaptic plasticity. In addition, there is some evidence to
suggest that saccharides are important in neurotransmitter and receptor actions.
Research suggests that saccharides can increase neurotransmitter release, thus
increasing neurotransmitter availability at post-synaptic receptors. Abnormalities in the
availability and the metabolism of saccharides in the brain have been related to poor
neurological functioning. In addition, a recent study has shown that poor metabolism of
saccharides in the brain may be related to poor mental functioning in individuals with
schizophrenia, Down's syndrome, and dementia, as well as among older adults. The
results suggest that compromised metabolism of saccharides may disturb neuronal
function and, in conjunction with neuron degeneration, may be related to brain disorder.

The lipids can be classified as total cholesterol (TC) and its derivatives such as;
triglycerides (TAG), low density lipoprotein (LDL), high density lipoprotein (HDL) and
very low density lipoprotein (VLDL) cholesterol. Lipids are insoluble in water but are
soluble in alcohol and other solvents. Hence, they are transported around the body as
lipoproteins. Lipids originate from two sources: endogenous lipids, synthesized in the
liver, and exogenous lipids, which are ingested and absorbed in the intestine.

In brain, lipids account for 20–25% of the dry weight and an enormous variety of
molecular species exists. Cholesterol is a major constituent of the human brain, and the
brain is the most cholesterol-rich organ. Numerous lipoprotein receptors and
apolipoproteins are expressed in the brain. Cholesterol is tightly regulated between the
major brain cells and is essential for normal brain development.

The different glycerophospholipid species, together with cholesterol and sphingolipids,


are the major components of cell membranes and determine the structure and fluidity of
somatodendritic and axonal membranes of neurons, and also of those of the glia and
organelles. Brain lipids are also involved in important metabolic pathways, usually
related to the energetic homeostasis of cells. Whereas the role of brain lipids in
intracellular signaling processes has been known for decades (e.g., phosphoinositides),
their implication in neurotransmission is acquiring relevance due to their participation in
cell proliferation, growth, and neuroprotection. Some lipid molecules mediate their
action by binding to specific receptors. The enormous abundance of lipid molecules in
the central nervous system (CNS) suggests that their role is not limited to be structural
and energetic components of cells. Over the last decades, some lipids in the central
nervous system (CNS) have been identified as intracellular signalers, while others are
known to act as neuromodulators of neurotransmission through binding to specific
14
receptors. Neurotransmitters of lipidic nature, currently known as neurolipids, are
synthesized during the metabolism of phospholipid precursors present in cell
membranes. Most of the glycerophospholipids are involved in cellular processes in the
central nervous system (CNS), including regulation of cell growth, signal transduction,
cell adhesion and neuronal plasticity.

Defects in cholesterol metabolism lead to structural and functional central nervous


system (CNS) diseases such as Smith-Lemli-Opitz syndrome, Niemann-Pick type C
disease, and Alzheimer's disease. These diseases affect different metabolic pathways
(cholesterol biosynthesis, lipid transport and lipoprotein assembly, apolipoproteins,
lipoprotein receptors, and signaling molecules). Defects in lipids metabolism lead to
structural and functional central nervous system diseases. Lipid storage diseases, or the
lipidoses, are a group of inherited metabolic disorders in which harmful amounts of fatty
materials (lipids) accumulate in various cells and tissues in the body. People with these
disorders either do not produce enough of one of the enzymes needed to break down
(metabolize) lipids or they produce enzymes that do not work properly. Over time, this
excessive storage of fats can cause permanent cellular and tissue damage, particularly in
the brain.

15
CHAPTER ONE
ANATOMY OF BRAIN

1.Structure of Brain
The central nervous system (CNS) is composed of the brain and spinal cord. The
peripheral nervous system (PNS) is composed of spinal nerves that branch from the
spinal cord and cranial nerves that branch from the brain. The peripheral nervous system
(PNS) includes the autonomic nervous system, which controls vital functions such as
breathing, digestion, heart rate, and secretion of hormones. The big hole in the middle of
the skull (foramen magnum) is where the spinal cord exits. The brain communicates with
the body through the spinal cord and twelve pairs of cranial nerves. Protected within the
skull, the brain is composed of the cerebrum, cerebellum, and brainstem. The brain
receives information through five senses: sight, smell, touch, taste, and hearing, often
many at one time. The right and left hemispheres of the brain are joined by a bundle of
fibers called the corpus callosum that delivers messages from one side to the other. Each
hemisphere controls the opposite side of the body. The left hemisphere controls speech,
comprehension, arithmetic, and writing. The right hemisphere controls creativity, spatial
ability, artistic, and musical skills. The left hemisphere is dominant in hand use and
language in about 92% of people.

The brain and spinal cord are covered and protected by three layers of tissue called
meninges. The meninges are three connective tissue membranes that lie just external to
the brain. The function of these layers are to;

1- Cover and protect the brain.


2- Protect blood vessels and enclose venous sinuses.
3- Contain cerebrospinal fluid .
4- Form partitions within the skull.

The major meninges:


1- Dura mater: external leathery tissue layer; protects brain, encloses venous sinuses,
and form partitions within the skull.
2- Arachnoid mater: middle tissue layer forming loose brain covering, houses
cerebrospinal fluid.
3- Pia mater: innermost delicate tissue layer adhered tightly to brain; contains many
blood vessels.

16
Figure (1): Meninges of brain (www.google.com)

There are two special dural folds, the falx and the tentorium. The falx separates the right
and left hemispheres of the brain and the tentorium separates the cerebrum from the
cerebellum. The space between the dura and arachnoid membranes is called the subdural
space. The space between the arachnoid and pia is called the subarachnoid space. It is
here where the cerebrospinal fluid bathes and cushions the brain.

The brain is connected by commissural, association, and projection tracts. The main
commissural tracts (interconnecting both hemispheres) are: corpus callosum, and
anterior and posterior commissures. The major association tracts (interconnecting
different regions of the same hemisphere) are: superior longitudinal, middle
longitudinal, inferior longitudinal, superior occipito-frontal, inferior occipito-frontal,
and uncinate fascicles. The main projection tracts (connecting the cortex with
subcortical structures) contain: cortico-spinal, cortico-thalamic (including optic
radiation ), cortico-bulbar, and cortico-pontine tracts as well as auditory radiation.

The brain is made up of two types of cells: nerve cells (neurons) and glia cells. Nerve
cells consist of cell body, dendrites, and an axon. Neurons transmit their energy, or talk,
to each other across a tiny gap called a synapse. The dendrites pick up messages from
other nerve cells. These messages are passed to the cell body, which determines if the
message should be passed along. Important messages are passed to the end of the axon
where sacs containing neurotransmitters open into the synapse. The neurotransmitter
molecules cross the synapse and fit into special receptors on the receiving nerve cell,
which stimulate that cell to pass the message. Glia cells are brain cells that provide
neurons with nourishment, protection, and structural support. Glia cells include:
17
1- Astroglia or astrocytes: transport nutrients to neurons, hold neurons in place, digest
parts of dead neurons, and regulate the blood brain barrier.
2- Oligodendroglia cells: provide insulation (myelin) to neurons.
3- Ependymal cells line the ventricles and secrete cerebrospinal fluid (CSF).
4- Microglia: digest dead neurons and pathogens.

1.1 Cerebrum
Cerebrum is divided into two hemispheres. The cerebrum is the largest region of the
human brain-the two hemispheres together account for about 85% of total brain mass.
The cerebrum forms the superior part of the brain, covering and obscuring the
diencephalon and brainstem. The surface of the cerebrum has a folded appearance called
the cortex. The cortex contains about 70% of the one hundred billion nerve cells. The
nerve cell bodies color the cortex grey-brown giving it its name gray matter. The gray
matter contains in addition to nerve cell bodies, the unmyelinated axons arranged in six
discrete layers. Beneath the cortex are long connecting fibers between neurons, the
myelinated axons, which make up the white matter. The folding of the cortex increases
the brain's surface area allowing more neurons to fit inside the skull and enabling higher
functions. Elevated ridges of cortex tissue are called gyri (singular: gyrus) separated by
shallow grooves called sulci (singular sulcus) mark nearly the entire surface of the
cerebral hemispheres. Deeper grooves, called fissures, separates large regions of the
brain. The cerebral hemispheres have distinct fissures which divide the brain into lobes.
Corpus callosum is the main bridge of white fibers that connect the two hemispheres of
the cerebrum.

The cerebrum comprises:

1- Left and right cerebral hemispheres.


2- Interbrain between the cerebrum and the brainstem termed diencephalon.
3- Deep gray nuclei.

The cerebral hemispheres are the largest compartment of the brain and are parcellated
into five lobes and each lobe is divided into areas that serve specific functions. These
lobes are:

1- Frontal lobe: region of cerebrum located under the frontal bone, contains the
primary motor cortex (precentral gyrus) and is involved in complex learning.
2- Temporal lobe: region of the cerebrum located under temporal bone; processes
information associated with hearing and equilibrium.
3- Parietal lobe: region of the cerebrum located under parietal bone, contains the
primary sensory cortex (postcentral gyrus) and is involved in language acquisition.
4- Occipital lobe: region of the cerebrum located under occipital bone; processes
visual information and is related to our understanding of the written word.
5- Limbic lobe: included in the limbic system are the cingulate gyri, hypothalamus,
amygdala, and hippocampus.

18
Figure (2): Cerebrum (www.google.com)

Figure (3): Limbic system (www.google.com)

Fornix is a bridge of white matter inferior to the corpus callosum; links regions of the
limbic system (emotional brain) together.

The insula is a region of the cerebrum deep within the lateral sulcus which is a deep
groove that separates the frontal and parietal lobes from the temporal lobe of the
cerebrum. The insula is sometimes classified as the central or insular lobe. The insula
processes information associated with hearing and equilibrium. The central sulcus
separates the frontal lobe anterior from the parietal lobe posterior. The Sylvian (lateral)
fissure demarcates the temporal lobe below from the frontal and parietal lobes above.
The posterior-occipital fissure separates the parietal lobe anterior from the occipital lobe
posterior. The cingulate sulcus separates the frontal lobe above from the limbic lobe
below.

The cortex has three surfaces: Lateral, medial, and inferior (also called basal or ventral).
Moreover, the transitional areas form the frontal, temporal, and occipital poles. These
surfaces are:
19
1- Lateral surface: four lobes are present on the lateral surface, which include: frontal,
temporal, parietal, and occipital.
2- Medial surface is where the frontal, parietal, occipital, and limbic lobes are present.
3- Inferior surface includes the frontal, temporal, and occipital lobes.

Diencephalon is surrounded by the cerebral hemispheres, the diencephalon forms the


central core of the brain. Consisting of largely of three paired structures, the thalamus,
hypothalamus, and epithalamus, the diencephalon plays a vital role in integrating
conscious and unconscious sensory information and motor commands.

Major anatomical regions of the diencephalon are:

1- Thalamus: composes 80% of diencephalon; major relay point and processing


center for all sensory impulses (excluding olfaction).
2- Intermediate mass: a flattened gray band of tissue connecting the two halves of the
thalamus.
3- Hypothalamus: region inferior to thalamus; main regulatory center involved in
visceral control of the body and maintenance of overall homeostasis.
4- Infundibulum: connects pituitary gland to hypothalamus.

Figure (4):Diencephalon (www.google.com)

The main deep gray nuclei are:


a- Basal ganglia (nuclei): a group of sub-cortical gray matter lie deep within the
cerebral white matter and is a third basic region of the cerebrum. Basal ganglia
comprise:
1- Caudate nucleus: basal nucleus initiates voluntary movements and coordinates
slow skeletal muscle contractions (e.g. posture and balance).
2- Putamen: basal nucleus initiates voluntary movements and coordinates slow
20
skeletal muscle contractions (e.g. posture and balance).
3- Globus pallidus: basal nucleus initiates voluntary movements and coordinates slow
skeletal muscle contractions (e.g. posture and balance).

The lentiform nucleus lies beneath the insula. It is said to resemble a lens. The outer
portion of the lentiform nucleus, immediately beneath the insula, is the putamen. The
inner part is the globus pallidus. The lentiform nucleus with the caudate nucleus forms
the striatum.

b- Thalamus: the largest subdivision of the diencephalon that consists chiefly of an


ovoid mass of nuclei in each lateral wall of the third ventricle and serves chiefly to
relay impulses and especially sensory impulses to and from the cerebral cortex.
c- Hippocampus: a curved elongated ridge that extends over the floor of the
descending horn of each lateral ventricle of the brain, that consists of gray matter
covered on the ventricular surface with white matter, and that is involved in
forming, storing, and processing memory

Amygdala (amygdaloid body): it is an almond-shaped section of nervous tissue located


in the temporal (side) lobe of the brain. There are two amygdalae per person normally,
with one amygdala on each side of the brain. They are thought to be a part of the limbic
system within the brain, which is responsible for emotions, survival instincts, and
memory. The amygdala is responsible for the perception of emotions such as anger, fear,
and sadness, as well as the controlling of aggression.

Figure (5): Basal nuclei (www.google.com)

21
Three types of white matter connections are distinguished in the cerebral hemispheres:
1- Commissural tracts: connect corresponding cortical areas in the two hemispheres.
They cross from one cerebral hemisphere to the other through bridges called
commissures. The great majority of commissural tracts pass through the corpus
callosum. A few tracts pass through the much smaller anterior and posterior
commissures. Commissural tracts enable the left and right sides of the cerebrum to
communicate with each other.
2- Associated tracts: The tracts that connect cortical areas within the same
hemisphere. Long association fibers connect different lobes of a hemisphere to each
other whereas short association fibers connect different gyri within a single lobe.
Among their roles, association tracts link perceptual and memory centers of the
brain. The cingulum is a major association tract. The cingulum forms the white
matter core of the cingulate gyrus and links from this to the entorhinal cortex.
3- Projection tracts: connect the cerebral cortex with the corpus striatum,
diencephalon, brainstem and the spinal cord. The corticospinal tract for example,
carries motor signals from the cerebrum to the spinal cord. Other projection tracts
carry signals upward to the cerebral cortex. Superior to the brainstem, such tracts
form a broad, dense sheet called the internal capsule between the thalamus and
basal nuclei, then radiate in a diverging, fanlike array to specific areas of the cortex.

Ventricles situated within the brain are central hollow civilities. These ventricles are
continuous with one another and with the central canal of the spinal cord. The hollow
ventricular chambers are filled with cerebrospinal fluid (CSF), a fluid that forms a liquid
cushion for the brain. The cerebrospinal fluid (CSF) helps nourish the brain and there is
some evidence that hormones circulate in the brain via this pathway. Cerebrospinal fluid
(CSF) is secreted in the choroid plexus (a network of vessels) and circulates from the
lateral ventricles through the paired interventricular foramen (foramen of Monro) to the
third ventricle, and then via the aqueduct of Sylvius to the fourth ventricle.

The left and right lateral ventricles are the largest and each contains:
1- Body (or centralportion)
2- Atrium (or trigon)
3- Horns which comprise:
- Frontal (anterior)
- Occipital (posterior)
- Temporal (inferior)

1.1Cerebellum
The cerebellum is approximately one-tenth of the cerebrum in size and weight and is
situated in the posterior cranial fossa. The cerebellum contains almost 80% of the total
brain neurons. Its circuitry is classically viewed to be involved in motor control and
motor learning. The cerebellum does not contribute to movement initiation, and thus its
damage is not associated with paralysis. However, coordinated, precise, and smooth
execution of voluntary movements and their adaptive modification rely on an intact
cerebellum. The cerebellum provides the precise timing and appropriate patterns of
skeletal muscle contraction for smooth, coordinated movements and agility needing for
our daily lives (e.g., driving). Cerebellar activity occurs subconsciously. The cerebellum

22
is located dorsal to the pons and medulla of brainstem and protrudes under the occipital
lobes of the cerebral hemispheres, from which it is separated by the transverse fissure.
Additionally, The cerebellum is located posterior to the fourth ventricle and is rostrally
separated from the cerebrum by an extension of the dura matter called tentorium
cerebelli. It consists of two lateral hemispheres and a narrow midline zone (i.e. the
cerebellar vermis), and its surface has many parallel thin transverse fold called folia. The
cerebellum consists of an outer layer of highly convoluted gray matter (cerebellar
cortex) surrounding a highly branched body of white matter known as the arbor vitae,
which in turn surrounds the three pairs of deep cerebellar nuclei embedded in the central
cerebellar white matter (corpus medullare). From medial to lateral, the deep nuclei are
the fastigial, interposed (consisting of globose and emboliform nuclei), and dentate
nuclei. Anatomically, the cerebellum is divided into three lobes by two transverse
fissures. The cerebellum is further subdivided into ten transverse lobules marked by
Roman numerals (lobules I-X). The cerebellum is attached to the brainstem through its
three pairs of peduncles (inferior, middle, and superior). All efferents and afferents of
cerebellum pass through these peduncles to reach their targets. Aside from the
flocculonodular lobe (archicerebellum/vestibulocerebellum, a small and evolutionarily
primitive part of the cerebellum), the medial (vermis) and intermediate (paravermis)
lobes are most commonly called the spinocerebellum, with the lateral zone
(hemispheres) regarded as the cerebrocerebellum.

There is a role for cerebellum in higher cognitive functions. Cerebellar activation during
cognitive tasks involving working memory, language, time perception, executive
functioning, and emotional processing.

Figure (6): Cerebellum (www.google.com)

1.2.1 Cognition
Cognition is the activity of knowledge: the acquisition, organization, and use of
knowledge. Cognition indeed refers to the mental process by which external or internal
input is transformed, reduced, elaborated, stored, recovered, and used. As such, it
involves a variety of functions such as perception, attention, retention, recall, decision
making, reasoning and computation, imaging, planning and executing actions, the
23
formation of knowledge, memory and working memory, judgment and evaluation,
problem solving, comprehension and production of language. Cognitive processes use
existing knowledge and generates new knowledge.

1.3 Brainstem
The brainstem connects the cerebrum and cerebellum to the spinal cord. The brainstem
participates in a wide range of functions including control of movement, modulation of
pain, autonomic reflexes, arousal, and consciousness. The brainstem performs many
automatic functions such as breathing, heart rate, body temperature, wake and sleep
cycles, digestion, sneezing, coughing, vomiting, and swallowing. The brainstem is
subdivided into:

1- Midbrain: region of brainstem between the diencephalon and pons; contains


multiple fiber tracts running between higher and lower neural centers.
- Superior colliculus: part of midbrain (corpora quadrigemina); contains nerve reflex
centers involved in coordinate eye movements, focusing, and papillary responses.
- Inferior colliculus: part of the midbrain (corpora quadrigemina); contains nerve
reflex centers involved in auditory reflexes.
2- Pons: region of brainstem between the midbrain and medulla oblongata; serves as
the bridge (connection) between the two regions, and the cerebellum.
3- Medulla oblongata: the most inferior portion of the brainstem; contains the cardiac,
vasomotor, and respiratory centers.

Ten of the twelve cranial nerves originate in the brainstem. The nuclei for cranial nerves
III, IV, and part of V (sensory) are found in the midbrain. Other key midbrain structures
include the substantia nigra and ventral tegmental area ( both dopamine-containing), the
pedunculopontine nucleus (acetyl choline-containing) and the first of the raphe nuclei
(serotonin-containing). Caudal to the midbrain is the pons. The nuclei for cranial nerves
VI, VII, VIII, and part of V (motor) are found in the pons. Caudal to the pons is the
medulla, containing the nuclei for cranial nerves IX, X, XI, XII, and a portion of V.

Figure (7): Brainstem (www.google.com)


24
2.Blood Supply and Drainage

2.1Blood Supply
The internal carotid arteries supply oxygenated blood to the front of the brain and the
vertebral arteries supply blood to the back of the brain. These two circulations join
together in the circle of Willis, a ring of connected arteries that lies in the interpeduncular
cistern between the midbrain and pons.

The internal carotid arteries are branches of the common carotid arteries. They enter the
cranium through the carotid canal, travel through the cavernous sinus and enter the
subarachnoid space. They then enter the circle of Willis, with two branches, the anterior
cerebral arteries emerging. These branches travel forward and then upward along the
longitudinal fissure, and supply the front and midline parts of the brain. One or more
small anterior communicating arteries join the two anterior cerebral arteries shortly after
they emerge as branches. The internal carotid arteries continue forward as the middle
cerebral arteries. They travel sideways along the sphenoid bone of the eye socket, then
upwards through the insula cortex, where final branches arise. The middle cerebral
arteries send branches along their length.

The vertebral arteries emerge as branches of the left and right subclavian arteries. They
travel upward through transverse foramina– spaces in the cervical vertebrae and then
emerge as two vessels, one on the left and one on the right of the medulla. They give off
one of the three cerebellar branches. The vertebral arteries join in front of the middle part
of the medulla to form the larger basilar artery, which sends multiple branches to supply
the medulla and pons, and the two other anterior and superior cerebellar branches.
Finally, the basilar artery divides into two posterior cerebral arteries. These travel
outwards, around the superior cerebellar peduncles, and along the top of the cerebellar
tentorium, where it sends branches to supply the temporal and occipital lobes. Each
posterior cerebral artery sends a small posterior communicating artery to join with the
internal carotid arteries.

2.2Blood Drainage
Cerebral veins drain deoxygenated blood from the brain. The brain has two main
networks of veins: an exterior or superficial network, on the surface of the cerebrum that
has three branches, and an interior network. These two networks communicate via
anastomosing (joining) veins. The veins of the brain drain into larger cavities the dural
venous sinuses usually situated between the dura mater and the covering of the skull.
Blood from the cerebellum and midbrain drains into the great cerebral vein. Blood from
the medulla and pons of the brainstem have a variable pattern of drainage, either into the
spinal veins or into adjacent cerebral veins.

The blood in the deep part of the brain drains, through a venous plexus into the cavernous
sinus at the front, and the superior and inferior petrosal sinuses at the sides, and the
inferior sagittal sinus at the back. Blood drains from the outer brain into the large
superior sagittal sinus, which rests in the midline on top of the brain. Blood from here
joins with blood from the straight sinus at the confluence of sinuses.

25
Blood from here drains into the left and right transverse sinuses. These then drain into the
sigmoid sinuses, which receive blood from the cavernous sinus and superior and inferior
petrosal sinuses. The sigmoid drains into the large internal jugular veins.

2.3 The Blood–Brain Barrier


The blood-brain barrier (BBB), a unique component of the central nervous system, is
comprised of tight junction containing endothelium, astrocyte foot processes lining the
microvasculature, and pericytes wrapping capillary endothelial cells. The structure is
responsible for the brain being the most tightly regulated organ regarding what
substances are able to enter and exit, requiring specific transporters and/or active
processes to shuttle water soluble molecules and compounds into and out of the
parenchyma.

The barrier is less permeable to larger molecules, but is still permeable to water, carbon
dioxide, oxygen, and most fat-soluble substances (including anaesthetics and alcohol).
The blood–brain barrier is not present in areas of the brain that may need to respond to
changes in body fluids, such as the pineal gland, area postrema, and some areas of the
hypothalamus. There is a similar blood–cerebrospinal fluid barrier, which serves the
same purpose as the blood–brain barrier, but facilitates the transport of different
substances into the brain due to the distinct structural characteristics between the two
barrier systems.

For further readings:


1. Arvanitakis, Z. (2018). Review: relationship of type2 diabetes to human brain
pathology. Neuropathology and Applied Neurobiology, 44:347-362.
2. Braudimonte, M.; Bruno, N.; Collina, S. (2006). Psychological concepts: an
international historical perspective. Editors: Pawlik, P.; d'Ydewalle, G. Psychology
Press, UK.
3. Gray, H. (1918). Anatomy the human body. Twentieth edition . LEA & FEBIGER.
Philadelphia & New York.
4. Hines, T. (2018). Anatomy of the brain. Mayfield Clinic.com.
5. Hurley, R.; Flashman, L.; Chow, T.; Taber, K. (2010). Window to the brain the
brainstem: anatomy, assessment, and clinical syndromes. J Neuropsychiatry Clin
Neurosci, 22(1).
6. Marieb, E.; Hoehn, K. (1994). Human anatomy and physiology. 9th ed. Pearson,
ISBN 13:9780321743268.
7. Nowinski, W.; Miller, K. (2011). Biomechanics of the brain, biological and medical
physics, biomedical engineering. DOI 10.1007/978-1-4419-9997-9-2. Singapore.
8. Roostaei, T.; Nazeri, A.; Sahraian, M.; Minagar, A. (2014). The human cerebellum
a review of physiologic neuroanatomy. Neuro Clin, 32:859-869.
9. Saladin, Kenneth (2012). Anatomy & Physiology: The Unity of Form and
Function. New York: McGraw Hill. p. 531. ISBN 978-0-07-337825-1.
10. Standring, Susan (2005). Gray's Anatomy: The Anatomical Basis of Clinical
Practice (39th ed.). Churchill Livingstone. p. 411. ISBN 9780443071683.

26
CHAPTER TWO
SACCHARIDES AND BRAIN

1.Carbohydrates
Carbohydrates are the most abundant biomolecules belonging to class of organic
compounds found in living organisms on earth. Carbohydrates are linked with amino
acid polymers (proteins) forming glycoproteins and with lipids as glycolipids.

Carbohydrates are polyhydroxylated aldehydes or ketones and their derivatives. Simple


carbohydrates or the entire carbohydrate family may also be called saccharides. In
general carbohydrates have the empirical formula (CH2O)n.

Monosaccharides the simplest and smallest unit of the carbohydrates is the


monosaccharide from which disaccharides, oligosaccharides, and polysaccharides are
constructed. Monosaccharides are either aldehydes or ketones with one or more
hydroxyl groups. One of the carbon atoms is double-bonded to an oxygen atom to form a
carbonyl group; each of the other carbon atoms has a hydroxyl group. If the carbonyl
group is at an end of the carbon chain, the monosaccharide is an aldehyde and is called an
aldose; if the carbonyl group is at any other position, the monosaccharide is a ketone and
is called a ketose. The carbon atoms, to which hydroxyl groups are attached, are often
chiral centers, and stereoisomerism is common among monosaccharides. The symbols
D and L designate the absolute configuration of the assymetric carbon farthest from the
aldehyde or ketone group. The cyclic structure of monosaccharides is represented by
Haworth projection, where the α-anomer has the OH- of the anomeric carbon under the
ring structure, and the β-isomer has the OH- of the anomeric carbon on top of the ring
structure.

A disaccharide consists of two monosaccharides joined by O-glycosidic bond.

An oligosaccharide is a saccharide polymer containing a small number (typically three


to ten) of component sugars.

Polysaccharides are relatively complex carbohydrates. They are polymers made up of


many monosaccharides joined together by glycosidic bonds. They are very large and
often branched macromolecules. When all the monosaccharides in a polysaccharide are
of the same type, the polysaccharide is called a homopolysaccharide and when more than
one type of monosaccharide is present, they are called heteropolysacchraides.

2.Glucose
Systemic name for glucose is D-glucose and the molecular formula is C6H12O6. The
living cells use it as a source of energy and metabolic intermediate. D-glucose is often
referred to as dextrose monohydrate or simply dextrose. The mirror image of the
molecule, L-glucose, cannot be metabolized by cells in the biochemical process known
as glycolysis.
27
The mammalian brain depends on glucose as its main source of energy. In the adult brain,
neurons have the highest energy demand, requiring continuous delivery of glucose from
blood. In humans, the brain accounts for about 2% of the body weight, but it consumes
about 20% of glucose-derived energy making it the main consumer of glucose (about
5.6mg glucose per 100g human brain tissue per minute. Glucose metabolism provides
the fuel for physiological brain function through the generation of adenosine
triphosphate (ATP), the foundation for neuronal and non-neuronal cellular maintenance,
as well as the generation of neurotransmitters. Therefore, tight regulation of glucose
metabolism is critical for brain physiology and disturbed glucose metabolism in the
brain underlies several diseases affecting both the brain itself as well as the entire
organism.

Brain function is optimal between blood glucose levels of 80-270mg/100ml. The


maintenance of blood glucose within the normal range (fasting blood glucose level
between 70-100mg/100ml; 2 hours postprandial level between 70-145mg/100ml)
involves the hormones insulin, glucagon, and cortisol. It can be seen that rather wide
ranges of blood glucose levels are associated with normal functioning of the brain.

Specialized centers in the brain, including proopiomelanocortin (POMC) and agouti-


related peptide (AgRP) neurons in the hypothalamus, sense central and peripheral
glucose levels and regulate glucose metabolism through the vagal nerve as well as
neuroendocrine signals. Glucose supply to the brain is regulated by neurovascular
coupling and may be modulated by metabolism-dependent and –independent
mechanisms. Glucose enters the brain from the blood by crossing the blood brain barrier
(BBB) through glucose transporter1 (CLUT1). Glucose and other metabolites (e.g.
lactate, Lac) are rapidly distributed through a highly coupled metabolic network of brain
cells. Glucose provides energy for neurotransmission. The largest proportion of energy
in the brain is consumed for neuronal computation and information processing, e.g. the
generation of action potentials and postsynaptic potentials generated after synaptic
events, and the maintenance of ion gradients and neuronal resting potential.
Additionally, glucose metabolism provides the energy and precursors for the
biosynthesis of neurotransmitters.

Figure (8): Chemical structure of glucose molecule (www.google.com)

28
2.1Homeostatic Regulation Glucose Metabolism
For a continuous and constant supply of glucose to the brain to occur, a coordinated
function of several organs such as the liver, white and brown adipose tissues, muscles,
and the brain itself is essentially required. Function of all these organs is again regulated
by glucose itself, which behaves similar to a regulatory signal that controls endocrine
cells secretion and activation of neurons in the peripheral and central nervous systems.
Glucose sensing system in the central nervous system controls not only the feeding
behavior, but also glucose and energy homeostasis.

The brain, particularly the hypothalamus, has a key role in the homeostatic regulation of
energy and glucose metabolism. The brain modulates various aspects of metabolism,
such as food intake, energy expenditure, insulin secretion, hepatic glucose production,
and glucose/fatty acid metabolism in adipose tissue and skeletal muscle. Highly
coordinated interactions between the brain and peripheral metabolic organs are critical
for the maintenance of energy and glucose homeostasis. In normal individuals, food
intake and energy expenditure are tightly regulated by homeostatic mechanisms to
maintain energy balance. The brain, particularly the hypothalamus is primarily
responsible for the regulation of energy homeostasis. The brain monitors changes in the
body energy state by sensing alterations in the plasma levels of key metabolic hormones
and nutrients. Specialized neuronal networks in the brain coordinate adaptive changes in
food intake and energy expenditure in response to altered metabolic conditions.

A specialized neuronal population in the brain senses hormones (insulin and leptin) and
nutrients (glucose and fatty acid) to regulate glucose homeostasis. The major sites of
convergence of these metabolic signals are the hypothalamus and brain stem.

Brain regions related to the control of glucose metabolism contain neurons whose
excitability changes with alterations in glucose concentrations in the extracellular fluid..
These glucose sensing neurons are found in the hypothalamic nuclei and brainstem,
which are also important areas in the control of energy balance. Glucose sensing neurons
are subgrouped into two types, glucose-excited neurons are excited when extracellular
glucose levels increase. In contrast, glucose-inhibited neurons are activated by a fall in
extracellular glucose concentration. Glucose-excited neurons are mostly located in the
ventromedial hypothalamus (VMH), the hypothalamic arcuate nucleus (ARC) and the
paraventricular nucleus (PVN), whereas glucose-inhibited neurons are distributed in the
lateral hypothalamus (LH),hypothalamic arcuate nucleus (ARC), and paraventricular
nucleus (PVN). Both types of neurons are also located in the dorsal vagal complex in the
brainstem, which encompasses the nucleus of the solitary tract (NTS), area postrema and
the dorsal motor nucleus of the vagus.

The brain has been recognized to be a site of insulin action with regard to glucose
homeostasis. Insulin acts on the brain to modulate hepatic glucose metabolism. A role is
for hypothalamic insulin actions in controlling glucose metabolism in peripheral organs.
Central insulin actions are mediated via neuronal KATP channel-vagus nerve-hepatic
interleukin-6/ signal transducer and activator of transcription 3 (IL-6/STAT3) signaling.

The function of the glucose-sensing neurons in the brain is basically to regulate and

29
maintain glucose homeostasis. Glucose-sensing neurons or the glucose-excited neurons
are possible players in meal initiation and termination. Glucose-inhibited neurons
decrease their action potential frequency when glucose levels increase from 0.1 to
2.5mM. It must be noted that the responses of the glucose-excited neurons and glucose-
inhibited neurons can be biphasic, because of the existence of presynaptic inputs and
postsynaptic conditions.

Studies have reported that all neurons are silent when they are exposed to extremely low
glucose level below 1mM for more than 12-15min.It was reported that exposure to
0.1mM glucose for over 15min leads to irreversible damage for most neurons.

When plasma glucose level falls below 5mM, a counterregulatory response is initiated,
which consists of activation of α-cell of the pancreas to secrete glucagon and activation
of the sympathoadrenal system. The activation of the sympathoadrenal system increases
plasma epinephrine, norepinephrine, and glucagon, which then stimulate hepatic
glycogenolysis and inhibit insulin secretion by the pancreas. It was shown that the
secretion of counterregulatory hormones, epinephrine and glucagon, is influenced by
neurons in the ventromedial hypothalamic nucleus (VMN). It was pointed out that the
neurons influencing counterregulation are the glucose-excited neurons in the
ventromedial hypothalamic nucleus (VMN).

Glucose sensing in the hypothalamus is important in glucose homeostasis. The glucose-


sensing mechanisms in the hypothalamic neurons are similar to those in pancreatic β-
cells. Glucose signaling in glucose-excited neurons requires glucose uptake via the
type2 glucose transporter, which is followed by glucose phosphorylation by
glucokinase, intramitochondrial glucose oxidation, and an increased cellular adenosine
triphosphate/adenosine diphosphate (ATP/ADP) ratio. This leads to the closure of ATP-
sensitive KATP channels, depolarization of the membrane potential, and influx of Ca++
through voltage-dependent calcium channels, which stimulate neuronal activity and
neurotransmitter release. The role of hypothalamic type2 glucose transporter,
glucokinase, and KATP channels in sensing hypoglycemia and counter-regulatory
hormone responses has been demonstrated. How glucose inhibits neuronal activity in
glucose-inhibited neurons is unclear. One possibility is that glucose increases the
adenosine triphosphate/adenosine diphosphate (ATP/ADP) ratio, which stimulates the
Na+/K+-ATPase pump and triggers hyperpolarizing currents. Alternatively, glucose-
induced activation of ATP-dependent Cl- channels may induce hyperpolarization of the
plasma membrane.

Leptin has an important role in the control of glucose metabolism. Leptin regulates
glucose homeostasis independently of the anorectic effects. The hypothalamus is a key
site of action of leptin-mediated control of glucose metabolism. Leptin signaling in the
hypothalamic arcuate (ARC) nucleus is critical for the maintenance of glucose
homeostasis. Leptin-mediated regulation of glucose metabolism is mediated by
hypothalamic signal transducer and activator of transcription 3 (STAT3) and P13K
signaling pathways. P13K-AKt signaling mediates leptin actions on glucose
homeostasis.

30
Hypothalamic lipid sensing regulates glucose homeostasis via a mechanism involving
the esterification of long chain fatty acids (LCFAs) to LCFA-CoAs, intact KATP channels
and vagal outflow to the liver.

Factors such as rising glucose levels, increased plasma concentrations of insulin, leptin,
and others such as meal consumption generate diverse signals that activate the brain-
centered glucoregulatory system (BCGS), which then take part in the glucose disposal
via insulin-dependent or insulin-independent mechanisms in concert with the pancreatic
islet responses. These revelations lead to a proposal of a two-system control of glucose
homeostasis, which consists of the pancreatic islet responses and the brain-centered
glucoregulatory system (BCGS) responses. Brain-centered glucoregulatory system
(BCGS) function depends on the normal function of the islet. Brain-centered
glucoregulatory system (BCGS) is shown to be relying on inputs from insulin and other
hormones whose secretion depends on normal islet function. Brain-centered
glucoregulatory system (BCGS) is capable of lowering blood glucose levels via insulin-
dependent and insulin-independent mechanisms. The insulin-dependent mechanism
control glucose homeostasis by regulating hepatic glucose production (HGP) through
direct action on hepatocytes and via another proposed indirect mechanism that regulate
hepatic glucose production (HGP) through insulin action at a remote site. Direct action
of insulin on hepatocytes activates insulin receptor substrate phosphatidylinositol 3-OH
(IRS-PI (3)K), which in turn activates AKt (Protein kinase B (PKB), also known as Akt)
that inhibits Forkhead box protein 01 (Fox01), a transcription factor that stimulates
gluconeogenesis and glycogenolysis, the two determinants of hepatic glucose
production (HGP). Another mechanism controlling glucose homeostasis is the insulin-
independent glucose disposal by Brain-centered glucoregulatory system (BCGS).

Figure(9): Brain-centered glucoregulatory system (BCGS) (www.google.com)


31
The astrocyte end-feet may even cover the entire surface of the capillary. The end-feet
show the presence of glucose transporter 1 (GLUT1) type glucose transporters and are
sites of glucose uptake. Astrocytes, on one hand, can communicate with capillaries, and
on the other are associated with neurons and synaptic processes. The metabolic brain
concept is based on this integrated cooperation between astrocytes and neurons.

2.2 Glucose Metabolism


Disturbed glucose metabolism can be the foundation for the development of a large
variety of disorders of the brain.

Most of the energy consumed in the brain is used on synaptic activity. The human cortex
alone requires approximately 3x1023ATP/s/m3, and the energy expenditure to release one
synaptic vesicle is roughly calculated to be 1.64x105molecules ATP. Consequently, a
model of energy use in the brain suggests that considerably larger amount of energy is
spent in the grey matter compared with the white matter. In essence, the brain increases
its utilization of glucose upon activation.

The large blood-to-brain concentration gradient drives the facilitative transport of


glucose across the endothelial membranes via glucose transporter 1 (GLUT1) glucose
transporters into extracellular fluid. The steady-state brain tissue glucose concentration
is about 20% of that in arterial plasma. Glucose transporter 1 (GLUT1) further mediates
glucose uptake from extracellular fluid into astrocytes, oligodendroglia, and microglia,
whereas glucose transporter 3 (GLUT3), which has a much higher transport rate than
glucose transporter 1 (GLUT1), facilitates neuronal glucose uptake. Astrocytes are
generally believed to be involved in the uptake and distribution of brain metabolites.

Hexokinase uses adenosine triphosphate (ATP) to phosphorylate glucose to glucose-6-


phosphate (Glc-6-P) in the first irreversible step of the glycolytic pathway. glucose-6-
phosphate (Glc-6-P) regulates hexokinase activity by feedback inhibition, and it is a
branch-point metabolite that has alternative metabolic fates. glucose-6-phosphate (Glc-
6-P) can continue down the glycolytic pathway to generate pyruvate that can then be
used in mitochondria by oxidative metabolism via the tricarboxylic acid (TCA) cycle. It
can also enter the pentose phosphate shunt pathway (PPP) to generate
dihydronicotinamide-adenine dinucleotide phosphate (NADPH) for management of
oxidative stress and precursors for nucleic acid biosynthesis, and, in astrocytes, it is a
precursor for glycogen. Pentose phosphate shunt pathway (PPP) runs into two steps, first
the oxidative step (irreversible)-wherein dihydronicotinamide-adenine dinucleotide
phosphate (NADPH) is generated, and a second-the non-oxidative step (reversible) –
where Pentose phosphate shunt pathway (PPP) is regulated by glucose-6-phosphate
dehydrogenase (G6PD); the regulation occurs via dihydronicotinamide-adenine
dinucleotide phosphate : nicotinamide adenine dinucleotide phosphate (NADPH:
NADP+) ratio and the allosteric induction by nicotinamide adenine dinucleotide
phosphate (NADP+). Most of the glucose carbon derived from the pentose phosphate
shunt pathway (PPP) re-enters the glycolytic pathway downstream of glucose-6-
phosphate (Glc-6-P). The glycolytic pathway produces a net of 2ATP per molecule of
glucose and oxidation of pyruvate via acetyl coenzymeA (acetyl CoA) in the
tricarboxylic acid (TCA) cycle produces about 30ATP for a total of about 32ATP.

32
Formation of pyruvate from glucose requires regeneration of nicotinamide adenine
dinucleotide (NAD+) from reduced nicotinamide adenine dinucleotide (NADH)
produced by the glyceraldehyde-3-phosphate dehydrogenase reaction by the malate
aspartate shuttle (MAS). Reduced nicotinamide adenine dinucleotide (NADH) cannot
cross the mitochondrial membrane, and the malate aspartate shuttle (MAS) transfers
cytoplasmic reduced nicotinamide adenine dinucleotide (NADH) to the mitochondria
where it is oxidized via the electron transport chain (ETC). When glycolytic flux exceeds
that of the malate aspartate shuttle (MAS) or the tricarboxylic acid (TCA) cycle rate, or
during hypoxic or anoxic conditions. Nicotinamide adenine dinucleotide (NAD+ )is
regulated by the lactate dehydrogenase (LDH) reaction that converts pyruvate to lactate.
Because intracellular accumulation of lactate would cause reversal of the lactate
dehydrogenase (LDH) reaction, lactate must be released from the cell by
monocarboxylic acid transporters (MCT). Exit of lactate eliminates pyruvate as an
oxidizable substrate for that cell and limits the adenosine triphosphate (ATP) yield per
glucose to two.

Figure (10): Major energy pathways in brain (www.google.com)

2.3 Fate of Lactate Generated in Brain


The astrocyte-to-neuron lactate shuttle (ANLS) model is where glucose is transported
from blood vessels to astrocytes by endothelial cells. After entering the astrocyte via
glucose transporter 1 (GLUT1), part of the glucose is metabolized to lactate via pyruvate
by the isoenzyme of lactate dehydrogenase (LDH5). Then lactate is transported outside
the astrocyte through the monocarboxylate (MCT) transporter and is captured by
neurons, also via the monocarboxylate (MCT) transporter. Intracellular lactate in the
neuron is oxidized to pyruvate by the other isoenzyme of lacate hydrogenase (LDH) and
is metabolized along the oxygen pathway. Glucose may be also transported directly into
neuronal cells and penetrates into these cells via glucose transporter 3 (GLUT3). The
astrocyte-to-neuron lactate shuttle (ANLS) model states that Na+-dependent uptake of
neurotransmitter glutamate from the synaptic cleft by astrocytes generates a demand for
2ATP in astrocytes, one to extrude Na+ and one to convert glutamate into glutamine in the
glutamate-glutamine cycle. The model states that this adenosine triphosphate (ATP) is

33
generated by the glycolytic pathway and is associated with release of lactate from
astrocytes and its uptake by nearby neurons where it is oxidized. Thereby astrocyte-
neuron metabolic coupling is linked with the glutamate-glutamine cycle and excitatory
neurotransmission. Thus, during brain activation glycolytic upregulation is stated to
occur in astrocytes-derived lactate providing the major fuel for neurons.

The neuron-to-astrocyte lactate shuttle (NALS) model is based on kinetics of glucose


uptake into brain cells in response to increased metabolic demand and different model
assumptions compared with astrocyte-to-neuron lactate shuttle (ANLS). Here, glucose
is predicted to be predominantly taken up into neurons due to their high energy demand
and the higher transport rate of the neuronal glucose transporter, glucose transporter 3
(GLUT3), compared with the astrocytic glucose transporter, glucose transporter 1
(GLUT1). Lactate is posited to be generated by neurons and taken up by astrocytes. The
lactate release model is based on the observed mismatch between total glucose
utilization and oxidative metabolism and measured lactate release from brain during
brain activation in vivo. If lactate were produced and locally oxidized, total and
oxidative metabolism would be similar in magnitude. However, the rise in oxidative
metabolism varies with experimental condition and pathways stimulated, it is much less
than that of total glucose utilization. Astrocytes have a much faster and greater capacity
for lactate uptake from extracellular fluid, and for lactate dispersal among gap junction-
coupled astrocytes compared with neuronal lactate uptake and shuttling of lactate to
neurons. Astrocytic end feet surround the vasculature, and can discharge lactate to
perivascular fluid for efflux from brain.

Local rates of glucose utilization are driven by functional activities that consume
adenosine triphosphate (ATP) and generate adenosine diphosphate (ADP), which is an
obligatory co-substrate for energy-producing reactions. Glucose metabolism is also the
source for biosynthesis of other compounds required by the brain, including complex
carbohydrates that are components of glycoproteins and glycolipids, amino acids, one-
carbon donors for methylation reactions, and the supply of neurotransmitter precursors.
Both neurons and astrocytes have been described as the main consumers of glucose.

Figure (11): Lactate shuttle in brain (www.google.com)

34
2.3.1The Hypothesis of Lactate Transfer Between Astrocyte and Neuron
Until recently it has been believed that due to their high energy demnands neurons
synthesize energy primarily by the oxidative metabolism of glucose (Krebs cycle and
respiratory chain). However a lot of evidence suggests that neurons can efficiently utilize
lactate and, in addition, have a preference for lactate if both glucose and lactate are
present. It has been shown that the enzyme phosphofructokinase B3, connected with the
glycolysis input pathway for glucose, is practically absent in neurons because of its solid
proteasomal degradation, whereas astrocytes show high levels of expression of this
enzyme. It has been demonstrated that neurons exhibit a slower rate of glycolysis as a
result of low production of fructose-2,6-biphosphate which is the strongest activator of
phosphofructokinases a key enzyme of glycolysis, in contrast to astrocytes, which
indicates an increased amount of activator and thus the rate of glycolysis.

Interestingly, excessive activation of glycolysis in neurons leads to oxidative stress and


apoptosis of neurons. This suggests that neurons cannot afford to maintain a high rate of
glycolysis. At the same time it has been shown that increasing the glucose concentration
starts the hexose monophosphate pathway (HMP) in neurons. It follows that the balance
between the glycolytic pathway and the hexose monophosphate pathway (HMP) must
be strictly maintained in neurons to meet their energy needs, while retaining their
antioxidant status. Accordingly, the use of lactate as a substrate may be a source of an
oxidant to produce large amounts of adenosine triphosphate (ATP), apart from the
glycolytic pathway, which can save neurons glucose for the hexose monophosphate
(HMP) cycle, and thus protect neurons.

Astrocytes display a very high glycolytic activity. Although in comparison with neurons
they have lower rates of oxygen metabolism, they very quickly metabolize glucose via
the glycolytic pathway. The glycolytic nature of astrocytes and their preferences for the
production and release of lactate are also conductive to the production of pyruvate,
which then included in the krebs cycle.

Glucose is transported to astrocytes through glucose transporter 1 (GLUT1) and then


metabolized to lactate. Lactate is transported outside astrocytes and taken up by neurons
by monocarboxylate transporters (MCTs). Intracellular lactate in neurons is oxidized to
pyruvate and metabolized along the oxygen pathway.

It has been shown that the activity of glutamatergic neurons increases the concentration
of extracellular glutamate that is taken up by astrocytes using sodium-dependent
glutamate transporters by increasing the concentration of Na+ in astrocytes. This in turn
activates the Na+ /K+ ATP ase and thereby stimulates glucose uptake and glycolysis in
astrocytes. Then it leads to increased production of lactate, which is released into the
extracellular space, is obtained by neurons via monocarboxylate transporters (MCT) and
then metabolized. The uptake of glutamate by astrocytes increases the use of glucose and
lactate release from the cells. Astrocytes glucose is metabolized to lactate, which is then
used by neurons to maintain their synaptic activity. Many studies support the idea of net
energy transfer from astrocytes to neurons in the form of lactate as a result of
glutamatergic transmission.

35
2.4 Alternative of Glucose as Energy Substrate for The Brain
In the absence of glucose, the brain has alternative energy sources, including ketones
derived from fatty acid metabolism taking place mainly in the liver. These include 3-β-
hydroxybutyrate (3βHM), acetoacetate, and acetone. Ketones play a significant role
especially during the maturation of the brain and deliver 30%-70% of the required
energy to the immature brain. The concentration of ketones in the brain is regulated by
their concentration in the blood and the permeability of the blood-brain barrier, which
depends on the number of monocarboxylic acid transporters (MCTs). Β-
hydroxybutyrate is metabolized primarily in neurons and is converted to glutamate and
glutamine. Ketone bodies are metabolized to acetyl-CoA, then transferred to the Krebs
cycle (TCA) and oxidized in an amount sufficient to satisfy the high metabolic
requirements of the brain. In the mature brain, blood ketone body levels are usually low
and grow mainly as a result of prolonged fasting or a high-fat diet. It is believed that
ketone bodies are able to provide two thirds of the total energy required for the brain
during starvation. They significantly save the resources of glucose, as they can inhibit its
oxidation, probably via the inhibition of pyruvate dehydrogenase complex (PCD). In
this way, a certain amount of glucose can be maintained during prolonged fasting, and
metabolized via glycolysis. The presence of other non-glucose substrates such as
pyruvate or α-ketoglutarate, which can be metabolized without the presence of a
cytosolic nicotinamide adenine dinucleotide (NAD), can preserve the viability of cells.
It has been shown that they significantly reduce pyruvate hypoglycemic neuronal death
and improve the cognitive function of the brain. This indicates the ability of neurons to
adapt to a condition of reduced blood glucose. These metabolic changes allow neurons to
maintain synaptic activity and maintain an appropriate level of adenosine triphosphate
(ATP) for a long time.

Figure (12): Ketones as alternative energy sources (www.google.com)

36
3.Glycogen
Glycogen is the reserve carbohydrate in the animals and it is found in significant amounts
in the liver and muscle. Glycogen is made up of D-glucose residues. Upon hydrolysis, it
yields D-glucose as the product. Glycogen is a highly branched chain polysaccharide.
Glucose residues are linked together through α-1,4-glycosidic linkages except at the
branch point. The branch is linked to the main chain through α-1,6-glycosidic linkages.
Excess glucose is stored not as monomer but converted to polymeric forms for storage in
the liver and skeletal muscle but can also be made by the brain, uterus, and the vagina.
However, muscle glycogen is not generally available to other tissues, because muscle
lacks the enzyme glucose-6-phosphatase. Stores of glycogen in the liver are considered
the main buffer of blood glucose levels. The major site of daily glucose consumption
(75%) is the brain via aerobic pathways. Most of the remainder of it is utilized by
erythrocytes, skeletal muscle, and heart muscle.

Glycogen is not evenly distributed throughout the brain. Microscopic examinations


show that the concentrations of glycogen are highest in regions with the highest synaptic
density suggesting its role in synaptic transmission, with the concentrations in gray
matter about two times greater than in white matter. High levels of glycogen can be
found in the medulla oblongata, pons, cerebellum, hippocampus, hypothalamus,
thalamus, cortex, and striatum.

Figure (13): Chemical structure of glycogen (www.google.com)

3.1Glycogen Metabolism
Glycogen degradation and synthesis are relatively simple biochemical processes.
Glycogen synthesis differs from glycogen breakdown. Unlike breakdown, synthesis is
endergonic, meaning that glycogen is not synthesized without input of energy. Energy
for glycogen synthesis comes from Uridine-5'-triphosphate (UTP), which reacts with
glucose-1-phosphate, forming Uridine diphosphate glucose (UDP-glucose), in reaction
catalyzed by UDP-glucose pyrophosphorylase. Glycogen is synthesized from
monomers of UDP-glucose by the enzyme Glycogen synthase, which progressively
lengthens the glycogen chain. As glycogen synthase can only lengthen an existing chain,
the protein glycogenin is needed to initiate the synthesis of glycogen.

Glycogen degradation consists of three steps:


1. The release of glucose-1-phosphate from glycogen.
2. The remodeling of the glycogen substrate to permit further degradation.
37
3. The conversion of glucose-1-phosphate into glucose-6-phosphate for further
metabolism.

Glycogen is cleaved from the nonreducing ends of the chain by the enzyme glycogen
phosphorylase to produce monomers of glucose-1-phosphate that is then converted to
glucose-6-phosphate. A special debranching enzyme is needed to remove the α(1-6)
branches in branched glycogen and reshape the chain into linear polymer. Debranching
enzyme has two independent active sites, consisting of residues in different segments of
a single polypeptide chain, that catalyze α (1-6) glucosidase and transferase
(transglycosylase) reactions. The transferase of the debranching enzyme transfers three
glucose residues from a 4-residue limit branch, diminishing the limit branch to a single
glucose residue. The α(1-6) glucosidase moiety of the debranching enzyme then
catalyzes hydrolysis of the α(1-6) linkage, yielding free glucose. This is a minor fraction
of glucose released from glycogen. The major product of glycogen breakdown is
glucose-1-phosphate, arising from phosphorylase activity, which is subsequently
converted to glucose-6-phosphate. The glucose-6-phosphate monomers produced have
three possible fates:

1. Glucose-6-phosphate can continue on the glycolysis pathway and be used as fuel.


2. Glucose-6-phosphate can enter the pentose phosphate pathway via the enzyme
glucose-6-phosphate dehydrogenase to produce dihydronicotinamide-adenine
dinucleotide phosphate (NADPH) and 5-carbon sugars.
3. In the liver and kidney, glucose-6-phosphate can be dephosphorylated back to
glucose by the enzyme glucose-6-phosphatase. This is the final step in the
gluconeogenesis pathway.

Brain glycogen is metabolized by enzymes located in astrocytes, such as glycogen


phosphorylase (GP) and glycogen synthase (GYS). The brain isoform of glycogen
phosphorylase (GP) occurs mainly in astrocytes, but interestingly it can also be found in
several other cell types, such as the choroid plexus cells and ependymal cells. glycogen
synthase (GYS) also occurs in neurons. glycogen synthase (GYS) is mainly expressed in
the hippocampus, cerebellum, and olfactory bulbs, occurring in both inactive
phosphorylated (GYSb) and active dephosphorylaed (GYSa) forms, which allows a
precise regulation of glycogen metabolism. The transformations of phosphorylated/
dephosphorylated synthases are controlled by a whole family of phosphatases.

3.1.1Glycogen Metabolism in Astrocytes


Glucose is transported via the glucose transporter 1 (GLUT1) and possibly the insulin-
sensitive transporter 4 (GLUT4). Glucose is phosphorylated by hexokinase (HK) to
glucose-6-phosphate (G-6-P) , which is subsequently converted to glucose-1-phosphate
(g-1-P) by phosphoglucomutase (PGM) and then to uridine diphosphate glucose (UDP
glucose) by UDP glucose pyrophosphate (UDPGPP). The uridine diphosphate glucose
(UDP glucose) continues on to glycogen synthesis via the actions of glycogen synthase
(GYS), which can exist in two forms; the active dephosphorylated form (GYSa) or the
inactive phosphorylated form (GYSb). Protein phosphatase 1 (PP1) converts inactive
phosphorylated form (GYSb) to active dephosphorylated form (GYSa) via the
regulatory subunit Protein Targeting to Glycogen (PTG), resulting in glycogen
formation.
38
Figure (14): Glycogen synthesis (www.google.com)

Glycogen is broken down by glycogen phosphorylase (GP) which similar to glycogen


synthase exists in two forms: the active phosphorylated form (GPa), or the inactive
dephosphorylated form (GPb). Phosphorylase kinase (PK) dephosphorylates GPb to the
active form. A major glycogen-derived product is lactate which is transported into the
extracellular space via monocarboxylate transporters (MCT).

Figure (15): Brain glycogenolysis (www.google.com)

39
Among other things, glycogen levels depend on hormones such as adrenaline and
noradrenaline which have a glycogenolytic action, and insulin, promoting the synthesis
of glycogen. Insulin-like growth factor-I and II (IGFI, IGFII) as well as insulin, may
increase the levels of brain glycogen by their influence on insulin receptors.

4.Glycoproteins
Glycoproteins are proteins to which oligosaccharides are covalently attached. The
length of the glycoprotein's carbohydrate chain is relatively short (usually 2-10 sugar
residues in length, although they can be longer). The glycoprotein carbohydrate chains
are often branched instead of linear, and may or may not be negatively charged.

Membrane-bound glycoproteins participate in a broad range of cellular phenomena,


including cell surface recognition (by other cells, hormones, and viruses), cell surface
antigenicity (such as the blood group antigens), and as components of the extracellular
matrix and of the mucins of the gastrointestinal and urogenital tracts, where they act as
protective biologic lubricants. In addition, almost all of the globular proteins present in
human plasma are glycoproteins.

The oligosaccharide components of glycoproteins are generally branched


heteropolymers composed primarily of D-hexoses, with the addition in some cases of
neuraminic acid, and of L-fructose-a 6-deoxyhexose. The oligosaccharide may be
attached to the protein through an N- or an O- glycosidic link. In the former case, the
sugar chain is attached to the amide group of an aspargine side chain, and in the latter
case, to a hydroxyl group of either a serine or threonine R-group. A glycoprotein may
contain only one type of glycosidic linkage (N- or O- linke), or may have both O- and N-
linked oligosaccharides within the same molecule.

Figure (16): General structure of glycoproteins (www.google.com)

4.1 Glycoproteins in Brain


There is emerging evidence that saccharides are important for the growth and
development of the brain. Indeed, there appears to be some structural features and use of
saccharides in the formation of glycoconjugates, specifically, glycoproteins that are
unique to the brain, leading to the classification of some glycoproteins as being brain
specific. While the structural and functional role of saccharides in the brain remains to be
clarified, it is known from research in animals and humans that the concentration of
40
glycoproteins in the brain increases during postnatal development. Specifically, age
related increases in the levels of glucosamine, mannose, and galactose contained in
glycoproteins have been found in the developing rat brain, with concentrations of
glycoproteins in the human brain increasing from 12 weeks before to 8 months after
birth. Conversely, in the human brain, it appears that there is an age-related decline in
glycoprotein concentrations. For example, the concentration of glycoproteins in the
cerebral gray matter of 3-year-old human brain tissue has been shown to be between 7
and 15 times greater than the concentration in 8-year-old tissue and 15 to 30 times greater
than in 72-year old tissue.

4.2 Synthesis of Glycoproteins


Glycoproteins are synthesized on ribosomes attached to the rough endoplasmic
reticulum (RER). These glycoproteins contain specific signal sequences that act as
molecular "address labels", which direct the proteins to their proper destinations. A N-
terminal hydrophobic sequence initially directs these glycoproteins to the rough
endoplasmic reticulum (RER), allowing the growing polypeptide to be extruded into the
lumen. The glycoproteins are then transported via secretory vesicles to the Golgi
complex, which acts as a sorting center. In the Golgi these glycoproteins that are to be
secreted from the cell (or are targeted for lysosomes) are free in the lumen, whereas those
that are destined to become components of the cell membrane are integrated into the
Golgi membrane, either releasing the free glycoproteins, or adding the membrane-
bound proteins of the vesicle to the cell membrane. The membrane glycoproteins are
thus oriented with the carbohydrate portion on the outside of the cell.

4.2.1 Carbohydrate Components of Glycoproteins


The precursors of the carbohydrate components of glycoproteins are nucleotide sugars,
which include uridine diphosphate-glucose (UDP-glucose), uridine diphosphate-
galactose (UDP-galactose), uridine diphosphate –N-acetylglucosamine (UDP-N-
acetylglucosamine), and uridine diphosphate-N-acetylgalactosamine (UDP-N-
acetylgalactosamine. In addition, guanosine diphospho-mannose (GDP-mannose),
guanosine diphospho-L-fructose (GDP-L-fructose), and cytidine monophosphate-N-
acetylneuraminic acid (CMP-N-acetylneuraminic acid) may donate sugars to the
growing chain. The oligosaccharides are covalently attached to specific amino acid R-
groups of the protein, where the three dimensional structure of the protein determines
whether or not a specific amino acid R-group is glycosylated.

4.2.2 Synthesis of The N-Linked Glycosides


The synthesis of N-linked glycosides occurs in the lumen of the endoplasmic reticulum
(ER). These structures require the participation of a lipid (dolichol) and its
phosphorylated derivative, dolichol pyrophosphate. Protein is synthesized on the rough
endoplasmic reticulum (RER) and enters its lumen. A lipid-linked oligosaccharide is
first constructed. This consists of dolichol attached through a pyrophosphate linkage to
an oligosaccharide containing N-acetylglucosamine, mannose, and glucose. The sugars
to be added to the dolichol by the membrane-bound glycosyltransferases are first N-
acetylglucosamine, followed by mannose and glucose. The oligosaccharide is
transferred from the dolichol to an aspargine side group of the protein by a protein-
oligosaccharide transferase present in the endoplasmic reticulum (ER). After

41
incorporation into the protein, the N-linked oligosaccharide is processed by the removal
of specific mannosyl and glucosyl residues as the glycoprotein moves through the
endoplasmic reticulum (ER). Finally, the oligosaccharide chains are completed in the
Golgi by the addition of a variety of sugars to produce a complex glycoprotein, or they
are not processed further, leaving branched, mannose-containing chains in a high-
mannose glycoprotein.

Figure (17): Oligosaccharide synthesis of glycoproteins (www.google.com)

Figure (18): Synthesis and processing of N-linked glycoprotein in the


endoplasmic reticulum (www.google.com)

4.2.3 Synthesis of The O-Linked Glycosides


First, the protein to which the oligosaccharides are to be attached is synthesized on the
rough endoplasmic reticulum (RER), and extruded into its lumen. Glycosylation begins
with the transfer of a N-acetylgalactosamine [from uridine diphosphate-N-
42
acetylgalactosamine (UDP-N-acetylgalactosamine)] onto a specific seryl or threonyl R-
group.

Figure (19): Synthesis of O-linked glycoprotein in the endoplasmic reticulum


(www.google.com)

Figure (20): Differences between O- and N- linked glycoprotein synthesis


(www.google.com)

43
4.3 Lysosomal Degradation of Glycoproteins
The lysosomal hydrolytic enzymes are each generally specific for the removal of one
component of the glycoprotein. They are exoenzymes that remove their respective
groups in sequence in the reverse order of their incorporation. If any degenerative
enzyme is missing, degradation by other exoenzymes cannot continue.

Figure (21): Sequential degradation of glycoproteins (www.google.com)

For further readings:


1. Best, T.; Kemps, E.; Bryan, J. (2005). Effects of saccharides on brain function and
cognitive performance. Nutrition Reviews, 63(12):409-418.
2. Champe, P.; Harvey, R.; Ferrier, D. (2008). Biochemistry. 4th edition. Lippincott
Williams & Wilkins. Philadelphia.
3. Falkowska, A.; Gutowska, I.; Goschorska, M.; Nowacki, P.; Chlubek, D.;
Baranowska-Bosiacka, I. (2015). Int. J. Mol. Sci., 16:25959-25981.
4. Harris, JJ.; Attwell, D. (2012). The energetics of CNS white matter. J Neurosci,
32:356-371.
5. Khowala, S.; Verma, D.; Banik, S. (2008). Biomolecules: (introduction, structure &
function) carbohydrates.
6. Lyngdoh, J.; Marbaniang, E.; Lynrah, K.; Lyngdoh, M. (2015). The role of brain in
44
the regulation of glucose homeostasis. International Journal of Medical Science
and Public Health, 4(11):1477-1480.
7. Marty, N.; Dallaporta, M.; Thorens, B. 92007). Brain glucose sensing
counterregulation and energy homeostasis. Physiology, 22:241-251.
8. Mergenthaler, P.; Lindauer, U.; Dienel, G.; Meisel, A. (2013). Sugar for the brain:
the role of glucose in physiological and pathological brain function. Trends
Neurosci, 36(10):587-597.
9. Nehlig, A. (2004). Brain uptake and metabolism of ketone bodies in animal models.
Prostaglandins Leukot. Essent. Fat. Acids, 70:265-275.
10. Roh, E.; Song,D.; Kim, MS. (2016). Emerging role of the brain in the homeostatic
regulation of energy and glucose metabolism. Experimental & Molecular
Medicine, 48.
11. Routh, VH. (2002). Glucose sensing neurons : are they physiologically relevant?.
Physiol Behav, 76:403-413.
12. Schaafsma, I. (2007). Sugar and behavior.
13. Schwartz, MW.; Porte, D. (2005). Diabetes, obesity, and brain. Science, 307:375-
379.
14. Song, Z.; Levin, BE.; McArdle, JJ.; Bakhos, N.; Routh, VH. (2001). Convergence
of pre-and postsynaptic influences on glucosensing neurons in the ventromedial
hypothalamic nucleus. Diabetes, 50(12):2673-2681.

45
CHAPTER THREE
SACCHARIDES DISORDERS AND BRAIN DISEASES

1.Pancreatic β Cells
The pancreas is a fish-shaped organ that stretches across the back of the abdomen behind
the stomach. Within the pancreas there are areas that are called the islets of Langerhans.
The beta cells constitute the predominant type of cells in the islets. They make up
65–80% of the cells in the islets.

The primary function of a beta cell is to store and release insulin. Insulin is a hormone
that brings about effects which reduce blood glucose concentration. Beta cells can
respond quickly to spikes in blood glucose concentrations by secreting some of their
stored insulin while simultaneously producing more.

Voltage-gated calcium channels and ATP-sensitive potassium ion channels are


embedded in the cell surface membrane of beta cells. These ATP-sensitive potassium ion
channels are normally open and the calcium ion channels are normally closed.
Potassium ions diffuse out of the cell, down their concentration gradient, making the
inside of the cell more negative with respect to the outside (as potassium ions carry a
positive charge). At rest, this creates a potential difference across the cell surface
membrane of -70mV.

When the glucose concentration outside the cell is high, glucose molecules move into the
cell by facilitated diffusion, down its concentration gradient through the glucose
transporter 2 (GLUT2) . Since beta cells use glucokinase to catalyze the first step of
glycolysis, metabolism only occurs around physiological blood glucose levels and
above. Metabolism of the glucose produces adenosine triphosphate (ATP), which
increases the adenosine triphosphate (ATP) to adenosine diphosphate (ADP) ratio.

The ATP-sensitive potassium ion channels close when this ratio rises. This means that
potassium ions can no longer diffuse out of the cell. As a result, the potential difference
across the membrane becomes more positive (as potassium ions accumulate inside the
cell). This change in potential difference opens the voltage-gated calcium channels,
which allows calcium ions from outside the cell to diffuse in down their concentration
gradient. When the calcium ions enter the cell, they cause vesicles containing insulin to
move to, and fuse with, the cell surface membrane, releasing insulin by exocytosis.

It is hypothesized that the occurrence of a decrease in the mass of these cells, observed
even in subjects with normal glucose tolerance, may contribute to the reduced secretion.
This decrease is caused by apoptosis due to the presence of high blood glucose
(glucotoxicity) or a large amount of free fatty acids (lipotoxicity). In addition, it is
observed that amyloid plaques are present in the β cells of individuals with diabetes and

46
these plaques are able to destroy cells and eliminate their secretory activity. These
plaques consist of sets of islet associated polypeptide, that emerge from a normal protein
which is co-secreted by the β-cells with insulin and is maintained in the granules of
insulin. A complementary explanation for the β-cell apoptosis is due to toxic oxygen
species (mainly produced in the mitochondria), which are excessively produced during
the course of disease.

Figure (22): β cell glucose sensing via GLUT2 transporter (www.google.com)

2.Insulin Resistance
Insulin is a critical polypeptide hormone, produced by pancreatic beta islet cells,
secreted in response to elevated blood glucose. As a key anabolic hormone, it readily
binds to cell membrane receptors affecting a multitude of cellular and systemic
processes, including glucose uptake, glycogen synthesis, gluconeogenesis, lipid
metabolism, hunger, cell growth, and gene expression.

Animal and in vitro studies have suggested a role for insulin in the regulation of brain
dopaminergic activity featuring a reciprocal regulation between the two chemicals.

Insulin has significant neurothropic properties in the brain. The hormone is rapidly
transported to the level of the central nervous system (CNS) through the blood- brain
barrier (BBB) by a transport mechanism mediated by insulin receptors. It is interesting to
note that these receptors are mainly localized at the level of the hippocampus, entorhinal
cortex, and frontal areas known to be involved in functions such as memory and learning.
Insulin is also involved in the production of important neurotransmitters such as
acetylcholine and norepinephrine.

Several studies have shown the presence of insulin receptors in numerous brain regions
such as the cerebral cortex, choroid plexus, hypothalamus, hippocampus, olfactory
regions, amygdaloid complex, entorhinal cortex, cerebellum, among other regions.
Despite that prevalence of receptors, the role of insulin in the brain is far from being
completely understood. It is reported that insulin acts in the food intake and body weight.
47
Other actions are also mentioned in neurotrophic role, increase of activity of choline
acetyl transferase, influence on the development of cholinergic and dopaminergic
neurons and increase in neurotransmitters release. It was demonstrated the presence of
insulin receptor substrate, tyrosine kinase P53-P58, and insulin receptor in the synapses
within the hippocampus and cerebellum, suggesting a signaling role for insulin.

A state of chronic hyperinsulinemia, as it occurs in insulin-resistance conditions and


type2 diabetes mellitus may determine a downregulation of the insulin receptors at the
blood-brain barrier, thus reducing the transport of insulin in the brain.

Insulin and insulin like growth factor-1 (IGF-1) mediate their effects by activating
complex intracellular signaling pathways starting with ligand binding to cell surface
receptors, followed by autophosphorylation and activation of the intrinsic receptor
tyrosine kinases. Insulin / IGF-1 receptor tyrosine kinases phosphorylates insulin
receptor substrate molecules, which transmit signals downstream by activating the
extracellular signal-related kinase/mitogen-activated protein kinase (ERK/MAPK) and
PI3 kinase/Akt pathways, and inhibit glycogen synthase kinase 3β (GSK-3β).

Major biological responses to signaling through insulin receptor substrate molecules


include increased cell growth, survival, energy metabolism, and cholinergic gene
expression, and inhibition of oxidative stress and apoptosis. These very same signaling
pathways are activated in various cell types, tissues, and target organs that express
insulin and universal. Moreover, these pathways are phylogenetically conserved and
have critical roles in regulating development, growth, survival, senescence,
carcinogenesis, and neurodegeneration.

Insulin resistance is clinically defined as the inability of a known quantity of exogenous


or endogenous insulin to increase glucose uptake and utilization in an individual as much
as it does in normal population. Insulin resistance is the event which consists of a poor
transport of glucose from the blood vessels to target tissues, causing an increase in the
concentration of blood glucose. This resistance to insulin is observed in adipocytes, in
the liver, and especially in muscle. Several factors have been suggested as responsible
for increased insulin resistance by target tissues. It is known that adipocytes from
individuals with diabetes or obesity are resistant to anti-lipolytic activity of insulin,
which contributes to the increase of free fatty acids in plasma. Studies have shown that
elevated concentrations of plasma free fatty acids increases the insulin resistance in
muscle and liver and also decrease insulin secretion. Furthermore, the accumulation of
triglycerides in liver and muscle appears to contribute to the insulin resistance. This has
recently been called lipotoxicity. Other studies have shown that a decrease in the number
of mitochondria in muscle or an impaired function of these may contribute to increased
insulin resistance. At the cellular level, proteins are part of the signaling cascade
responsible for the influx of glucose in the target tissues are also indicated as having
influence on the increase of insulin resistance. The insulin receptor substrate plays an
important role in the cascade of signaling events that occurs after insulin binding to
receptor in muscle. In individuals with diabetes, the phosphorylation of the insulin
receptor substrate and its activity are decreased, which is highly related to a decrease in
glucose transport into the muscle.

48
Insulin resistance is characterized by decreased responsiveness of cells to the hormone,
with associated hyperglycemia and hyperinsulinemia. Insulin resistance is a central
feature of type2 diabetes, though type2 diabetes involves many pathophysiological
processes beyond insulin resistance. In type2 diabetes, insulin secretion initially rises,
compensating for insulin resistance, but over time pancreatic beta islet cell dysfunction
progresses, insulin secretion decreases, and hyperglycemia worsens. Different cell types
and tissues react variably to insulin, so the degree and distribution of insulin resistance in
tissues is not uniform. Thus, brain insulin resistance could exist in tandem with, or
independently from peripheral insulin resistance and type2 diabetes. The majority of
insulin is thought to enter the brain from the periphery through a saturable transport
system of the blood-brain barrier (BBB) with some variation in permeability in different
regions of the brain. There is also significant evidence indicating at least some degree of
insulin production within the brain parenchyma.

The development of insulin resistance in humans is closely correlated with immune


infiltration and inflammation, with clear links between the development of obesity, type
2 diabetes mellitus, and cardiovascular disease. Evidence for mitochondrial dysfunction
in the development of insulin resistance has been obtained through measuring rates of in
vivo mitochondrial phosphorylation using proton1 (H) magnetic resonance
spectroscopy in the relatives to patients with type 2 diabetes mellitus, finding that rates of
mitochondrial adenosine triphosphate (ATP) production are reduced by 30% in the
muscle of lean pre-diabetic insulin resistant species. The more widespread development
of insulin resistance has also been proposed to be mediated through mitochondrial
dysfunction at least to a partial extent through muscle expression of transcriptional
regulators including peroxisome proliferator-activated receptor gamma coactivator 1-
alpha (PGC1 α), an important regulator of enzymes involved in mitochondrial
respiration. The earliest sign of insulin resistance was shown to be reduction in
expression of peroxisome proliferator-activated receptor gamma coactivator 1-alpha
(PGC1 α) and the mitochondrial gene nuclear respiratory factor1 (NRF1). In a whole
genome methylation analysis of skeletal muscle from type2 diabetes mellitus and
control subjects. It is well known that mutations in mitochondrial deoxyribonucleic acid
(DNA) can lead to a wide variety of phenotypes that commonly involve
neurodegeneration and diabetes. Normal mitochondrial biogenesis, respiration, and
metabolism of reactive oxygen species require intact expression of both nuclear and
mitochondrial encoded genomes now recognized as being regulated by peroxisome
proliferator-activated receptor gamma coactivator 1-alpha (PGC1α). It has been shown
that peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC1α)
has a powerful suppressive effect on reactive oxygen species production, in parallel to its
effects in elevating mitochondrial respiration. Insulin resistant patients show reduced
expression of peroxisome proliferator-activated receptor gamma coactivator 1-alpha
(PGC1α) and the mitochondrial encoded gene cyclooxygenase-1(COX1) while
reduction in peroxisome proliferator-activated receptor gamma coactivator 1-alpha
(PGC1α)-responsive genes has been shown among patients with type2 diabetes mellitus
and their asymptomatic relatives compared with healthy controls

49
Figure (23): PGC1α (www.google.com)

Figure (24): Insulin in the brain (www.google.com)

3.Islet Amyloid Polypeptide


Islet amyloid polypeptide (IAPP), otherwise known as amylin, is a hormone co-secreted
with insulin from pancreatic islet beta cells in the postprandial state. Islet amyloid
polypeptide (IAPP) helps regulate blood sugar by inducing satiation, delaying gastric
emptying and inhibiting glucagon secretion. In addition to peripheral effects, islet
amyloid polypeptide (IAPP) readily crosses the blood- brain barrier (BBB) and acts

50
centrally with receptors found in the area postrema and other locations in the brainstem.
As islet amyloid polypeptide (IAPP) is secreted with insulin, its blood levels become
elevated in those with insulin resistance and type2 diabetes mellitus. Islet amyloid
polypeptide (IAPP) is toxic to pancreatic beta islet cells. Islet amyloid polypeptide
(IAPP) deposition may also negatively affect the brain and its vasculature however this
remains unknown. Postmortem data has shown islet amyloid polypeptide (IAPP)
deposition in the temporal lobe grey matter in those with diabetes but not in controls,
with some degree of islet amyloid polypeptide (IAPP) and amyloid beta (Aβ) peptide
colonization.

Figure (25):Islet amyloid polypeptide (IAPP) apoptotic pathways of β cells of


type2 diabetes mellitus patients (www.google.com)

4.Diabetes Mellitus
Diabetes mellitus is a group of metabolic diseases characterized by hyperglycemia
resulting from defects in insulin secretion, insulin action, or both. Absence or reduced
insulin in turn leads to persistent abnormally high blood sugar and glucose intolerance.

In normoglycemic individuals in a fasting situation, there are a production of glucose by


the liver which accounts for approximately 85% of the glucose produced by the full
body. This glucose is used by the nervous system, internalized by the liver and
gastrointestinal tract and also by muscle. When glucose ingestion occurs, insulin is
secreted and inhibits hepatic glucose production. During diabetes, insulin resistance
occurs in the liver, and has not been observed a decrease in endogenous glucose
production. In a fasting situation, this can still be compensated by the high amount of
insulin secreted in patients with mild diabetes (fasting hyperglycemia ≤140mg/dl). In
subjects with moderate diabetes (fasting hyperglycemia of 140-200mg/dl) there is
increase in endogenous glucose production, which further increases the glucose
concentration in blood during fasting. The production of glucose by the liver is primarily
by gluconeogenesis or glycogeneolysis. Some risk factors are contributors to an increase
51
in gluconeogenesis as hyperglucagonemia increased sensitivity to glucagon, increased
fatty acids oxidation, and decreased insulin sensitivity. Some studies also showed
increased activity of enzymes that regulate gluconeogenesis and glucose output from the
liver.

Chronic hyperglycemia leads to oxidative stress and the production of reactive oxygen
species, which has been implicated in the ongoing β cell death in type2 diabetes mellitus.
While production of reactive oxygen species may also be a mechanism underlying
dopaminergic cell loss in hyperglycaemic animals, this remains speculation, and a
further possibility emerges from work demonstrating that in both insulin deficient rats
and insulin-resistant mice, diabetes impairs hippocampus-dependent memory, synaptic
plasticity, and adult neurogenesis.

Glucose metabolism has important implications for the functioning of the brain.
Diabetes has been known to have an effect on the brain for more than one hundred years.
In the event of disturbances, e.g. in diabetes mellitus, conditions of glucose delivery to
the brain may become disrupted. Chronic hyperglycemia in diabetes is associated with
abnormalities in many organs, for example microvasculature and macrovascular
changes in the brain. In diabetes both hypo-and hyperglycemia may contribute to
microangiopathy and its complications. Diabetes is a complex metabolic syndrome that
disrupts both signaling and metabolic pathways, which in turn affect the transport of
glucose. The predominant transporter proteins (GLUT) involved in cerebral glucose
utilization are glucose transporter 1 (GLUT1) and glucose transporter 3(GLUT3) with
glucose transporter 1 (GLUT1) present in all brain cells including the endothelial cells of
the capillaries (with very low neuronal expression in vivo), and glucose transporter
3(GLUT3) almost restricted to neurons. Glucose transporter 1 (GLUT1) is mainly
responsible for the facilitative transport of glucose across the blood-brain barrier (BBB).
The blood-brain barrier (BBB) which was considered to behave as a single transport
step, separates the blood circulation compartment from the brain aqueous phase that is
virtually separated from the metabolic pool where glucose is consumed. The transport of
glucose to the brain is downregulated in chronic hyperglycemia as well as during
experimentally induced hyperglycemia.

The activity of glycolysis differs between type1 and type2 diabetes. Some data show that
glycolytic activity is reduced in type1 diabetes, while other studies show that at least
initially glycolytic activity increases to a certain extent. The fact that glucose utilization
initially increases suggests an increase in the oxidative metabolism of glucose through
the tricarboxylic acid (TCA) cycle. In contrast, in type2 diabetes, tricarboxylic acid
(TCA) cycle activity is reduced. One of the subunits of the pyruvate dehydrogenase
complex (PDC) enzyme complex is downregulated.

Lactate concentration is reduced in the cerebral cortex in type 2 diabetes, in connection


with the inhibition of glycogen degradation. Lower amounts of lactate are also due to the
presence of a glycogen phosphorylase inhibitor. Glycolysis is activated in the presence
of glycogen phosphorylase inhibitor.

Neurons are not able to generate intermediates of the tricarboxylic acid (TCA) cycle or
glutamate from glucose, which shows the importance of a continuous supply of
precursor compounds from the astrocytes.
52
Summing up the impact of diabetes on metabolic processes, the brain appears to use
compensatory mechanisms to ensure an adequate supply of glucose in a diabetes state.
However, studies indicate that glucose transport itself and cellular metabolic
interactions are weakened. In addition, diabetes is associated with an alteration of both
turnover and the activity of key enzymes involved in glycogen metabolism. For
example, the amount of newly synthesized glycogen is reduced in type1 diabetes,
suggesting that the activity of glycogen synthase may be decreased. Besides, energy
homeostasis and homeostasis associated with neurotransmission may be disrupted.

Chronic hyperglycemia may result in cerebral metabolic alterations and central nervous
system (CNS) injury. Hyperglycemia induced (or associated) metabolic and vascular
disturbances are known and may increase the risk of stroke, seizures, diabetic
encephalopathy and cognitive compromise. These pathological conditions may result
from alterations in cerebral energy homeostasis and metabolism possibly through
mechanisms including changes in osmolar gradients in hyperglycemia, hormonal
regulation, glucose utilization, oxidative stress, and the levels of ketone bodies.

Hyperglycemia leads to osmolar gradients across cell membranes, triggering alterations


in cell volume regulation that shift water from the intracellular fluid space to the
extracellular fluid. The osmolar gradients due to the elevated glucose levels in
hyperglycemia may have led to increased levels of intracellular osmolytes including
glycerophosphocholine (GPC), Ins and Tau to promote the maintenance of neuronal and
glial cell volume. Glycemic normalization was able to restore the glutathione (GSH)
level but not that of N-Acetylaspartic acid (NAA), indicating potential irreversible
neuronal damage due to prolonged hyperglycemia.

Figure (26): pathophysiology of type1 and type2 diabetes mellitus


(www.google.com)
53
4.1 Type 1 Diabetes Mellitus
Type 1 diabetes mellitus (T1DM) is also known as juvenile diabetes. Type 1 diabetes
mellitus (T1DM) is an absolute deficiency of insulin secretion. Type 1 diabetes mellitus
(T1DM) is an autoimmune disorder resulting from a cellular mediated autoimmune
destruction of pancreatic β cells. Type 1 diabetes mellitus (T1DM) is the most common
type of the diabetes in people with lower age groups. Type 1 diabetes mellitus (T1DM) is
commonly diagnosed during childhood and adolescence. This is a period of rapid
developmental changes in the central nervous system and there has been concern that the
younger brain may be more susceptible to extremes of glycemia. The prevalence of type
1 diabetes mellitus (T1DM) is increasing in both prosperous and poor countries.

Etiology of type1 diabetes mellitus involves:


1- Autoimmune response: it is expressed as a consequence autoimmune disease,
where the β cells of pancreas are slowly demolished by the body's own immune
system which reduces insulin production. In the development of type 1 diabetes
mellitus (T1DM) both the environmental and genetic predisposing factors are
significant, but the exact correlation is still unknown.
2- Genetic factors: investigators have commenced at least eighteen genetic positions
are designated as IDDM1-IDDM18 (insulin dependent diabetes mellitus), which
are related to type 1 diabetes. The IDDM1 (insulin dependent diabetes mellitus 1)
region contains the HLA (human leukocyte antigen) genes that encode proteins
called major histocompatibility complex (MHC). In this location immune
responses are affected by these genes. Other chromosomes and genes continue to be
identified. We can predict the onset of clinical diabetes as much as three years by the
appearance of islet cell antibodies (ICAs) frequently.
3- Environmental factors: due to abrupt stress like an infection where the β cells of
pancreas falls below 5-10%. Coxsackie viruses are a family of enteric viruses
which attack the intestinal tract leads to the destruction of insulin producing
pancreatic β cells.

4.2 Type 2 Diabetes Mellitus


Type 2 diabetes mellitus (T2DM) is a chronic disease characterized by progressive
insulin resistance and hyperglycemia with major complications as retinopathy,
neuropathy, and cerebrovascular disease (CVD). Three pathophysiological events are
classically identified as responsible for the development of type 2 diabetes mellitus
(T2DM): insulin resistance, β cell secretory dysfunction, and increased glucose
production by the liver. About 85-95% of type 2 diabetes mellitus (T2DM) is
predominant in developing countries. Type 2 diabetes mellitus (T2DM) is caused by
insulin resistance in peripheral tissues and is most frequently associated with aging, a
family history of diabetes, obesity, and failure to exercise. Type 2 diabetes mellitus
(T2DM) is associated with a strong genetic predisposition. It has been estimated that
about 1.5% of type 2 diabetes mellitus (T2DM) is attributable to the mitochondrial
A3243G mutation (the cause of mitochondrial encephalopathy, lactic acidosis, and
stroke-like episodes). Insulin resistance in type 2 diabetes mellitus (T2DM) is partly
mediated by reduced insulin receptor expression, insulin receptor tyrosine kinase
activity, insulin receptor substrate (IRS) type 1 expression, and/or phosphatidyl-
inositol-3 (PI3) kinase activation in skeletal muscle and adipocytes.

54
4.3 Type 3 Diabetes Mellitus
Type 3 diabetes mellitus (T3DM) corresponds to a chronic insulin resistance plus insulin
deficiency state that is largely confined to the brain but can overlap with type 2 diabetes
mellitus (T2DM). It has been proposed that type 3 diabetes mellitus (T3DM) represents a
major pathogenic mechanism of Alzheimer disease (AD) neurodegeneration.

4.4 Gestational Diabetes Mellitus


Gestational diabetes is a temporary and appears during pregnancy usually develops
during third trimester of pregnancy. After delivery blood glucose levels generally return
to normal. Gestational diabetes is pregnancy associated and caused by insulin deficiency
and hyperglycemia. Approximately 87.5% of maternal diabetes are due to gestational
diabetes mellitus (GDM), defined as glucose intolerance with onset or first recognition
during pregnancy. Maternal diabetes has been associated with an increased risk of poor
neurodevelopmental outcomes in the offspring. Maternal hyperglycemia can result in
hypoxia in the fetus, and a depleted oxygen supply to the fetus may impair
neurodevelopment and thus contribute to a greater risk of ASD (autism spectrum
disorder). Maternal altered autoimmunity because of hyperglycemia may influence
brain development in the offspring either by creating a hostile intrauterine environment
or by modifying the offspring's autoimmunity in early development through
immunoglobulin G. Epigenetic modification by hyperglycemia may also be implicated
in the pathogenesis of ASD (autism spectrum disorder), but the current evidence is still
sparse.

Figure (27): Gestational Hyperglycemia (www.google.com)

5.Hypoglycemia
Hypoglycemia, also known as low blood sugar, is when blood sugar decreases to below
normal levels. This may result in a variety of symptoms including clumsiness, trouble
talking, confusion, loss of consciousness, seizures or death. A feeling of hunger,
sweating, shakiness and weakness may also be present. Symptoms typically come on
quickly.
55
The most common cause of hypoglycemia is medications used to treat diabetes mellitus
such as insulin and sulfonylureas. Risk is greater in diabetics who have eaten less than
usual, exercised more than usual or have drunk alcohol. Other causes of hypoglycemia
include kidney failure, certain tumors, such as insulinoma, liver disease,
hypothyroidism, starvation, inborn error of metabolism, severe infections, reactive
hypoglycemia and a number of drugs including alcohol. Low blood sugar may occur in
otherwise healthy babies who have not eaten for a few hours.

The glucose level that defines hypoglycemia is variable. In people with diabetes, levels
below 3.9 mmol/L (70 mg/dL) is diagnostic. In adults without diabetes, symptoms
related to low blood sugar, low blood sugar at the time of symptoms and improvement
when blood sugar is restored to normal confirm the diagnosis. Otherwise, a level below
2.8 mmol/L (50 mg/dL) after not eating or following exercise may be used. In newborns,
a level below 2.2 mmol/L (40 mg/dL), or less than 3.3 mmol/L (60 mg/dL) if symptoms
are present, indicates hypoglycemia. Other tests that may be useful in determining the
cause include insulin and C peptide levels in the blood.

Hypoglycemia, including iatrogenic hypoglycemia in people with diabetes, causes brain


fuel deprivation that initially triggers a series of physiological and behavioral defenses
but if unchecked results in functional brain failure that is typically corrected after the
plasma glucose concentration is raised. Hypoglycemia is a distinctly uncommon clinical
event except in people who use drugs that lower the plasma glucose concentration.
Hypoglycemia causes recurrent morbidity in most people with type 1 diabetes mellitus
(T1DM) and in many with advanced type 2 diabetes mellitus (T2DM) and is sometimes
fatal.

Recent interest in alternative brain fuels (including lactate) derived from glucose largely
within the brain notwithstanding, glucose is an obligate metabolic fuel for the brain
under physiological conditions. Because the brain cannot synthesize glucose or store
substantial amounts as glycogen in astrocytes, the brain requires a virtually continuous
supply of glucose from the circulation. Facilitated diffusion of glucose from the blood
into the brain is a direct function of arterial plasma glucose concentration. The rate of
blood-to-brain glucose transport exceeds the rate of brain glucose metabolism at normal
(or elevated) plasma glucose levels, but it falls and becomes limiting to brain glucose
metabolism when arterial glucose concentrations fall to low levels. Thus, hypoglycemia
causes brain fuel deprivation and, as a result, functional brain failure. Initially, declining
plasma glucose levels activate defenses against hypoglycemia. Physiological defenses
normally include decrements in pancreatic β cell insulin secretion as glucose levels
decline within the physiological post-absorptive plasma glucose concentration range
(approximately 3.9-6.1mmol/L [70-110mg/dL]). The glycemic threshold for decreased
insulin secretion is approximately 4.5mmol/L (81mg/dL). Increments in pancreatic β
cell glucagon and adrenomedullary epinephrine secretion among other neuroendocrine
responses normally occur as glucose levels fall just below the physiological range
[threshold equal to approximately 3.8mmol/L (8mg/dL)]. If these defenses fail to abort
the hypoglycemic episode, lower glucose levels trigger a more intense sympathoadrenal
response that causes neurogenic (or autonomic) symptoms; neuroglycopenic symptoms
occur at about the same glucose level (threshold equal to approximately 3.0mmol/L
56
(54mg/dL). The perception of symptoms, particularly neurogenic symptoms, prompts
the behavioral defense, the ingestion of food. The plasma if all of these defenses fail,
lower glucose levels cause overt functional brain failure that can progress from
measurable cognitive impairments (threshold equal to approximately 2.8mmol/L
[50mg/dL]) to aberrant behaviors, seizure, and coma. Coma occur at glucose levels in
the range of 2.3-2.7mmol/L (41-49mg/dL) as well as lower glucose levels. All of these
responses are typically corrected after the plasma glucose concentration is raised.
Episodes of hypoglycemia are a fact of life for most people with type 1 diabetes mellitus
(T1DM) and many with advanced type2 diabetes mellitus (T2DM). In type 1 diabetes
mellitus (T1DM), plasma glucose concentrations may be less than 2.8mmol/L
(50mg/dL) as much as 10% of the time; the average patient suffer two episodes of
symptomatic hypoglycemia per week and one episode of severe temporarily disabling
hypoglycemia per year. Although from the iatrogenic deaths do result adverse effects of
drug therapy (the mechanisms are unclear but could include cardiac arrhythmias),
seemingly complete recovery from hypoglycemia-induced functional brain failure after
the plasma glucose concentration is raised is the rule. Permanent neurological damage is
rare. Profound prolonged hypoglycemia can cause brain death. Plasma glucose
concentrations of less than 1.0mmol/L (18mg/dL) occur occasionally in people with
diabetes, and dying brain cells presumably neurons, have been reported following
episodes of hypoglycemia at plasma glucose levels of 1.7-1.9mmol/L (30-35mg/dL) but
not following episodes of hypoglycemia at plasma glucose levels of 2.5mmol/L
(45mg/dL)-in rats. Complete recovery follows the vast majority of episodes of clinical
hypoglycemia.

Figure (28): Hypoglycemia in type1 diabetes mellitus (www.google.com)


57
5.1Hypoglycemia and Brain Energy Metabolism
Human brain glucose metabolism changes under hypoglycemic conditions.
Hypoglycemia may influence the cognitive and memory processes that are associated
with changes in the energy metabolism of the brain. Despite the high sensitivity of
neurons to hypoglycemic states, certain neurochemical changes allow the cells to
become resistant to low concentrations of glucose.

Due to slow glycogen metabolism at rest and a rapid mobilization during energy crisis or
hypoglycemia, glycogen is considered to be the main emergency energy substrate. Its
physiological role also includes effective support of brain activity when glucose cannot
meet the high energy demands. Lactate derived from glycogen metabolism is able to
maintain neuronal function in the absence of glucose. Consciousness disorders
associated with hypoglycemia are characterized by an inability to detect hypoglycemia,
which may result from the limited availability of alternative energy substrates and
reduced activity of glucose itself. If the concentration of intracellular glucose stores (i.e.,
glycogen) increases during hypoglycemia, this might contribute to the masking of an
increased demand by other cells for glucose, thereby causing hypoglycemia-related
disturbances in consciousness. This condition is associated with hypoglycemic
autonomic failure and is very dangerous as there are no symptoms of hypoglycemia
before the onset of cognitive impairment, which in turn can lead to sudden hypoglycemia
and a coma. Glycogen super compensation i.e., an increase in glycogen content after
hypoglycemia, provides evidence of the mobilization of brain glycogen during
hypoglycemia.

Normal cerebral glucose metabolism during hypoglycemia may be maintained by


several potential mechanisms. One possible explanation is a compensatory increase in
the cerebral uptake of lactate. As plasma lactate increases by about 50% in response to
hypoglycemia, it may be used by the brain as an alternative source of energy and reduce
need for glucose in the brain. When converted to pyruvate, lactate carbons may enter the
tricarboxylic acid (TCA) cycle, similar to glucose carbons. Mason et al. (2006) indicated
that during hypoglycemia the levels of transporters for monocarboxylic substrates such
as lactate can increase two times, which is consistent with the proposition that lactate
consumption increases during hypoglycemia. Moreover, even under resting conditions,
increased lactate availability stimulates its consumption by the brain while glucose
utilization decreases. A few recent in vivo studies have shown the neuroprotective effect
of lactate during hypoglycemia or cerebral ischemia episodes, the brain preferentially
utilized lactate over glucose, and so was able to sustain the activity of neurons for hours,
even without glucose. According to some papers, neurons are the preferential site of
lactate oxidation, while lactate itself is produced by astrocytes. Although the lactate in
the brain was long considered a sign of cerebral harm and hypoxia, lactate is now
postulated to have physiological functions in the central nervous system (CNS). It seems
to be an important element in the metabolic cooperation between neurons and glia cells,
especially during increased brain activity. In addition, in order to coordinate brain
activity and energy supply, lactate levels are detected by a specific type of neurons
(orexin neurons in the lateral hypothalamus).

Glucose is used by neurons to maintain their antioxidant status via the pentose phosphate

58
pathway (PPP) that cannot be fueled by lactate. In the case of low plasma glucose levels
and low concentration gradient, glucose transport into neurons is not sufficient to
stimulate their antioxidant pentose phosphate pathway (PPP), and some neurons,
especially vulnerable to reactive oxygen and nitrogen species, may not completely avoid
oxidative damage. Glucose is also needed by the astrocyte to pump glutamate and as
such plays a significant role in functional neuroenergetics. Lactate is transported by
monocarboxylate transporters (MCTs) in a co-transport with protons. An increased
lactate concentration may bring about changes in lactate influx and then in intracellular
and extracellular pH in neurons. These changes in the proton gradient could interfere
with nerve condition and lead to a delayed response. The extracellular lactate
concentration seems to increase during stimulation, and there is also some evidence for
translocation of monocarboxylate transporter-2 (MCT2) to the membrane surface
during stimulation. This process would increase the transport of lactate and result in a
higher intracellular lactate concentration. The increase in the conversion of lactate to
pyruvate is enhanced by an increased lactate/pyruvate ratio, the lower pyruvate level
results from a decreased glycolytic flux. Neurons must rely on lactate as an energy
substrate, especially under hypoglycemic conditions. Even in normoglycemia, it was
observed an increased turnover of lactate during increased activation.

In conclusion, lactate is capable of maintaining neuronal integrity in glucose deficiency.


Furthermore, lactate is preferred over glucose if both these substances are available. The
healthy human brain seems to be able to withstand moderate reductions in plasma
glucose.

Figure (29): Lactate and hypoglycemia (www.google.com)


59
5.2 Hypoglycemia and Cerebral Blood Flow
Cerebral blood flow (CBF) rises during hypoglycemia have been reported in several
studies. The increase in cerebral blood flow (CBF) during hypoglycemia can be
considered neuroprotective because it represents an attempt to increase capillary
glucose concentration for improved glucose supply to the brain when circulating glucose
levels are below a critical level.

6.Brain Disorders and Diabetes Mellitus

6.1 Alzheimer's Disease


Alzheimer's disease (AD) is age related condition characterized by increased incidence
and prevalence with aging. Alzheimer's disease (AD) is the most common form of
dementia among elderly population. Alzheimer's Disease (AD) is characterized by
progressive loss of memory and cognition. Alzheimer's disease (AD) can only be
diagnosed with certainty postmortem demonstration of abundant neurofibrillary tangles
and neuritic plaques with accompanying accumulation of amyloid precursor protein,
amyloid-β (APP-Aβ) deposits in plaques and vessel walls in selected regions of the
brain. Dementia-associated structural lesions are caused by neuronal cytoskeletal
collapse and accumulation of hyperphosphorylated and polyubiquitinated microtubule-
associated proteins, such as tau, resulting in the formation of neurofibrillary tangles,
dystrophic neuritis, and neuropil threads. These features are accompanied by
mitochondrial dysfunction and alterations in neuronal synapses. Progressive loss of
fibers and cells and disconnection of synaptic circuitry mediate the cerebral atrophy that
worsens over time.

Figure(30): Alzheimer's disease (www.google.com)

A close relationship of Alzheimer's disease (AD) is with diabetes mellitus. Persons with
diabetes have been reported to hold a higher incidence of cognitive decline and
Alzheimer's disease (AD); Type2 diabetes mellitus (T2DM) has been strongly
associated with an increased risk of developing all types of dementia, including
Azheimer's disease (AD). In a longitudinal cohort study, lasting up to 9 years, the risk of
developing Alzheimer's disease (AD) was 65% higher in persons with diabetes than in
non-diabetic control. Several studies have suggested that longer diabetes duration is
60
generally associated with a higher risk for developing dementia. Cognitive deficits in
Type2 diabetes mellitus (T2DM) mainly affected the areas of psychomotor efficiency,
attention, learning and memory, mental flexibility, and speed and executive function.
Insulin has been reported to activate signaling pathways associated with long term
memory and learning, and regulate plasticity, energy metabolism and neuronal survival
required for memory and learning. Adversely, insulin resistance or lack of insulin have
been reported to contribute towards the process of degeneration in brain. A series of
longitudinal studies has shown that glucose intolerance and impairment of insulin
secretion are associated with a higher risk to develop dementia or Alzheimer's disease
(AD).

The increased risk of dementia in type2 diabetes mellitus (T2DM) could be linked to
chronic hyperglycemia, peripheral insulin resistance, oxidative stress, accumulation of
advanced glycation end products, increased production of pro-inflammatory cytokines,
and/or cerebral microvascular disease. Type2 diabetes mellitus (T2DM) and
Alzheimer's disease (AD) are age-related conditions, both characterized by increased
incidence and prevalence with aging. A magnetic resonance imaging study
demonstrated that older adults with type2 diabetes mellitus (T2DM) have a moderately
increased risk for developing lacunes and hippocampal atrophy and that the severity of
those lesions increases with the duration and progression of type2 diabetes mellitus
(T2DM). Persons with diabetes have increased incidence of cognitive decline and
Alzheimer's disease (AD). The relationship of diabetes mellitus (DM) and Alzheimer's
disease (AD) was also suggested by the fact that insulin receptor in the brain is
distributed abundantly in synaptic membranes of hippocampus and cerebral cortex. The
neuronal insulin receptor does not seem to have a role in glucose metabolism but seem to
be involved in more diverse brain functions, including synaptic activities required for
learning and memory. Insulin receptors by binding to insulin have been reported to
impact cognition, as they are mostly located in hypothalamus, olfactory bulb,
cerebellum, cerebral cortex and hippocampus; thus making it logical to explore various
aspects of association between cognition and insulin. A massive reduction in the
expression of insulin receptor and insulin level results in deficiency of downstream
signaling pathways in the brain of Alzheimer's disease (AD) patients. Reduced
utilization of glucose by brain is a crucial factor in Alzheimer's disease (AD)
development.

Recent studies confirmed the role of insulin as possible link between type2 diabetes
mellitus (T2DM) and Alzheimer's disease (AD). Low insulin level and insulin resistance
provide the characteristics of Alzheimer's disease (AD) in the central nervous system
(CNS) because it is produced not just in the pancreas but also in the brain. Altered insulin
signaling may contribute to Alzheimer's disease (AD) biochemical and
histopathological lesions. Insulin resistance has been strongly implicated as a possible
link between type2 diabetes mellitus (T2DM) and Alzheimer's disease (AD).A condition
of hyperinsulinemia, regardless of the presence of type2 diabetes mellitus (T2DM),
appears to be associated with a worse cognitive performance. A state of chronic
hyperinsulinemia, as it occurs in insulin-resistance conditions and in type2 diabetes
mellitus (T2DM) may determine a down-regulation of the insulin receptors at the blood-
brain barrier (BBB), thus reducing the transport of insulin in the brain. Hyperglycemia
61
and hyperinsulinemia may accelerate brain aging by inducing tau hyperphosphorylation
and amyloid oligomerization, as well as by leading to widespread brain
microangiopathy. Persons with diabetes are more prone to develop accelerated
leukoaraiosis (white matter high-intensity lesions). Insulin resistance can lead to
mitochondrial dysfunction and increased oxidative stress, thus resulting in a reduction in
synaptic plasticity in the hippocampus. The aging process and the use of fat rich diet is a
factor exacerbating oxidative stress and formation of lipid peroxidation products.

Figure(31): The role of Type2 diabetes mellitus in Alzheimer's disease


(www.google.com)

De la Monte et al. (2008) first used the term type 3 diabetes mellitus (T3DM) for sporadic
Alzheimer's disease (AD), because the pathological features of sporadic Alzheimer's
disease (AD) harbors of both type1 diabetes mellitus (T1DM) and type2 diabetes
mellitus (T2DM), characterized by reduced insulin production and resistance to insulin
receptors. Alzheimer's disease (AD) is typically characterized by a reduced glucose
utilization. Alzheimer's disease (AD) patients have been reported to have fewer insulin
receptors and lesser insulin than the normal individuals. Type3 diabetes mellitus
(T3DM) corresponds to a chronic insulin resistance plus insulin deficiency state that is
largely confined to the brain but, can overlap with type2 diabetes mellitus (T2DM. Type3
diabetes mellitus (T3DM) represents a major pathogenic mechanism of Alzheimer's
disease (AD) neurodegeneration. Report by Talbot and Wang (2014), the concept of
62
type3 diabetes mellitus is no longer correct for a couple of reasons including the fact that:

1- hyperglycemia- which is the key diagnostic features of diabetes-is not present in


acetoacetate decarboxylase-like superfamily (ADCSF).
2- It is not clearly demonstrated that Alzheimer's disease (AD) brain insulin-deficient.

6.1.1Amyloid-Beta Protein and Alzheimer's Disease


Amyloid beta (Aβ or A beta) denotes peptides of 36–43 amino acids that are crucially
involved in Alzheimer's disease as the main component of the amyloid plaques found in
the brains of Alzheimer patients. The peptides derive from the amyloid precursor protein
(APP), which is cleaved by beta secretase and gamma secretase to yield amyloid beta
(Aβ). Amyloid beta (Aβ) molecules can aggregate to form flexible soluble oligomers
which may exist in several forms.

Most studies have suggested that the deposit of the toxic amyloid-beta peptide caused by
an abnormal processing of amyloid-beta precursor protein my initiate and/or contribute
to the pathogenesis of Alzheimer's disease (AD).

Evidence is growing to link an alteration of metabolism and the deposition of precursors


of amyloid in the brain that may occur in persons with diabetes, which is suggested as the
pathogenesis of Alzheimer's disease (AD) in type2 diabetes mellitus (T2DM). The
amyloid precursor protein (APP) is a transmembrane protein consisting of 770 amino
acids; it is known to be the precursor of the amyloid beta involved in the etiopathogenesis
of Alzheimer's disease (AD). Whereas insulin is a neurothrophic factor at moderate
concentrations, hyperinsulinemia with elevated concentrations of insulin in the brain
may be associated with reduced amyloid-beta (Aβ) clearance due to competition for
their common and main degrading mechanisms-the "Insulin-Degrading Enzyme"
(IDE). Insulin modulates metabolism of amyloid precursor protein decreasing
intracellular accumulation. Insulin is degraded by the Insulin-Degrading Enzyme (IDE),
which is also involved in the metabolism and degradation of amyloid beta (Aβ). This
multifunctional enzyme degrades insulin and amylin, peptides related to the pathology
of type2 diabetes mellitus (T2DM), together with amyloid-beta peptide in the brain.
Hyperinsulinemia may elevate amyloid beta (Aβ) through insulin's competition with
amyloid beta (Aβ) for Insulin-Degrading Enzyme (IDE). Therefore, it has been
suggested that the link between hyperinsulinemia and Alzheimer's disease (AD) may be
Insulin-Degrading Enzyme (IDE). Since Insulin-Degrading Enzyme (IDE) is much
more selective for insulin than for amyloid beta (Aβ), brain hyperinsulinemia may
deprive amyloid beta (Aβ) of its main clearance mechanism, favoring its accumulation
in the brain, and its consequent neurotoxic effects.

Amyloid beta (Aβ) is associated with the formation of reactive oxygen species (ROS)
and reactive nitrogen species (RNS), and induced calcium-dependent excitotoxicity,
impairment of cellular respiration, and alteration of synaptic functions associated with
learning and memory.

63
Figure (32): Impact of amyloid beta on memory (www.google.com)

6.1.2 Oxidative Stress and Alzheimer's Disease


Reactive oxygen species/ reactive nitrogen species (ROS/RNS) play a dual biological
role in living systems since they can be either beneficial or detrimental. Low levels of
reactive oxygen species (ROS) exert beneficial physiological roles in cellular responses
to stress and in the activation of several cellular signaling pathways, such as synaptic
signaling, where reactive oxygen species (ROS) are messenger molecules in long-term
potentiation (LTP). Moreover, moderate reactive oxygen species (ROS) levels are
thought to improve peripheral insulin sensitivity. On the other hand, imbalance between
the production of reactive oxygen species/ reactive nitrogen species (ROS/RNS) as a
consequence of mitochondrial dysfunction and the intracellular antioxidant capacity
leads to abnormally elevated reactive oxygen species (ROS) levels and a condition
known as oxidative stress (OS) that is followed by oxidative damage to cells and,
eventually death. Increased oxidative stress (OS) has been implicated in the pathology of
several diseases and the accumulation of oxidatively damaged proteins, lipids, and
nucleic acids correlates with the onset of age-related cellular alterations, especially in
diabetes and Alzheimer's disease (AD). Oxidative damage to lipids and protein of
neuronal membrane affects activities of membrane-bound enzymes, ion channels and
receptors. Reactive oxygen species/ reactive nitrogen species (ROS/RNS) attack
biological component of the cell resulting in the formation of stable adducts with
deoxyribonucleic acid (DNA) and ribonucleic acid (RNA) and can damage cell or
organelle membranes directly (e.g., through lipid peroxidation), or reacting with metals,
nitrogen or carbon to form intermediates that react with proteins (e.g., through nitration,
carbonylation, nitrosylation or reactive alkenals by Michael addition). Oxidation of
amino acids can lead to the formation of advanced glycation end products (AGEs),
advanced oxidation protein products (AOPPs), peroxides and carbonyls that can attack
other molecules and generate radicals resulting in protein unfolding and in rendering the
protein inactive and prone to aggregation.

Aging-related alterations of mitochondrial function represent the driving force of


increased oxidative stress (OS) in type2 diabetes mellitus (T2DM) contributing to the
progression and development of Alzheimer's disease (AD) pathology.

Patients with Alzheimer's disease (AD) pathology demonstrated the concomitant

64
presence of mitochondrial bioenergetics failure, increased oxidative stress(OS) and
reduced insulin signaling as consequence of increased amyloid beta (Aβ) and levels of
hyperphosphorylated tau.

Alzheimer's disease (AD) is considered a brain form of diabetes, with features of both
insulin resistance and insulin deficiency. The progressive worsening of insulin
resistance along stages of Alzheimer's disease (AD) correlates with the increased
oxidative stress (OS), deoxyribonucleic acid (DNA) damage and protein oxidation
demonstrated by 4-hydroxynonenal (HNE), protein carbonyls (PCO), and 3-
nitrotyrosine (3NT) accumulation. It was proposed that insulin resistance together with
decreased brain insulin levels might lead to accumulation of amyloid beta (Aβ) and
consequently Alzheimer's disease (AD).

Figure (33): Type2 diabetes mellitus develops Alzheimer's disease


(www.google.com)

6.1.3 Advanced Glycation End Products and Alzheimer's Disease


The association between advanced glycation end products (AGEs) and vascular
dementia is well established for long time but, the relation between advanced glycation
end products (AGEs) and Alzheimer's disease remains uncertain. Recently, several
reports suggested that diabetes mellitus as well as Alzheimer's disease (AD) are
associated with enhanced oxidative stress and intense glycation end products (AGEs)
production. Advanced glycation end products (AGEs) have been reported to be present
in central nervous system (CNS), retinal vessels, kidneys, and peripheral nerves of
diabetic mellitus (DM) patients, and exhibited that they associate with free radicals to
promote oxidative damage and cause injury to cells. Oxidative stress itself also results in
the formation of advanced glycation end products (AGEs), thus creating a vicious cycle.
Advanced glycation end products (AGEs) have been reported to present in and modify
65
neurofibrillary tangles (NFTs) and senile plaques. Since neurofibrillary tangles (NFTs)
as well as plaques are implicated in Alzheimer's disease (AD), so the patients suffering
from diabetes can have an enhanced Alzheimer's disease (AD) risk via advanced
glycation end products (AGEs) production. Since type2 diabetes mellitus enhances
advanced glycation end products (AGEs) production, it may act as another causative
factor for Alzheimer's disease (AD) development. The measurement of toxic advanced
glycation end products (AGEs) in cerebrospinal fluid and serum have been proposed to
be a potential biomarker for early Alzheimer's disease (AD) detection.

Diabetic patients could have an increased risk of Alzheimer's disease (AD) via advanced
glycation end products (AGEs) production since the modification of amyloid beta (Aβ)
by advanced glycation end products (AGEs) accelerates amyloid beta (Aβ) aggregation
and the glycation of tau stabilizes neurofibrillary tangles. High levels of advanced
glycation end products (AGEs) immunoreactivity are present in Alzheimer's disease
(AD) plaques and neurofibrillary tangles. Receptor of advanced glycation end products
(RAGE) has been found to be a receptor for amyloid beta (Aβ) and mediates amyloid
beta (Aβ) induced microglia activation and subsequent inflammation in Alzheimer's
disease (AD).

Figure (34): Advanced glycation end products result in Alzheimer's disease


(www.google.com)

6.2 Parkinson's Disease


Parkinson's disease (PD) is the second most common neurodegenerative disease
afflicting about 1% of people over 65 years old and 4-5% of people over 85 years.
Typically, Parkinson's disease (PD) is the result of the degeneration of neurons in the
substantia nigra pars compacta (SNpc), which leads to the subsequent reduction of
dopaminergic input to the striatum. Moreover, there is a degeneration of neurons of
selected brainstem nuclei (locus coeruleus, raphe nuclei, dorsal motor nucleus of the
66
vagus), cortical neurons (particularly within the cingulated gyrus and the entorhinal
cortex), the nucleus basalis of Meynert and preganglionic sympathetic and
parasympathetic neurons. In the soma of these neurons, the existence of intracellular
proteinaceous inclusions, called Lewy bodies and Lewy neuritis, mainly composed of α-
synuclein, have been observed. The characteristic distribution of these aggregations is
considered to be the most classical neuropathological hallmark of Parkinson's disease
(PD).

Several reports discuss that the mechanism of neuronal death in Parkinson's disease (PD)
starts with an otherwise healthy dopaminergic neuron being hit by an etiological factor,
such as mutant α-synuclein. Besides, type2 diabetes mellitus, chronic renal failure, past
brain insults, or genetically determined differences in drug metabolism were also
suggested as a risk factor for Parkinson's disease (PD). Also, the coexistence of
dopaminergic neurons and insulin receptors in the substantia nigra pars compacta
(SNpc) reinforce the occurrence of a direct association between the two diseases. There
are various ways in which a shared pathogenesis of diabetes, dementia, and Parkinson's
disease (PD) may occur. One is that there might be an underlying disorder of
mitochondrial bioenergetics, manifest in pancreatic beta-cells and adipose tissue; this
might be attributable to limited activation of peroxisome proliferator-activated receptor-
γ (PPAR-γ), PPAR coactivator-1α (PGC1α) and its link to adenosine monophosphate
(AMP) kinase in the substantia nigra pars compacta (SNpc) and dopaminergic neurons.
Another overlapping cytotoxic disorder is that of abnormal protein folding, which is
associated with amylin-derivative effects on pancreatic beta-cells in diabetes, the
neurodegenerative taupathies (hyperphosphorylation of tau, low levels of soluble tau),
the formation of amyloid precursor protein and with synucleinopathies in
neurodegenerative disorders characterized by neurofibrillary aggregates of α-synuclein
protein in neurons and glial cells in parkinson's disease (PD).

Neuropathological studies of patients with Parkinson's disease have shown that insulin
receptors are densely represented on the dopaminergic neurons of the substantia nigra
pars compacta (SNpc) and loss of insulin receptor immunoreactivity and messenger
ribonucleic acid (mRNA) in the substantia nigra pars compacta (SNpc) of patients with
Parkinson's disease coincides with loss of tyrosine hydroxylase messenger ribonucleic
acid (mRNA) (the rate-limiting enzyme in dopamine synthesis). Indeed abnormal
glucose utilization has been specifically shown in the brains of patients with Parkinson's
disease undergoing either magnetic resonance spectroscopy or fluorodeoxyglucose-
PET demonstrating increased lactate concentrations and glucose hypometabolism,
supporting the hypothesis that Parkinson's disease is a systemic disorder characterized
by a derangement of oxidative energy metabolism. Further evidence for a positive
association between Parkinson's disease and type 2 diabetes mellitus (T2DM) has been
obtained from prospective cohort studies. A statistically significant direct association
between triceps skinfold thickness and the risk of Parkinson's disease has been found in
the Honolulu Heart Program and both excess weight and type 2 diabetes mellitus
(T2DM) itself were associated with an increased risk of Parkinson's disease in a
population-based prospective cohort of Finnish males and females. This association was
independent of the known modifying factors such as smoking status, coffee and alcohol
consumption and body weight, tempting speculation regarding common pathways
67
underlying the development of these conditions. Nevertheless, a recent cohort study
could not replicate an association between either type 2 diabetes mellitus (T2DM) or
obesity and Parkinson's disease risk although the authors acknowledge that diagnosis of
type 2 diabetes mellitus (T2DM) was entirely based on self-report.

Any link between Parkinson's disease and type 2 diabetes mellitus (T2DM) may relate to
abnormalities in a common pathway or indirectly as a consequence of chronic
hyperglycaemia or hyperinsulinaemia. Animal and in vitro studies have suggested a role
for insulin in the regulation of brain dopaminergic activity featuring a reciprocal
regulation between the two chemicals. In the rat, elevation of glucose concentrations in
the blood, to levels equivalent to those produced by a meal or stress, suppresses the firing
of dopamine-containing neurons located within the substantia nigra and prevents or
reverses the increase in discharge rates of dopaminergic cells normally elicited by the
dopamine receptor antagonist haloperidol. Administration of glucose to rats has also
been shown to produce a significant decrease of dopamine turnover in both striatum and
olfactory tubercle.

Converging evidence suggests that cellular pathways leading to either insulin resistance
or neurodegeneration involve mitochondrial mechanisms. It is well known that
mutations in mitochondrial deoxyribonucleic acid (DNA) can lead to a wide variety of
phenotypes that commonly involve neurodegeneration and diabetes. However, these
mutations do not account for a significant proportion of cases of sporadic Parkinson's
disease. Normal mitochondrial biogenesis, respiration and metabolism of reactive
oxygen species (ROS) requires intact expression of both nuclear and mitochondrial
encoded genomes now recognized as being regulated by Peroxisome proliferator-
activated receptor gamma coactivator 1-alpha (PGC1). It has previously been shown
that Peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC1) has
a powerful suppressive effect on reactive oxygen species production, in parallel to its
effects in elevating mitochondrial respiration. This occurs through the PGC1-mediated
expression of genes involved in reactive oxygen species detoxification, as well as
peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC1)
expression that is rapidly induced by these proteins following a single bout of endurance
exercise in vivo. Insulin resistant patients show reduced expression of Peroxisome
proliferator-activated receptor gamma coactivator 1-alpha (PGC1) and the
mitochondrial encoded gene Cyclooxygenase-1 (COX1), while reduction in PGC1-
responsive genes has been shown among patients with type 2 diabetes mellitus (T2DM)
and their asymptomatic relatives compared with healthy controls. Indeed
polymorphisms in peroxisome proliferator-activated receptor gamma coactivator 1-
alpha (PGC1) have been associated with an increased risk for type 2 diabetes mellitus
(T2DM) in diverse populations. Peroxisome proliferator-activated receptor gamma
coactivator 1-alpha (PGC1) has also been implicated as having a major role in
Parkinson's disease pathogenesis. Meta-analysis of gene expression data using
microarrays has utilized post-mortem brain homogenates to look at gene expression in
the substantia nigra pars compacta of patients with confirmed SNCA-positive Lewy
body Parkinson's disease. Robust three tiered analysis from separate genome wide
datasets have indicated that 'gene sets' involved in mitochondrial electron transport,
mitochondrial biogenesis, glucose utilization and glucose sensing were strongly
68
associated with Parkinson's disease. Included amongst these gene sets were 10 PGC1
responsive genes. Furthermore it was shown that over-expression of peroxisome
proliferator-activated receptor gamma coactivator 1-alpha (PGC1) was able to protect
dopamine cell loss induced by the mitochondrial toxin rotenone. In parallel with these
discoveries was the identification of a zinc-finger protein, parkin interacting substrate
(PARIS), which is up-regulated 3-fold in the nigra of patients with not only parkin-
related parkinsonism but also sporadic Parkinson's disease, and is both necessary and
sufficient for the neurodegeneration associated with parkin animal models. The same
group further identified that PARIS represses the expression of Peroxisome proliferator-
activated receptor gamma coactivator 1-alpha (PGC1) and PGC1 target genes playing an
important role in mitochondrial function including Nuclear respiratory factor 1 (NRF1),
and the oxidative phosphorylation regulator ATP5b. The site of interaction between
PARIS and Peroxisome proliferator-activated receptor gamma coactivator 1-alpha
(PGC1) is a sequence that is involved in the regulation of transcripts involved in insulin
responsiveness and energy metabolism. Although other pathways are undoubtedly also
relevant, it is clear that parkin, PARIS, Peroxisome proliferator-activated receptor
gamma coactivator 1-alpha (PGC1) and Nuclear respiratory factor 1 (NRF1) contribute
to the pathogenesis of Parkinson's disease. Loss of expression of Peroxisome
proliferator-activated receptor gamma coactivator 1-alpha (PGC1) controlled genes
may therefore be a key link between abnormal mitochondrial function, abnormal
glucose utilization and Parkinson's disease. Hypermethylation of Peroxisome
proliferator-activated receptor gamma coactivator 1-alpha (PGC1) during life may
follow either genetic or environmental influences that promote accumulation of free
fatty acids, tumor necrosis factor (TNF) and ceramides, which might then lead to
decompensation of mitochondrial bioenergetics and the onset of Parkinson's disease.

Figure (35): Role of diabetes mellitus in Parkinson's disease (www.google.com)

69
6.3 Autism Spectrum Disorders
Autism spectrum disorders (ASD) are a group of developmental disabilities
characterized by deficits in socialization, communication, and repetitive or unusual
behaviors. A substantial upward trend for autism spectrum disorders (ASD) prevalence
has been reported since the 1960s, with a recent estimate of about 113 cases per 10,000
(one in 88) children in the United States [Autism and Developmental Disabilities
Monitoring Network Surveillance Year 2008 Principal Investigators and Centers for
Disease Control and Prevention (2012)]. Although the underlying mechanisms still
remain to be elucidated, autism spectrum disorders (ASD) is regarded as multifactorial,
with genetic and non-genetic risk factors acting together to produce the phenotype.
Inspired by the conceptual framework of the developmental origins of health and disease
(DOHaD) hypothesis and exciting findings on the effects of early intrauterine
environment insults on brain development, increasing research initiatives have
concentrated on the identification of early life determinants for autism spectrum
disorders (ASD) risk.

Diabetes affects up to 15 % of pregnant women worldwide. Approximately 87.5 % of


maternal diabetes are due to gestational diabetes (GDM), defined as glucose intolerance
with onset or first recognition during pregnancy, 7.5 % to pre-existing type 1 diabetes
(T1DM), and 5 % to pre-existing type 2 diabetes (T2DM). The prevalence of maternal
diabetes is increasing. Maternal diabetes has been associated with an increased risk of
miscarriage, macrosomia and other adverse fetal outcomes, as well as poor
neurodevelopmental outcomes in the offspring. Furthermore, a positive association
between maternal diabetes and autism spectrum disorders (ASD) risk in the offspring
was recently reported in a case–control study and two subsequent cohort studies.
However, the results from other studies have been conflicting and inconclusive.

In utero exposure to hyperglycemia, a consequence of any types of maternal diabetes,


may increase the risk of offspring autism spectrum disorders (ASD) involving several
plausible biological mechanisms. First, maternal hyperglycemia can result in hypoxia in
the fetus, and a depleted oxygen supply to the fetus may impair neurodevelopment and
thus contribute to a greater risk of autism spectrum disorders (ASD). Second, maternal
hyperglycemia has been associated with increased free-radical production and impaired
antioxidant defense system that lead to oxidative stress in the cord blood and placental
tissue. The positive association between oxidative stress and autism spectrum disorders
(ASD) in children was reported in several studies. Third, excessive adiposity that
commonly accompanies type 2 diabetes mellitus (T2DM) and gestational diabetes are
inducers of chronic inflammation. A number of studies have identified both neuro and
systemic markers of inflammation in children with autism. Also, it has been
hypothesized that insulin signaling may contribute to development of autism in
genetically susceptible individuals via activation of PI3K/Tor pathway in neurons.
Fourth, type1 diabetes mellitus (T1DM) is an autoimmune disorder resulting from a
cellular-mediated autoimmune destruction of pancreatic beta-cells. Maternal altered
autoimmunity may influence brain development in the off-spring either by creating a
hostile intrauterine environment or by modifying the offspring's autoimmunity in early
development through immunoglobulin G. Fifth, epigenetic modification by
hyperglycemia may also be implicated in the pathogenesis of autism spectrum disorders
(ASD), but the current evidence is still sparse.
70
Figure(36): Possible pathophysiology of autism spectrum disorder in preterm
infants (www.google.com)

Figure (37): Impact of maternal diabetes on incidence of autism spectrum


disorders (www.google.com)

6.4 Mood and Psychotic Disorders and Type2 Diabetes Mellitus


Major depressive disorder is defined as recurrent periods of depressive episodes that
interfere with psychosocial, interpersonal or vocational functioning. A depressive
episode includes the presence of depressed mood and/or anhedonia for a duration of ≥2
weeks, with accessory symptoms such as sleep disturbance, diminished energy, poor
concentration, thoughts of hopelessness and suicidal ideation. The lifetime prevalence
of major depressive disorder is estimated to be 10 to 25% in women and 5 to 12% in men.
Bipolar I disorder is defined as the presence of a manic or mixed episode. A manic
episode is an expansive, euphoric and/or irritable mood for a period of ≥1 week (or any
duration if hospitalization is required), associated with accessory symptoms such as
decreased need for sleep, racing thoughts as evidenced by rapid speech, impulsive
behavior, profligate spending and increased overall activity. A mixed episode is the
simultaneous presence of a manic and depressive episode. Bipolar I disorder is common,
affecting approximately 1 to 2% of the general population. Schizophrenia, a prototypical
71
psychotic disorder, is defined as a minimum 6-month disturbance during which ≥2 of the
following symptoms—delusions, hallucinations, disorganized speech, disorganized
behavior or negative symptoms—are present for at least 1 month. In addition, symptoms
are associated with impairment of the major spheres of functioning. Schizophrenia
affects an estimated 1% of the general population. Contrary to major depressive
disorder, bipolar I disorder and schizophrenia appear to affect both genders equally.
Similar to type 2 diabetes, mood and psychotic disorders are also associated with
obesity, sedentary lifestyle and genetic etiology. The etiology of mood and psychotic
disorders is unknown, but is likely multifactorial, with genetics accounting for an
estimated 70% of the vulnerability for bipolar I disorder and schizophrenia and 30 to
40% of that for major depressive disorder. Collectively, mood and psychotic disorders
are associated with significant psychosocial impairment, human suffering and economic
costs. Results of a recent investigation undertaken by the World Health Organization
indicate that major depressive disorder, bipolar disorder and schizophrenia are each
ranked among the top 10 leading causes of disability and premature death among
persons age 15 to 44, globally. Unnatural death (e.g. suicide) accounts for a significant
proportion of this increased risk; however, mortality from natural causes [e.g.
cardiovascular disease (CVD), diabetes] accounts for the majority of excess and
premature deaths in mood and psychotic disorders. Adverse effects of type 2 diabetes on
psychiatric status and psychological well-being have been reported; obesity, a well-
known risk factor for type 2 diabetes, has been found to be associated with worse self-
reported mental health ratings. Most investigations examining the prevalence of
psychiatric disorders in populations with diabetes have focused on depression. Recently,
studies evaluated the prevalence of depression in adults with diabetes. The odds of
depression in the group with diabetes was twice that of the non-diabetic comparison
group and did not differ by sex, type of diabetes or assessment method. Results from
well-characterized investigations have sharpened the cross sectional and longitudinal
associations between mood and psychotic disorders and type 2 diabetes. Extant study
results indicate that the prevalence of glucose metabolism disturbances in populations
with mood and psychotic disorders is approximately 2-to 4-fold greater than the age-
adjusted general population estimate. Patients with mood and psychotic disorders have
cluster risk factors (e.g. smoking, obesity) for metabolic disorders, which likely
contribute to this increased risk.

6.4.1 Defects in Bioenergetics Coupling in Schizophrenia


Synaptic neurotransmission relies on maintenance of the synapse and meeting the
energy demands of neurons. Defects in excitatory and inhibitory synapses have been
implicated in schizophrenia, likely contributing to positive and negative symptoms as
well as impaired cognition. Recently, accumulating evidence has suggested that
bioenergetic systems, important in both synaptic function and cognition, are abnormal in
psychiatric illnesses such as schizophrenia. Animal models of synaptic dysfunction
demonstrated endophenotypes of schizophrenia as well as bioenergetic abnormalities.

72
Figure (38): Defects in bioenergetics coupling in schizophrenia
(www.google.com)

6.5 Effects of Diabetes on Cognitive Function and Brain Structure


Diabetes mellitus is associated with decrements in cognitive function and changes in
brain structure. People with type1 and type 2 diabetes have been shown to have mild-to-
moderate reductions in cognitive function. Type 2 diabetes (T2DM) has been associated
with 50% increased risk of dementia. Researchers and clinicians recognized that people
with diabetes complained of poor memory and attention.

6.5.1 Cognitive Dysfunction in Type 1 Diabetes Mellitus


Type 1 diabetes mellitus (T1DM) is commonly diagnosed during childhood and
adolescence. This is a period of rapid developmental changes in the central nervous
system and there has been concern that the younger brain may be more susceptible to
extremes of glycemia. Subjects with type 1 diabetes mellitus (T1DM) have performance
deficits in multiple cognitive domains, including information-processing speed,
psychomotor efficiency, memory, attention, visuospatial abilities, and executive
function. Children with type 1 diabetes mellitus (T1DM) who experience severe
hypoglycemic episodes before the age of 5 have deficits in spatial intelligence and
delayed recall, suggesting that the developing brain at very younger ages may be
susceptible to the effects of hypoglycemia. Other factors, including poor glycemic
control and presence of microvascular complication like neuropathy and retinopathy,
have been associated with cognitive dysfunction in subjects with type 1 diabetes mellitus
(T1DM). People with type 1 diabetes mellitus (T1DM) have mild-to-moderate declines
in multiple domains, including intelligence, speed of information processing,
psychomotor efficiency, attention, cognitive flexibility, and visual perception. Lowered
cognitive performance in diabetic patients appear to be associated with the presence of
73
microvascular complications but not with the occurrence of severe hypoglycemic
episodes or with poor metabolic control. It is suggested that early age of onset may be an
important variable of cognitive dysfunction in children with type 1 diabetes mellitus
(T1DM).

6.5.2 Cognitive Dysfunction in Type 2 Diabetes Mellitus


It is evaluated the impact of type 2 diabetes mellitus (T2DM) on cognitive function. The
studies have been done mostly in middle-age to older adults. People with type 2 diabetes
mellitus (T2DM) perform less well in the cognitive domains of information-processing
speed, memory, attention, and executive function. Mental flexibility and global
cognitive function have been shown to be affected in some, but not in all studies.
Decrements in cognitive function in subjects with type 2 diabetes mellitus (T2DM) have
been associated with increased duration of diabetes and poor glycemic control.
Dementia caused by either Alzheimer's disease or vascular disease have been linked to
type 2 diabetes mellitus (T2DM) in longitudinal studies. Cognitive deficits is noted in the
domains of processing speed and executive function and is associated with duration of
diabetes. The presence of type 2 diabetes mellitus (T2DM) almost double the risk of
dementia. Cross-sectional studies have shown that subjects with type 2 diabetes mellitus
(T2DM) performed poorly in several cognitive domains including attention, executive
function, information processing, memory, psychomotor efficiency, verbal fluency, and
learning. These reductions have been associated with poor glycemic control, longer
duration of diabetes, and the presence of microvascular complications such as diabetic
retinopathy and peripheral neuropathy. Epidemiological studies have shown that
comorbidities such as hypertension, dyslipidemia, and depression are associated with
poor cognitive function in subjects with type 2 diabetes mellitus (T2DM). It was
reported that risk of dementia was increased by 50-100% in people with type 2 diabetes
mellitus (T2DM) relative to people without diabetes.

Figure (39): Cognitive disorders in diabetes (www.google.com)

74
Figure (40): Contributors to cognitive dysfunction in diabetes mellitus
(www.google.com)

6.5.3 Impact of Type 1 Diabetes Mellitus on Brain Structure


Studies have shown lower gray and white volumes in subjects with type 1 diabetes
mellitus (T1DM) compared to non-diabetic controls. Patients with diabetes have lower
gray matter density, primarily in the posterior, temporal and cerebellar regions of the
brain. Lower gray matter density is associated with poor glycemic control, higher
frequency of severe hypoglycemic events, age of onset, and duration of diabetes. The
frontal lobe is the location of reduced volumes in patients with type 1 diabetes mellitus
(T1DM). Reduced white matter volumes were identified in subjects with type 1 diabetes
mellitus (T1DM) and this volume loss was associated with lower performance of
attention, speed of information processing, and executive function. Age of onset,
duration of diabetes, and presence of retinopathy have been associated with structural
changes in type 1 diabetes mellitus (T1DM). Patients with type 1 diabetes mellitus
(T1DM) have been shown to have increased white matter lesions (WMLs), which may
represent vascular abnormalities in intraparenchymal cerebral arterioles. In patients
with type 1 diabetes mellitus (T1DM) increased severity of white matter lesions
(WMLs) compared to controls has been reported in some but not all studies. It was found
that hippocampal gray matter volume was larger in children with type 1 diabetes mellitus
(T1DM) with history of three or more severe hypoglycemic episodes in the past. The
changes in brain structure have been associated with decline in cognitive performance. It
was reported that in patients with type 1 diabetes mellitus (T1DM) white matter
microstructural deficits in the posterior corona radiate and the optic radiation correlated
with lower performance in cognitive functions were thought to be associated with white
matter function. Studies on type 1 diabetes mellitus (T1DM) patients with neuropathic
pain with or without microangiopathy revealed abnormalities in networks involving
attention, working memory, auditory and language processing, and motor and visual
75
areas. Reduced functional connectivity in the attention network was found in diabetics
with microangiopathy compared to controls, but not in patients who did not have
microangiopathy. Subclinical macroangiopathy was also found to be a factor that likely
contributed to development of diabetes-related cognitive changes in type 1 diabetes
mellitus (T1DM).

6.5.4 Impact of Type 2 Diabetes Mellitus on Brain Structure


People with type 2 diabetes mellitus (T2DM) have been shown to have brain atrophy,
including lower total and regional white and gray matter volumes compared to non-
diabetic controls. Patients with type 2 diabetes mellitus (T2DM) was reported to have
lower total gray, white, and hippocampal volumes. Regions with loss of gray matter
include the medial temporal, anterior cingulate, and medial frontal lobes. White matter
loss was found in the frontal and temporal regions. These investigators determined that
brain volume loss was associated with poor performance in cognitive performance in
patients with type 2 diabetes mellitus (T2DM). Other studies have suggested that
atrophy may be greater in the hippocampal region in patients with type 2 diabetes
mellitus (T2DM). Brain atrophy and white matter lesions (WMLs) have been associated
with cognitive dysfunction in some but not all studies. Diabetes-related risk factors,
including hypertension, duration of diabetes, glycemic control, and retinopathy, have
been associated with brain structural changes . It was reported microstructural
abnormalities and disruptions in the white matter network in individuals with type 2
diabetes mellitus (T2DM) compared with controls. These abnormalities were related to
slowing of information-processing speed. A study demonstrated that the decline in
cognitive performance in type 2 diabetes mellitus (T2DM) was associated with a
reduction in functional connectivity of the hippocampus. Patients with type 2 diabetes
mellitus (T2DM) have an increased incidence of both Alzheimer's disease and vascular-
type dementia. A study presented a decreased amplitude of low-frequency fluctuations
observed in the postcentral gyrus and occipital lobe was in patients with type 2 diabetes
mellitus (T2DM). This was associated with worse memory performance and executive
functioning. The microvascular complications of diabetes largely contribute to the
development of brain functional abnormalities, which possibly precede the cognitive
decline observed in type 2 diabetes mellitus (T2DM).

7.Hypoglycemia and Brain Function Failure


Hypoglycemia, including iatrogenic hypoglycemia in people with diabetes, causes brain
fuel deprivation that initially triggers a series of physiological and behavioral defenses
but if unchecked results in functional brain failure that is typically corrected after the
plasma glucose concentration is raised. Rarely, profound, prolonged, hypoglycemia
causes brain death.

Profound, prolonged hypoglycemia can cause brain death. In studies of insulin induced
hypoglycemia in monkeys, 5–6 hours of blood glucose concentrations of less than 1.1
mmol/L (20 mg/dL) were required for the regular production of neurological damage;
the average blood glucose level was 0.7 mmol/L (13 mg/dL). Fortunately, hypoglycemia
of that magnitude and duration occurs rarely in people with diabetes. The mechanisms of
the common, hypoglycemia-induced functional brain failure and of the rare,
hypoglycemia-induced brain death that occurs at very low, and at least in primates

76
prolonged, plasma glucose concentrations differ. The former is the result of brain fuel
deprivation per se, but the latter is not. As summarized in a study a variety of mechanisms
are thought to be involved in the pathogenesis of hypoglycemic neuronal death. These
include glutamate release and activation of neuronal glutamate receptors, production of
reactive oxygen species, neuronal zinc release, activation of poly(ADP-ribose)
polymerase, and mitochondrial permeability transition. In a report, was described
additional studies of the mechanisms of hypoglycemia-induced neuronal necrosis.
Based on systematic cell culture and in vivo rodent studies of glucose deprivation
followed by glucose provision, they provide evidence that hypoglycemic superoxide
production and neuronal death were increased by nicotinamide adenine dinucleotide
phosphate (NADPH) oxidase activation during glucose reperfusion. These effects were
reduced by an inhibitor of nicotinamide adenine dinucleotide phosphate (NADPH)
oxidase, deficiency of a subunit of the enzyme, and blockade of nicotinamide adenine
dinucleotide phosphate (NADPH) regeneration, among other findings. Notably,
superoxide formation and neuronal death increased with increasing glucose
concentrations during the post-hypoglycemic reperfusion period. That finding is
generally consistent with earlier findings by investigators. In the in vivo studies, blood
glucose concentrations averaged 0.4mmol/L(7 mg/dL), causing an isoelectric
electroencephalogram (EEG), during hypoglycemia and approximately 7.5 mmol/L
(135 mg/dL) during glucose reperfusion that was documented to cause detrimental
effects. Superoxide production, and presumably neuronal death, occurred as a result of
hypoglycemia, but these occurred to a greater extent with glucose reperfusion, less so
when post-hypoglycemic blood glucose concentrations were raised to the range of
1.0–2.0 mmol/L (18–36 mg/dL) than when they were raised to the range of 5.0–10.0
mmol/L (90–180mg/dL). Studies involving less profound hypoglycemia were not
reported. The distinction between the common hypoglycemia-induced functional brain
failure and the rare hypoglycemia-induced brain death drawn here is admittedly
arbitrary. Plasma glucose concentrations of less than 1.0 mmol/L (18 mg/dL) occur
occasionally in people with diabetes, and dying brain cells, presumably neurons, have
been reported following episodes of hypoglycemia at plasma glucose levels of 1.7–1.9
mmol/L (30–35 mg/dL) — but not following episodes of hypoglycemia at plasma
glucose levels of 2.5 mmol/L (45 mg/dL) — in rats. Thus, it could be reasoned that these
categories are not binary and that there is a continuous spectrum with increasing risk of
neuronal death at progressively lower plasma glucose concentrations. Nonetheless,
seemingly complete recovery follows the vast majority of episodes of clinical
hypoglycemia. The appropriate clinical extrapolation of these data is not entirely clear.
As the authors point out, plasma glucose concentrations must be raised in hypoglycemic
patients. In the common clinical setting of hypoglycemia-induced functional brain
failure, plasma glucose levels should be raised into the physiological range promptly
with the expectation that recovery of brain function will follow. At this point there is no
clear evidence that post-treatment hyperglycemia is detrimental to recovery, but there is
no reason to think it is beneficial in that setting. On the other hand, under-treatment will
delay recovery. In the rare clinical setting of profound, prolonged hypoglycemia, where
the risk of neuronal death is higher, the data suggest that plasma glucose levels should be
raised cautiously with avoidance of hyperglycemia. Nonetheless, it would seem
reasonable to raise the plasma glucose level into the physiological range [e.g., >3.9
mmol/L (70 mg/dL)] promptly.
77
Figure (41): Impact of hypoglycemia on different body systems
(www.google.com)

8.Hypometabolism of Glucose in Anorexia Nervosa


Anorexia nervosa is a complex syndrome where the most prominent symptom is weight
loss. Different clinical characteristics, however, such as disturbances in the perception of
body shape, high levels of motor restlessness, and hyperactivity are associated with
strategies to induce weight loss such as restricting food, vomiting, purging, or abuse
diuretics or appetite suppressants. A number of metabolic and endocrine abnormalities,
mainly at the hypothalamic level, have been described in anorexia nervosa. Concerning
central nervous system (CNS) neurotransmitters, a significant reduction of
cerebrospinal fluid (CSF), metabolites of serotonine and dopamine, or opiate-related
peptides have been observed in anorexia nervosa. These levels do normalize with weight
restoration, however, long-term weight –restored anorectics have elevated
concentrations of cerebrospinal fluid 5-hydroxyindoleacetic acid (CSF 5-HIAA) and a
reduction of cerebrospinal fluid (CSF) norepinephrine levels. These findings might be
correlated to some clinical characteristics such as altered mood, anxiety, or obsessional
thinking observed in anorexia nervosa. Structural brain changes in anorexia nervosa are
characterized by a reversible decrease of brain volume. Ventricular and sulcal
enlargements seem related to malnourishment and are partly reversible after weight
restoration. Neuropsychological evaluation of patients with anorexia nervosa showed
impaired cognitive performance which appeared to be a sign of poor prognosis at follow-
up. In low- weight anorectic patients it was found an absolute global and regional
hypometabolism of glucose which is most important in frontal and parietal cortices. It
has been suggested that acido-ketosis and high plasma levels of ketone bodies could
explain hypometabolism of glucose selectively at the cortical level. Some central
neurotransmitters systems have proved to be altered in anorectic patients with significant
reductions of cerebrospinal fluid (CSF) 5-hydroxyindole-acetic acid, norepinephrine
and 3-methoxy-4-hydroxyphenylglycol (MHPG) or homovanillic acid levels in the
78
underweight anorectics. Abnormalities of these modulatory amine transmitter systems
could contribute to the absolute global hypometabolism but not the relative regional
glucose hypometabolism. Abnormal hypothalamic functioning, characterized by
neuroendocrine dysfunctions such as amenorrhea, was observed in underweight
anorectics. Different psychopathological disturbances such as depression or anxiety
have been demonstrated to influence cerebral glucose metabolism in different ways. A
hypometabolism in the parietal and superior frontal cortices is associated with a relative
hypermetabolism in the caudate nuclei and in the inferior frontal cortex was observed
and it tended to remain abnormal after weight gain particularly in the parietal and inferior
frontal cortices. The relative hypermetabolism in the caudate nuclei which normalized
after weight gain has also been observed. Obsessional features focused on eating
patterns were common in anorexia nervosa and a relative hypermetabolism of the orbito
frontal gyri, corresponding to inferior frontal cortex, and of the head of caudate nuclei
has been described in obsessive compulsive patients. Persistence of these abnormalities
after weight gain might be associated with the persistence of obsession concerning food,
weight, and body shape in weight restored anorectic patients. A trend towards the
persistence of an absolute or relative glucose parietal hypometabolism is observed in
weight-gained anorectics. Cognitive studies showing distortions of body image
associated with the impairment of the awareness of emaciation and particular arithmetic
low performances tend to support a parietal dysfunction in anorexia nervosa.

Figure (42): A model of anorexia nervosa as an attempt to preserve mental


homeostasis (www.google.com)

9.Carbohydrate-Deficient Glycoprotein Syndromes


A group of metabolic disorders characterized by a defect in the modification of
glycoproteins by carbohydrates. Four types of carbohydrate-deficient glycoprotein
syndrome have been recognized, depending on the isoelectric focusing pattern of serum
sialotransferrins. The basic defect in type Ia is usually a deficiency of
phosphomannomutase, and in type II, a deficiency in Golgi localized N-
acetylglucosaminyltransferase II. In the other two types the cause has not been found;
they present as a severe encephalopathy with intractable epilepsy and a different
cathodal shift on serum transferrin isoelectric focusing from type Ia. The commonest
79
variety of the carbohydrate-deficient glycoprotein syndromes is type Ia, and in most
instances a deficiency of phosphomannomutase has been recognized. The inheritance is
autosomal recessive, and the gene responsible for the disorder has been mapped to
chromosome 16p. However, in a study from Canada involving 17 families, one of them
showed no linkage to chromosome 16p; it is of interest that this family was of French-
Canadian origin and the only non-European one, which suggests genetic heterogeneity.
Another phosphomannomutase gene has been located on chromosome 22q. A study
cloned the phosphomannomutase gene on chromosome 22q (PMM1), and identified a
second human gene on chromosome 16p (PMM2) which encodes a protein 66%
identical to PMM1. It was also found 11 different missense mutations in the PMM2 gene
in 16 patients suffering from the carbohydrate-deficient glycoprotein type Ia syndrome.
Missense mutations in this gene have also been reported in Japanese patients, and it is
suggested that the principle defect is a disturbance in the synthesis of the dolichol-
pyrophosphateoligosaccharide precursor. The typical pathology on examination of the
brain is atrophy of the brainstem and cerebellum. In about a third there is also atrophy of
the cortex. There was report on patients with typical type Ia syndromes and radiological
evidence of olivopontocerebellar atrophy. There was also supratentorial atrophy but it
was considered that the variable mental retardation with special impairment of
visuoperceptual skills, eye–hand coordination, visual memory and language may
support the role of the cerebellum and brainstem in the acquisition of these skills. It was
reported the findings in 13 patients with type Ia disease who had passed the age of 15
years. All had presented with the early onset of delayed development, mostly with mild
facial dysmorphic features and some degree of hepatic dysfunction; and one patient had
a pericardial effusion. About half of them had subcutaneous lipodystrophy and comatose
or stroke-like episodes during childhood. After the age of 15 the disease was mainly
characterized by non-progressive ataxia with cerebellar hypoplasia, stable mental
retardation, variable peripheral neuropathies, and strabismus. One-third of them had
generalized seizures, and they all had retinal pigmentary degeneration and thorathic
deformities. None of the females had passed puberty. All had raised serum
concentrations of carbohydrate-deficient transferrin. The authors concluded that this
type of the syndrome is mainly a non-progressive condition compatible with a socially
functioning but dependent life-style. A type Ib disease has been reported with a
deficiency of phosphomannose isomerase due to mutations in the PMI1 gene. This
phenotype is characterized by a protein-losing enteropathy, but neurological disorders
are usually absent. In a report of three siblings suffering from this type it was confirmed
that phosphomannose isomerase was deficient and that phosphomannomutase was
normal. In type II, the gene encoding for the Golgi localized N-acetylglucosaminyl
transferase enzyme has been cloned, and located on chromosome 14q. The clinical
picture in the few children described is much the same, with marked dysmorphic features
and severe developmental retardation, but it is of interest that the two children described
in a study had marked stereotypical behavior, which in the case of the girl included
tongue thrusting and hand-washing movements, similar to those of children with Rett
syndrome. Affected child, reported was severely retarded, but there was no peripheral
neuropathy and the cerebellum was normal. Two examples of type III were published
again there were dystrophic features, and initially hypotonia and severe non-progressive
retardation, but tetraparesis with brisk reflexes developed, as well as cerebral and optic
atrophy, infantile spasms, and pigmentary changes suggestive of incontinentia pigmenti.
80
The first girl had intermittently elevated liver enzyme levels and at the age of 8 years had
a stationary, severe, flaccid tetraparesis with areflexia, daily salaam fits, stable café-au-
lait skin changes and scoliosis. The second girl had hepatomegaly, a Dandy Walker
malformation and hypoplasia of the corpus callosum. Dystonic features have now
developed and the clinical picture is very different from type I (Hagberg, personal
communication). There is no evidence of polyneuropathy, retinal degeneration, or
cerebellar hypoplasia. Two infants with a fourth type have been described. Both were
microcephalic with severe epilepsy and delayed development, and showed dysmorphic
features. Several glycoproteins were abnormal, and it was considered that the clinical
and glycoprotein differences were sufficient to establish a separate entity.

In this condition there is a deficiency of the carbohydrate moieties of secretary


glycoproteins. Serum transferrin shows the most pronounced carbohydrate defect, both
quantitatively and qualitatively, with half of this glycoprotein apparently missing two or
four of its terminal trisaccharides. The deficiency of sialic acid, galactose, and N-
acetylglycosamine is found in both serum and the liver, and this failure of
glycoconjugate synthesis with an abnormal sialic acid transferrin pattern is a highly
specific marker for the syndrome. Many glycoproteins are abnormal, including transport
proteins, glycoprotein hormones, complement factors, lysosomal and other enzymes,
and enzyme inhibitors. Most enzymes are decreased in the serum but some, such as
arylsulphatase A, are increased. No doubt these changes are responsible for many of the
syndrome's manifestations, especially the phosphomannomutase deficiency shown to be
the major cause of type Ia carbohydrate-deficient glycoprotein syndrome. This enzyme
is essential for the biosynthesis of N-linked glycans. It can occasionally show normal
activity, and in type Ib there is a deficiency of phosphomannose isomerase. In type II
there is a deficiency of Golgi-localised N-acetylglucosaminyltransferase II,
demonstrated in fibroblasts and mononuclear blood cells, and necessary for the
biosynthesis of complex and hybrid N-linked glycans. There is a high mortality in the
first few years of life due to severe infections, status epilepticus, liver insufficiency, renal
failure, or cardiomyopathy. It was recorded a mortality of about 20% from these causes.
Post-mortem findings include olivopontocerebellar atrophy, loss of neurons and gliosis
in the cerebral cortex, basal ganglia, thalamus, and spinal cord. The peripheral nerves
show decreased myelin and multivacuolar inclusions in the Schwann cells. Other
findings include renal cysts, fibrosis of the liver with steatosis and glycogen storage,
fibrosis of the testes, and lymph node abnormalities. The neuropathy suggests an error of
macromolecular metabolism, and electron microscopy of liver cells shows lysosomal
vacuoles with concentric electron dense membranes and variable electron-lucent and
electron-dense material. In Carbohydrate-deficient glycoprotein syndromes one report,
there was no evidence of progression on a second biopsy. The presence of lysosomal
storage of an unknown nature affecting the anterior horn cells has been reported as an
unusual finding. There is no doubt that olivopontocerebellar atrophy is a common
finding in this condition, and the cause of the ataxia, although as it can be absent it is not a
reliable marker for the disorder.

81
Figure (43): Carbohydrate-deficient glycoprotein syndromes (www.google.com)

Figure (44): Carbohydrate-deficient glycoprotein type I (www.google.com)

10.Glycoproteinoses
Glycoproteinoses constitute a peculiar group of lysosomal storage disorders, in which
the common feature is the genetic abnormality of a lysosomal protein involved in the
catabolic pathway of glycoproteins. The deficient protein may be a glycosidase, a
cofactor or a lysosomal membrane carrier which delivers catabolic products to the
cytosol. Most lysosomal glycosidases are exo-hydrolases that degrade step-wise the
glycan at the nonreducing end. The lack of a single enzyme leads to the complete

82
blockage of the catabolic chain and results in the accumulation of undegraded
oligosaccharides in lysosomes. An elevated urinary excretion of carbohydrate material
is also observed in patients.

Figure (45): Glycoproteinoses (www.google.com)

10.1 Mannosidosis

10.1.1. α-Mannosidosis
The main clinical symptoms are progressive mental retardation, immunodeficiency, and
impaired hearing and Hurler-like skeletal changes with typical vertebral bodies, which
are prominently involved with ovoid configurations. Other findings are lens opacities,
muscular hypotonia, macroglossia and prognatheism. The clinical severity of α-
mannosidosis ranges in a continuum from mildly affected to severely affected patients,
nevertheless the disease is classically divided into two phenotypes: a severe infantile
phenotype referred to as type I and a milder juvenile-adult phenotype referred to as type
II .

The difference in the clinical course of the disease may be in part due to external factors
not relevant to the primary genetic defect. It is indicated that most patients do not show
any enzyme activity, although they are mildly affected. This suggests that other genetic
or environmental factors influence the clinical phenotype. Particularly it has been
advanced that the storage of oligomannosides in serum of the patients may interfere with
molecules of the immune system leading to the observed immunodeficiency among
mannosidosis patients.

Histological investigations demonstrate the presence of vacuolated lymphocytes,


hepatocytes and Kupffer cells.

83
Multiple forms of α-mannosidases with different subcellular locations have been
described. Lysosomal α-mannosidase (LAMAN) is affected in mannosidosis. This
enzyme presents a broad substrate specificity, hydrolyzing α(1–2), α(1–3) and α(1–6)
mannosyl linkages found in high-mannose and hybrid type glycans . This enzyme is
distinguished from other cellular α-mannosidase activities by a combination of a low pH
optimum (pH 4.5), Zn2+ dependence, a broad natural substrate specificity, an activity
towards the artificial substrate p-nitrophenyl-α-mannoside and an inhibition by
swainsonine. The enzyme belongs to the class II α-mannosidases and presents sequence
similarity to the Golgi α-mannosidase II, Golgi mannosidase IIx, mannosidase from
epididymis and the cytosolic/endoplasmic reticulum α-mannosidase . According to the
classification of Henrissat and Bairoch the enzyme belongs to family 38 of the
glycosylhydrolases. The lysosomal α-mannosidase has been purified from a number of
mammalian sources and is generally isolated as a complex consisting of 2 to 10 different
peptides. Biosynthetic studies in human fibroblasts have demonstrated that the enzyme
is synthesized as a 110 kDa precursor and is proteolytically cleaved into fragments of
40–46 kDa and 63–67 kDa upon transport to lysosomes.

Within a single tissue, acidic α-mannosidase occurs as isoenzymes: two isoenzymes


designated A and B have been isolated by ion-exchange chromatography and differ only
in their isoelectric point according to their degree of sialylation and/or phosphorylation
of the mannose residues.

In addition to the 'broad spectrum' acid α-mannosidases, a 'core specific' α-mannosidase,


hydrolyzing the Man(α1–6) linkage of the common trimannosyl core, has also been
characterized. Mannosidosis patients accumulate oligomannosides in tissues and urine .
These substrates originate from the incomplete catabolism of high-mannose and hybrid-
type glycans. High-mannose type oligomannosides are normally poorly present on
mature glycoproteins but are found on newly synthesized glycoproteins undergoing the
glycan biosynthetic trimming inside rough endoplasmic reticulum. They also occur on
polyprenic lipid intermediates (dolichol-oligosaccharides). One possible origin of high-
mannose oligosaccharides undergoing the lysosomal catabolism by α-mannosidase is in
the so-called 'quality control' of newly biosynthesized glycoproteins.

It is now well established that misfolded glycoproteins are retained inside the
endoplasmic reticulum (ER) by the two lectins calnexin and calreticulin and are finally
catabolized both in endoplasmic reticulum (ER) and cytosol. The generated
oligosaccharides are further transferred to lysosomes. Another origin for
oligosaccharides is in the catabolism of dolichol-phospho-oligosaccharides due to the
hydrolytic activity of the oligosaccharidyl-transferase enzyme.

Several observations may be deduced from the primary structures of stored


oligosaccharides.

1- All the oligosaccharides are terminated at the reducing end by a single GlcNAc
residue, due to the activity of the lysosomal chitobiase.
2- The major compound is the trisaccharide Man(α1–3)Man(β1–4)GlcNAc; the Man
(α1–6) counterpart is never found as well as the trimannose structure Man(α1–3)

84
[Man(α1–6)]Man(β1–4)GlcNAc which would be expected as the major stored
oligosaccharide. These observations could be related to the presence of residual
α1–6 mannosidase activity.
3- A number of oligosaccharides possess 'linear' structures with α1–2 linked mannose
on one branch. Such oligosaccharides may originate from the previous cytosolic
catabolism by the soluble α-mannosidase activity . Indeed, a complementary action
of cytosolic and lysosomal α-mannosidases has been demonstrated .

Figure (46): Pathway in genetic α-mannosidosis leading to accumulation of


Manα1 → 3Manβ1 → 4GlcNAc (www.google.com)

10.1.2 β-Mannosidosis
Genetic disorders associated with β-mannosidases deficiency have been described in
Nubian goats, Salers cattle, and humans. Ruminants affected by the disease present
severe neurological deficiencies associated with myelin abnormalities leading to a rapid
neonatal death, as well as facial dysmorphism contractures and hyperextension of the
limb joints and deafness.

Unlike the severe clinical manifestations of the disease in ruminants, the human disease
phenotype is generally milder, and heterogeneous in its clinical spectrum, with a wide
range of symptoms and age of onset. It is still difficult to distinguish a consistent pattern
of clinical manifestation for β-mannosidosis. The disease is associated with a range of
neurological involvements including various degrees of mental retardation except for
two cases, hearing loss and speech impairment, hypotonia, epilepsy and peripheral
neuropathy. There is no evidence for severe dysmyelination as observed in the animal
disease. Other clinical symptoms are angiokeratoma, facial dysmorphism, and skeletal
abnormalities.

The lysosomal enzyme β-mannosidase cleaves the β-linked mannose residue present in
all types of N-glycosylprotein glycans.

It could assume that the mammalian β-mannosidases are synthesized as one chain
precursor of about 110 kDa and subjected to limited N-terminal and C-terminal
proteolysis in a species-specific way. Human, goat and bovine β-mannosidases have
been cloned and mutations leading to the human and goat diseases identified. The human
gene was localized on chromosome 4q22–25. Mutation analysis in two siblings
85
differently affected with β-mannosidosis demonstrated a homozygous A→G transition,
resulting in two abnormally spliced mutant mRNA species in both siblings. The
sialylated trisaccharide NeuAc(α2–6)Man(β1–4)GlcNAc isolated from human β-
mannosidosis urine probably results from an abnormal sialylation of the Man(β1–4)
GlcNAc precursor.

Figure (47):β-Mannosidosis (www.google.com)

86
10.2 Fucosidosis
Fucosidosis is an autosomal recessive lysosomal storage disease caused by defective
lysosomal α-L-fucosidase with accumulation of fucosylated glycoconjugates in tissues.

The clinical features of fucosidosis are progressive mental retardation and neurological
deterioration, coarse facies, growth retardation, recurrent infections, dysostosis
multiplex, angiokeratoma. Traditionally, two clinical subtypes (type 1 and 2) are
generally differentiated. Type 1 is described as the severe infantile form with a general
deterioration of the neurological functions at the age of 1 or 2 years. Type 2 is a milder
form; symptoms appear progressively and the patients survive into the second or third
decade of life. It has been suggested that these two clinical subtypes were genetically
determined by different mutations. Excellent review of 77 cases demonstrate a wide
range of severity of the disease existing in the same family, moreover that type 1 and 2
represent two extremes of a continuous clinical spectrum. The cause of the clinical
variability of fucosidosis is not understood and suggests the importance of other genetic
or nongenetic factors.

α-L-Fucosidase is involved in the cleavage of terminal fucose linked α1–2 to a galactose


residue and of fucose linked α1–3, α1–4 or α1–6 to an N-acetylglucosamine residue on
glycoproteins and glycolipids. The enzyme is a homotetramer composed of subunits of
approximately 50 to 60 kDa: these molecular mass differences are due to variation in N-
glycosylation and proteolytic processing.

The α-L-fucosidase gene (FUCA 1) has been localized to chromosome 1p34 and
contains eight exons spanning 23 kb [110]. cDNA encodes an unprocessed protein of
461 amino acids, including 22 amino acids of the signal peptide and 439 amino acids
encoding the mature protein.

A pseudogene was found on chromosome 2 and is 80% identical to α-L-fucosidase


cDNA: it contains no intron and has no open reading frame . A locus on chromosome 6
(FUCA2) linked to the plasminogen locus influences plasma fucosidase activity but not
leukocyte fucosidase activity. This gene is not involved in fucosidosis.

More than 20 mutations in the FUCA 1 gene have been identified leading either to
fucosidosis or to a polymorphism . The most frequent mutation is a C to T transition that
results in the generation of an 'in frame' stop codon (CAA→TAA at codon 422)(Q422X)
that is found in six different families in Italy and in families in France, Cuba and Canada.
The other mutations result in premature stop defect, frameshift and alteration of splicing.
Amino acid substitutions occurring in conserved regions result in a severe form of the
disease (G60D, S63L, N379Y). A mutation L405R was described in a 46-year-old
patient with a less severe clinical phenotype. Amino acid substitution Q281R is
responsible for the polymorphism detected by isoelectric focusing (Fu1/Fu2). The
observed clinical variability is therefore not due to the nature of fucosidosis mutation,
but to secondary unknown factors.

α-L-Fucosidase enzyme activity in fucosidosis patients is nearly absent and CRIM


(cross reacting immunological material) is below 6% of the normal mean.

87
Fucosidosis patients accumulate in tissues and urine fucosylated glycoconjugates
originating from incomplete catabolism of N- and O-glycosylproteins, glycolipids and
proteoglycans. There is a major accumulation of glycolipids expressing blood group
antigens (H, X) in liver. Oligosaccharides and glycoasparagines are stored in tissues and
abnormally excreted in urine.

In the urine of patients with fucosidosis a large amount of glycopeptides is present. These
glycopeptides are characterized by the presence of an α1–6-linked fucose to the
chitobiose core. The major glycopeptide, a glycoasparagine Fuc(α1–6)GlcNAcβ1-Asn
present in urine corresponds to the linkage region of the oligosaccharide . The large
amount of glycopeptides excreted is unique to fucosidosis and can be explained by the
steric inhibition of the glycosylasparaginase by the fucose residue; the nonfucosylated
glycoasparagines are normally processed and only fucosylated glycoasparagine and
oligosaccharides without α1–6 linked fucose residue accumulate.

Analysis of urinary glycoconjugates from two patients with different phenotypes of


fucosidosis, type I (acute) and type II (chronic) indicated an increased amount of
Fucα1–6 GlcNAcβ1-Asn glycoasparagine in the type II patient as compared to the type I
patient. On the other hand a hexasaccharide possessing a Lewis X determinant at the
periphery of the oligosaccharide was present in much greater quantities in urines of the
type I patient. An inverse relationship between terminal oligosaccharide chain
fucosylation and core α1–6 fucosylation has been shown in the case of human blood
group H fucosyltransferase (Fuc-TI). If a similar inhibition is found with α1–3
fucosylation, it would explain the urinary phenotypes observed and furthermore a
correlation between the expression of the different fucosyltransferases (Fuc-TI, to
FucTVII) and the clinical phenotypes (type I and II) or a relationship between the nature
of the stored lysosomal material and the severity of the disease could be made.

Figure (48): Fucosidosis (www.google.com)


88
Figure (49): Metabolic pathways disorders in oligosaccharidoses
(www.google.com)

10.3 α-N-Acetylgalactosaminidase Deficiency (Schindler Disease and Kanzaki


Diseases)
The deficiency of lysosomal α-N-acetylgalactosaminidase (α-NAGA) was first reported
in Germany . The clinical heterogeneity and genetic and biochemical data suggest that
the phenotypic variation of this disease is influenced by numerous factors other than α-
N-acetylgalactosaminidase (α-NAGA).

Two clinical subtypes have been described. Type I α-N-acetylgalactosaminidase (α-


NAGA) deficiency is an infantile-onset form, designated as Schindler disease,
characterized by a neurodegenerative syndrome resembling Seitelberger disease
(infantile neuroaxonal dystrophy) . The first two patients described in Germany had a
normal development until the age of one and thereafter a rapid regression of their
previously acquired mental skills took place. Other neurological signs observed are
myoclonic seizures, cortical blindness, spasticity and decorticate posturing. Contrary to
other lysosomal storage diseases, no visceral abnormalities are observed. This
neuropathology characterized by the presence of 'spheroids' in terminal axons, classified
this disease as a neuroaxonal dystrophy. Type II is an adult-onset disease with mild
89
phenotypic expression and was first described in a 46-year-old Japanese woman with
angiokeratoma corporis diffusum, mild intellectual impairment, and peripheral
neuroaxonal degeneration. Two other adult patients with a very similar phenotype were
described in Spain. Another intermediate phenotype was described in two Dutch
children. As previously described for β-mannosidosis, an extreme diversity of the
clinical phenotypes is observed among patients.

α-N-acetylgalactosaminidase is a lysosomal enzyme, initially thought to be an


isoenzyme of α-galactosidase and thereafter called also α-galactosidase B. The two
enzymes have similar physico-chemical properties, a subunit molecular mass of 49 kDa,
homodimeric structure and a similar amino acid composition. α-N-
Acetylgalactosaminidase has an acidic optimal pH (pH 4.6) and hydrolyzes aryl-α-N-
acetylgalactosaminides and aryl-α-galactosides and is inhibited by the product of
hydrolysis, N-acetylgalactosamine . α-N-Acetylgalactosaminidase is known to act on N-
acetylgalactosamine linked to serine or threonine as an exoglycosidase. The α-N-
acetylgalactosaminidase gene has been localized to chromosome 22 and contains nine
exons spanning 14 kb. Four different mutations have been described. The German type I
probands had a G to A transition at nucleotide 973 that resulted in a substitution of a
glutamic acid to a lysine residue (E325K) . The type II Japanese patient had a substitution
of an arginine to a tryptophan residue (R329W). The Dutch patients showed two
mutations, the previously described E325K and a new one, S160C. Finally, the Spanish
patients are homozygous for the nonsense mutation E193X. It was demonstrated the
absence in all the different patients of immunologically detectable α-N-
acetylgalactosaminidase; the activity of the enzyme was barely detectable and only one
Dutch patient had a residual activity of 4%. Finally, no correlation can be made at the
moment between the severity of the disease and the mutation of the enzyme suggesting
that α-N-acetylgalactosaminidase deficiency is a multifactorial disease. It should be
pointed out that α-N-acetylgalactosamine occurs in the blood group A substance and thus
the blood group status of the patient may influence the degree of severity of the disease.
Some mutations may also affect the enzyme folding in the endoplasmic reticulum (ER)
and impair their arrival in the lysosome.

α-Linked N-acetylgalactosamine occurs:


1- at the termini of glycoconjugates as part of the blood group A determinant
(GalNAcα1–3(Fucα1–2)Gal) in glycolipids or in O- and N-linked oligosaccharides
of glycoproteins or as part of Forssman antigen on glycolipids: (GalNAcα1–
3GalNAcβ1– 3Galα1–4Galβ1–4Glcβ1–1′Cer);
2- at the reducing end of O-linked oligosaccharide. One would expect that the
compound accumulating in this lysosomal disease would always possess a terminal
GalNAc residue. The blood group A German patient described by Schindler and
colleagues excreted in the urine the blood group A trisaccharide and glycopeptides
possessing the common element GalNAcα1-O-Ser/Thr, substituted or not.
Remarkably, most of these glycopeptides possess galactose and sialic acid residues
attached to the core N-acetylgalactosamine residue and are similar to those excreted
in sialidosis and galactosialidosis patients. This might indicate that
a- the α-N-acetylgalactosaminidase is part of an enzymatic protein complex with
sialidase, galactosidase and the protective protein and, as a consequence of the

90
mutation of the α-N-acetylgalactosaminidase, the protein complex structure is less
efficient in the degradation of O-linked sialylated glycopeptides;
b- there is a phenomenon of transglycosylation on the primarily accumulated
GalNAcα1→O-Ser/Thr as observed in other lysosomal diseases.

For Further Readings:


1. Abott, M.; Wells, D.; Fallon, J. (1999). The insulin receptor tyrosine kinase
substrate p58/53 and the insulin receptor are components of CNS synapses. J
Neurosci, 19: 7300-7308.
2. Anick, D.; Harry, H.; Erik, Q.; Peter, H.; Luc, B.; Daniel, P.; Frans, S. (1995).
Human and rat beta cells differ in glucose transporter but not in glucokinase gene
expression. Journal of Clinical Investigation, 96 : 2489–95. doi:10.1172/
JCI118308. PMC 185903.
3. Arvanitakis, Z.; Wilson, R.; Bienias, J.; Evans, D.; Bennett, D. (2004). Diabetes
mellitus and risk of Alzheimer disease and decline in cognitive function. Arch
Neurol,61:661-666.
4. Ashraf, G.; Chibber, S.; Mohammad, N.; Zaidi, S.; Tabrez, S.; Ahmad, A.; Shakil,
S.; Mushtaq, G.; Baeesa, S.; Kamal, M. (2016). Recent updates on the association
between Alzheimer's disease and vascular dementia. Med. Chem.,12(3):226-237.
5. Aviles-Olmos, I.; Limousin, P.; Lees, A.; Foltynie, T. (2013). Parkinson's disease,
insulin resistance and novel agents of neuroprotection. Brain, 136:374-384.
6. Barbagallo, M.; Dominguez, L. (2014). Type2 diabetes mellitus and Alzheimer's
disease. World J Diabetes, 5(6):889-893.
7. Bruce, D.; Davis, W.; Casey, G.; Starkstein, S.; Clarnette, R.; Foster, J.; Almeida,
O.; Davis, T. (2008). Predictors of cognitive impairment and dementia in older
people with diabetes. Diabetologia, 51:241-248.
8. Butterfield, D.; Domenico,F.; Barone, E. (2014). Elevated risk of type 2 diabetes for
development of Alzheimer disease: a key role for oxidative stress in brain.
Biochimica et Biophysica Acta, 1842:1693-1706.
9. Camarata, S. (2014). Validity of early identification and early intervention in autism
spectrumdisorders; future directions. International Journal of Speech-Language
Pathology, 16(1):61-68.
10. ClaudeMichalski, PJ.; Klein, A. (1999). Glycoprotein lysosomal storage disorders:
α- and β-mannosidosis, fucosidosis and α-N-acetylgalactosaminidase deficiency.
Biochimica et Biophysica Acta (BBA)-Molecular Basis of Disease, 1455(2-3):69-
84.
11. Craft, S.; Watson, G. (2004). Insulin and neurodegenerative disease: shared and
specific mechanisms. Lancet Neurol, 3:169-178.
12. Cryer, P. (2007). Hypoglycemia, functional brain failure, and brain death. J Clin
Invest, 117(4):868-870.
13. Deepthi, B.; Sowjanya, B.; Bhargavi, R.; Babu, P. (2017). A modern review of
diabetes mellitus: an annihilatory metabolic disorder. Journal of In Silico & In Vitro
Pharmacology, 3(1).
14. de la Monte, S.; Wands, J. (2008). Alzheimer's disease is type 3 diabetes-evidence
reviewed. J Diabetes Sci Technol, 2(6):1101-1113.
15. Delvenne, V.; Goldman, S.; De Maertelaer, V.; Simon, Y.; Luxen, A.; Loststra, F.
(1996). Brain hypometabolism of glucose in anorexia nervosa: normalization after

91
weight gain. BIOL PSYCHIATRY, 40:761-768.
16. Falkowska, A.; Gutowska, I.; Goschorska, M.; Nowacki, P. (2015). Energy
metabolism of the brain, including the cooperation between astrocytes and neurons,
especially in the context of glycogen metabolism. International Journal of
Molecular Sciences, 16:25959-25981.
17. Gaetano, S.; Gennaro, P.; Celestino, S.; Wenjun, X.; Steven, R.; Luca, DA.;
Michele, C.; Nicola, M.; Bruno, T.; Theresa, G.; Alain, L.; Andrew, M. (2015).
Calcium release channel RyR2 regulates insulin release and glucose homeostasis.
Journal of Clinical Investigation. 125 (5): 1968–78. doi:10.1172/JCI79273. ISSN
0021-9738. PMC 4463204. PMID 25844899.
18. Gordon, N. (2000). Carbohydrate-deficient glycoprotein syndromes. Postgrad Med
J, 76:145-149.
19. Keizer, J.; Magnus, G. (1989). ATP-sensitive potassium channel and bursting in the
pancreatic beta cell. A theoretical study. Biophysical Journal. 56 (2): 229–42.
Bibcode:1989BpJ....56..229K. doi:10.1016/S0006-3495(89)82669-4. PMC
1280472. PMID 2673420.
20. Khan, KJ.; and Myers, R. (1971). Insulin-induced hypoglycemia in the non-human
primate. 1. Clinical consequences. In Brain hypoxia. Brierly, J.; Meldrum, B.
William Heinemann Medical Books Ltd. London, United Kingdom. 185-194.
21. Lang, V.; Light, PE. (2010). The molecular mechanisms and pharmacotherapy of
ATP-sensitive potassium channel gene mutations underlying neonatal diabetes.
Pharmgenomics. Pers. Med. 3: 145–61. doi:10.2147/PGPM.S6969. PMC
3513215. PMID 23226049.
22. Lebovitz, H.(2001). Insulin resistance: definition and consequences. Exp Clin
Endocrinol Diabetes, 109 (suppl.2): S135-S148.
23. Leszek, J.; Trypka, E.; Tarasov, V.; Ashraf, G.; Aliev, G. (2017). Type 3 diabetes
mellitus: a novel implication of Alzheimers disease. Current Topics in Medicinal
Chemistry, 17:1331-1335.
24. Lima, M.; Targa, A.; Noseda, A.; Rodrigues, L.; Delattre, A.; Santos, F.; Fortes, M.;
Maturana, M.; Ferraz, A. (2014). Does Parkinson's disease and type 2 diabetes
mellitus present common pathophysiological mechanisms and treatments?. CNS &
Neurological Disorders.Drug Targets, 13:418-428.
25. Mason, G.; Petersen, K.; Lebon, V.; Rothman, D.; Shulman, G.(2006). Increased
brain monocarboxylic acid transport and utilization in type1 diabetes. Diabetes,
55:929-934.
26. Mclntyre, R.; Mancini, D.; Pearce, M.;Silverstone, P.; Chue, P.; Misener, V.;
Konaraki, J. (2005). Mood and psychotic disorders and type 2 diabetes: a metabolic
triad. CANADIAN JOURNAL of DIABETES, 29(2):122-132.
27. Mizugishi, K.; Yamanaka, K.; Kuwajima, K.; Yuasa, I.; Shigemoto, K.; Kondo, I.
(1999). Missence mutations in the phosphomannomutase 2 gene of two Japanese
siblings with carbohydrate-deficient glycoprotein yndrome type1. Brain Dev,
21:223-228.
28. Moheet, A.; Mangia, S.; Seaquist, E. (2015). Impact of diabetes on cognitive
function and brain structure. Ann N Y Acad Sci, 1353:60-71.
29. Palacios, N.; Gao, X.; McCullough, M.; Jacobs, E.; Patel, A.; Mayo, T. (2011).
Obesity, diabetes, and risk of Parkinson's disease. Mov Disord, 26:2253-2259.
30. Patti, M.; Butte, A.; Crunkhorn, S.; Kusi, K.; Berria, R.; Kashyap, S. (2003).

92
Coordinated reduction of genes of oxidative metabolism in humans with insulin
resistance and diabetes: potential role of PGC1and NRF1. Proc Natl Acad Sci,
100:8466-8471.
31. Petersen, K.; Dufour, S.; Befroy, D.; Garcia, R.; Shulman, G. (2004). Impaired
mitochondrial activity in the insulin-resistant offspring of patients with type2
diabetes. N Engl J Med, 350:664-671.
32. Pruzin, J.; Nelson, P.; Abner, E.; Arvanitakis, Z. (2018). Review: relationship of
type2 diabetes to human brain pathology. Neuropathology and Applied Neurology,
44: 347-362.
33. Ramaekers, VT.; Stibler, H.; Kint, J.; Jaeken, J. (1991). A new variant of the
carbohydrate deficient glycoprotein syndrome. J Inherit Metab Dis, 4:385-388.
34. Ronnemaa, E.; Zethelius, B.; Sundelof, J.; Sundstrom, M.; Degerman-Cunnasson,
M.; Berne, C.; Lannfelt, L.; Kilander, L. (2008). Impaired insulin secretion
increases the zheimer disease. Neurology, 71:1065-1071.
35. Schrier, W. (2007). The internal medicine casebook real patients, real answers (3rd
ed.). Philadelphia: Lippincott Williams & Wilkins. p. 119. ISBN 9780781765299.
Archived from the original on 1 July 2015.
36. Shaw, K. (2017). Type 3 diabetes: a brain insulin-resistant state linked to
Alzheimer's disease. PRACTICAL DIABETES, 34(6):187-188.
37. Stranahan, A.; Arumugam, T.; Cutler, R.; Lee, K.; Egan, J.; Mattson, M. (2008).
Diabetes impairs hippocampal function through glucocorticoid-mediated effects
on new and mature neurons. Nat Neurosci, 11:309-317.
38. Suh, S.; Gum, E.; Hamby, A.; Chan, P.; Swanson, R. (2007). Hypoglycemic
neuronal death is triggered by glucose reperfusion and activation of neuronal
NADPH oxidase. J. Clin. Invest., 117:910-918.
39. Sullivan, C.; Donovan, S.; McCullumsmith, R.; Ramsey, A. (2017). Defects in
bioenergetics coupling in schizophrenia. Society of Biological Psychiatry.
40. Talbot, K.; Wang, H. (2014). The nature, significance, and glucagon-like peptide-1
analog treatment of brain insulin resistance in Alzheimer's disease. Alzheimer's
Dement, 10:S12-S25.
41. Van den Bergh, B. (2011). Developmental programming of early brain and
behavior development and mental health: a conceptual framework. Developmental
Medicine and Child Neurology, 53(suppl 4):19-23.
42. Wyss, M.; Jolivet, R.; Buck, A.; Magistretti, P.; Weber, B. (2011). In vivo evidence
for lactate as a neuronal energy source. J. Neurosci. 31:7477-7485.
43. Xu, G.; Jing, J.; Bowers, K.; Liu, B.; Bao, W. (2014). Maternal diabetes and the risk
of autism spectrum disorders in the offspring: a systematic review and meta-
analysis. J Autism Dev Disord, 44(4):766-775.
44. Yanai, H.; Adachi, H.; Katsuyama, H.; Moriyama, S.; Hamasaki, H.; Sako, A.
(2015). Causative anti-diabetic drugs and the underlying clinical factors for
hypoglycemia in patients with diabetes. World Journal of Diabetes, 6 (1): 30–6.
doi:10.4239/wjd.v6.i1.30. PMC 4317315. PMID 25685276.
45. Yao, Y.; Walsh, W.; McGinnis, W.; Pratico, D. (2006). Altered vascular phenotype
in autism: correlation with oxidative stress. Archives of Neurology, 63:1161-1164.

93
CHAPTER FOUR
LIPIDS AND HUMAN BRAIN

1.Preview
Lipids are heterogeneous group of water-insoluble (hydrophobic) organic molecules.
Because of their insolubility in aqueous solutions, body lipids are generally found
compartmentalized as in the case of membrane-associated lipids or droplets of
triacylglycerol in adipocytes, or transported in plasma in association with protein as in
lipoprotein particles or on albumin. Lipids are major source of energy for the body.

Lipids originate from two sources: endogenous lipids synthesized in the liver and
exogenous lipids which are ingested and absorbed in the intestine.

The lipids can be classified as total cholesterol (TC) and its derivatives such as
triglycerides (TGs), low-density-lipoprotein-cholesterol (LDL-C), high-density-
lipoprotein-cholesterol (HDL-C), and very-low-density-lipoprotein-cholesterol
(VLDL-C).

Fat is an essential component in the diet, whether it is part of an oral diet, an enteral
nutrition formula, or part of a parenteral nutrition (PA) admixture. The human body
needs fat stores to cushion organs and provide insulation for temperature regulation. The
fat depot can be used for energy during times of starvation, although it is important to
recognize that some tissues in the body (brain and red blood cells) rely solely on glucose
for energy as fat cannot be metabolized to create glucose. Dietary fat is not only an
energy source through oxidation, it is also required to facilitate absorption of fat-soluble
vitamins in the small bowel. The cell membrane is composed of phospholipids, which
are sensitive to chemical signaling.

Fasting for more than 8h normally only occurs a few hours before breakfast. By contrast,
the non-fasting state predominates for most of a 24-h cycle. Therefore, because plasma
contains atherogenic lipoproteins of hepatic origin in the fasting state and additionally
those of intestinal origin in the non-fasting state, the non-fasting state may better capture
the total amount of atherogenic lipoproteins in plasma during the majority of a 24-h
period.

Plasma contains remnant lipoproteins of hepatic origin in the fasting state, whereas
remnant lipoproteins of intestinal origin additionally are present in the non-fasting state.
This means that from 1 to 7h after a habitual meal, plasma triglycerides and remnant
cholesterol are slightly elevated. A habitual meal have means whatever the person chose
to eat before blood sampling, which naturally will differ from person to person and from
country to country. The degree of plasma triglyceride (TG) elevation is related to levels
at baseline, the lower the baseline triglycerides (TGs) the smaller the postprandial effect
and vice versa per equivalent fat load.
94
Standard lipid profile involves triglycerides (TGs), total cholesterol (TC), low-density-
lipoprotein-cholesterol (LDL-C), high-density-lipoprotein-cholesterol (HDL-C),
remnant cholesterol, and non-high-density-lipoprotein-cholesterol (non-HDL-C).

Expanded lipid profile includes triglycerides (TGs), total cholesterol (TC), low-density-
lipoprotein-cholesterol (LDL-C), high-density-lipoprotein-cholesterol (HDL-C),
remnant cholesterol, non-high-density-lipoprotein-cholesterol (non-HDL-C),
lipoprotein (a) [Lp(a)], and apoB.

A minimal lipid profile consists of plasma total cholesterol and triglycerides. A standard
lipid profile also includes measurements of low-density lipoprotein (LDL), cholesterol,
and high-density lipoprotein cholesterol (HDL-C). Total cholesterol (TC), high-density
lipoprotein cholesterol (HDL-C), and triglycerides (TGs) are measured directly,
whereas low-density-lipoprotein-cholesterol (LDL-C) can either be measured directly
or calculated by the Friedewald equation if triglycerides (TGs) are <400mg/dL
(<4.5mmol/L)=Total cholesterol-HDL cholesterol-triglycerides/2.2 (all in mmol/L, or
–triglycerides/5 with values in mg/dL), with direct measurement of low-density-
lipoprotein-cholesterol at triglyceride concentrations ≥400mg/dL (4.5mmol/L).
Traditionally, the Friedewald equation has been applied to a fasting lipid profile,
however, calculated low-density-lipoprotein-cholesterol determined with this equation
at triglyceride concentrations <400mg/dL (4.5mmol/L) is similar to low-density-
lipoprotein-cholesterol (LDL-C) measured directly on both fasting and nonfasting lipid
profiles. An expanded lipid profile should be a standard one, with inclusion of
lipoprotein (a) [Lp(a)] measurement. Lipoprotein (a) [Lp(a)] determination should not
be included in repeated lipid profile measurements in the same patient, unless
therapeutic intervention is aimed at reducing lipoprotein (a) [lp(a)] concentrations.

Glycolipids were discovered and named by Ernst Klenk after their isolation from brain
tissue in 1942. They are ubiquitous membrane constituents, which are embedded in the
cell plasma membrane. Glycolipids are glycosyl derivatives of lipids. They are
collectively a part of a larger family of substances known as glycoconjugates.

Figure (50): classes of lipids (www.google.com)


95
Second only to adipose tissue, the brain is the most lipid concentrated organ in the body.
Nervous tissue contains 50% lipid on a dry weight basis, or 10% lipid on a weight/weight
basis. This lipid plays a role in modifying the structure, fluidity, and function of brain
membranes. A variety of complex lipids exists in the brain, and the composition and
metabolism of these lipids change with development and age. It should be noted that a
difference in the deposition and net absorption of fat exists between perinatal and
postnatal growth.

It is becoming increasingly apparent that the central nervous system (CNS) is a major
contributor in the regulation of systemic metabolism and lipid balance. In the central
nervous system (CNS) the nutritional status of the body is constantly being surveyed and
assessed by energy sensing regions of the brain, such as the hypothalamus. Key nuclei
within the hypothalamus, such as the ventromedial nucleus (VMH), arcuate nucleus
(ARC), dorsomedial hypothalamic nucleus (DMH), and the paraventricular nucleus
(PVN), integrate signals to elicit peripheral responses, such as changes in feeding
behavior, fuel mobilization, energy utilization, and energy storage. These nuclei detect
both nutrients and nutritionally regulated endocrine factors, such as insulin, ghrelin,
melanocortin (MC), and leptin, in order to regulate feeding and energy balance.

Figure (51): Map of hypothalamic areas involved in nervous control of energy


homeostasis (www.google.com).

2.Lipid Digestion, Absorption, Metabolism


Digestion of fat starts in the mouth with salivary lipase; when food enters the stomach
there is exposure to gastric lipase, although this primarily digests medium-and short-
chain fatty acids. As chyme is released from the stomach into the duodenum, fat is
emulsified by bile, while pancreatic lipase and colipase digest fat into free fatty acids and
monoglycerides that are packaged into micelles (~200 times smaller than emulsion
droplets). The micelles transport to free fatty acids and monoglycerides to the brush
border of the distal jejunum and ileum for absorption. Once inside the enterocyte,
monoglycerides and fatty acids are synthesized into triglycerides (TGs) with

96
cholesterol, fat-soluble vitamins, and phospholipids into chylomicrons. Chylomicrons
are transported via the lymphatic system to the liver, adipose, and muscle for additional
metabolism and/or storage. Within the cell, fatty acids are metabolized through
desaturation and elongation. When considering the essential fatty acids, ALA (an
omega-3fatty acid) is metabolized preferentially over LA (an omega-6 fat); when either
fat is not available or limited, oleic acid (an omega-9 fat) is metabolized.

Figure (52): Absorption of dietary lipids (www.google.com)

Figure (53): Emulsification and digestion of fat (www.google.com)


97
3.Lipid Entrance to Brain
The brain has the second highest lipid content behind adipose tissue, and brain lipids
constitute 50% of the brain dry weight. Unlike adipose tissue, which largely stores fatty
acids (FAs) as triglycerides (TGs) for subsequent utilization and mobilization to other
metabolic tissues, the brain is thought to mainly utilize acylated lipids to generate
phospholipids for cell membranes. The fatty acid (FA) composition of the brain is unique
and is rich in long-chain polyunsaturated fatty acids (LC-PUFAs), particularly
arachidonic acid (AA), eicosapentaconic acid, and docosahexacnoic acid (DHA).
Although some fatty acids (FAs) can be synthesized de novo, essential fatty acids (EFAs)
must be transported into the brain from the systemic circulation. A number of studies
have shown that fatty acids (FAs) are able to cross the blood-brain barrier (BBB) and
enter neurons. Fatty acids (FAs) cross the blood-brain barrier (BBB) and that albumin
may be important in this process.

3.1 The Membrane Localized Fatty Acid Transporter Proteins


Fatty acid transport protein 1(FATP1) and fatty acid transport protein 4 (FATP4) appear
to be the predominant fatty acid transport proteins expressed in the blood-brain barrier
(BBB). The fatty acid translocase/CD36 plays a prominent role in the transport of fatty
acids (FAs) across human brain microvessel endothelial cells.

Figure (54): Fatty acid transport into the brain (www.google.com)

4.Fatty Acids
Fatty acids (FAs) exist free in the body, that is, they are unesterified, and are also found as
fatty acid esters in more complex molecules, such as triacylglycerols (TGs). Low levels
of free fatty acids (FFAs)occur in all tissues, but substantial amounts can sometimes be
found in the plasma, particularly during fasting. Plasma free fatty acids (FFAs)
(transported by serum albumin) are in route from their point of origin (triacylglycerol of
adipose tissue or circulating lipoproteins) to their site of consumption (most tissues).
Fatty acids (FAs) can be oxidized by many tissues particularly liver and muscle to
provide energy. Fatty acids (FAs) are also structural components of membrane lipids,
such as phospholipids and glycolipids. Fatty acids are attached to certain intracellular
proteins to enhance the ability of those proteins to associate with membranes. Fatty acids
(FAs) are the precursors of the hormone-like prostaglandins. Esterified fatty acids in the
form of triacylglycerols stored in adipose cells serve as the major energy reserve of the

98
body. A fatty acid (FA) consists of a hydrophobic hydrocarbon chain with a terminal
carboxyl group. Fatty acid (FA) chains may contain no double bonds that is saturated or
contain one or more double bonds that is mono or polyunsaturated. Fatty acids (FAs) can
be classified by the number of double bonds: saturated fats -0, monounsaturated fats-1,
and polyunsaturated fats ≥2.

Figure (55): Classification of fatty acids based on number of double bonds


(www.google.com)

Figure (56): saturated and unsaturated fatty acids (www.google.com)

99
Figure (57): Types of fatty acids (www.google.com)

Fatty acids can be classified based on their length, with short chain fatty acids (FAs)
having 2-4 carbon atoms, medium chain fatty acids having 6-12 carbon atoms, and long-
chain fatty acids having 12-24 carbon atoms.

Figure (58): Classification of fatty acids based on their length (www.google.com)

The essential fatty acids (EFAs) are those that cannot be synthesized as humans lack the
enzymes required. Two fatty acids (FAs) are dietary essentials in humans, linoleic acid
(LA), which is the precursor of arachidonic acid, the substrate for prostaglandin
synthesis and α-linolenic acid (LNA) the precursor of other fatty acids (FAs) important
for growth and development. Essential fatty acid (EFA) deficiency can result in
neurologic abnormalities.
100
Figure (59): Role of essential fatty acids (www.google.com)

A large proportion of fatty acids (FAs) used by the body is supplied by the diet.
Carbohydrates, proteins, and other molecules obtained from the diet in excess of the
body's needs for those compounds can be converted to fatty acids which are stored as
triacylglycerols (book chemistryCH16).

4.1 Fatty Acid Synthesis


In adult humans, fatty acid (FA) synthesis occurs primarily in the liver and lactating
mammary glands and to a lesser extent in adipose tissue. The process incorporates
carbons from acetyl CoA into the growing fatty acid chain, using adenosine triphosphate
(ATP) and reduced nicotinamide adenosine dinucleotide phosphate (NADPH). Fatty
acids (FAs) are synthesized in the cytosol. Citrate carries two-carbon acetyl units from
the mitochondrial matrix to the cytosol. The regulated step in fatty acid synthesis is
catalyzed by acetyl CoA carboxylase which requires biotin. Citrate is the allosteric
activator and long-chain fatty acyl CoA is the inhibitor. The enzyme can also be activated
in the presence of insulin and inactivated by epinephrine or a rise in adenosine
monophosphate (AMP). The most of the steps in fatty acid synthesis are catalyzed by the
multifunctional enzyme, fatty acid synthase, which produces palmitoyl CoA from acetyl
CoA and malonyl CoA, with nicotinamide adenosine dinucleotide phosphate (NADPH)
as the source of reducing equivalents. Fatty acids (FAs) can be elongated and desaturated
in the endoplasmic reticulum (ER). When fatty acids are required by the body for energy,
adipose-cell hormone-sensitive lipase (activated by epinephrine or glucagon and
inhibited by insulin) initiates degradation of stored triacylglycerol. Fatty acids (FAs) are
carried by serum albumin to the liver and peripheral tissues, where oxidation of the fatty
acids provides energy.

101
Figure (60): Synthesis of fatty acids (www.google.com)

4.2 Oxidation of Fatty Acids


Fatty acid degradation (β-oxidation) occurs in mitochondria. The carnitine shuttle is
required to transport fatty acids (FAs) from the cytosol to the mitochondria. The
enzymes required are carnitine palmitoyl transferases I and II. Carnitine palmitoyl
transferase I is inhibited by malonyl CoA. This prevents fatty acids being synthesized in
the cytosol from malonyl CoA from being transported into the mitochondria where they
would be degraded.

Figure (61): Activation and transportation of fatty acid via the Carnitine shuttle
102
Once in the mitochondria, fatty acids are oxidized producing acetyl CoA, reduced
nicotinamide adenine dinucleotide (NADH), and dihydroflavin adenine dinucleotide
(FADH2). The first step in the β-oxidation pathway is catalyzed by one of a family of four
acetyl CoA dehydrogenases, each of which has a specificity for either short, medium,
long,or very-long-chain fatty acids.

Figure (62): β-oxidation of fatty acid (www.google.com)

Oxidation of fatty acids with an odd number of carbons proceeds two carbons at a time
(producing acetyl CoA) until three carbons remain (propionyl CoA). This compound is
converted to methyl malonyl CoA (a reaction requiring biotin), which is then converted
to succinyl CoA (a gluconeogenic precursor) by methyl malonyl CoA mutase (requiring
vitamin B12).

Figure (63): β-oxidation of odd carbon chain length fatty acids


103
Β-oxidation of very long chain fatty acids (VLCFA) and α-oxidation of branched-chain
fatty acids occurs in peroxisomes.

Figure (64): α-oxidation of branched chain fatty acids (www.google.com)

Liver mitochondria can convert acetyl CoA derived from fatty acid oxidation into the
ketone bodies, acetoacetate, and 3-hydroxybutyrate. Peripheral tissues processing
mitochondria can oxidize 3-hydroxybutyrate to acetoacetate, which can be converted to
acetyl CoA, thus producing energy for the cell. Unlike fatty acids, ketone bodies can be
utilized by the brain, and therefore are important fuels during fast. The liver lacks the
ability to degrade ketone bodies, and so synthesizes them specifically for the peripheral
tissues.

Free fatty acid (FFA) in the serum reflects the level of lipid metabolism and mediates cell
damage caused by oxidative stress. Brain free fatty acid (FFA) content is extremely low
under normal conditions, but it increases when circulation disorders occur in the brain.

4.3 Neuronal Uptake of Fatty Acids


Once fatty acids (FAs) have traversed the blood-brain barrier (BBB), fatty acid
transporters may play a key role in facilitating fatty acid (FA) uptake into neurons. For
example, dissociated neurons from the ventromedial nucleus (VMH) express both fatty
acid transporter protein 1 (FATP1) and CD36. Since CD38 is an established gustatory
lipid sensor, it is plausible that other lipid sensors involved in the chemoreception of long
chain fatty acids (LCFAs)/omega-3 fatty acids (ω−3 FAs), such as G-protein coupled
receptor 120 (GPR 120), are also involved in neuronal lipid sensing. Fatty acid binding
protein 3 (FABP3), which is localized in neurons, facilitates brain arachidonic acid (AA)
but not palmitic acid (C16.0) uptake and trafficking into specific brain lipid pools.

104
Lipid metabolism in astrocytes plays a key role in fatty acid (FA) sensing, since fatty acid
(FA) oxidation is thought to occur predominantly in the astrocyte rather than the neuron.
Both neurons of the ventromedial nucleus (VMH) and astrocytes have been shown to
express many of the fatty acid transporters, including fatty acid transporter protein 1
(FATP1), fatty acid transporter protein 4 (FATP4), and CD36. The fatty acid bind protein
7 (FABP7) is important for astrocyte-neuron lipid homeostasis.

4.4 Hypothalamic Fatty Acid Metabolism


Lipid utilization is a critical process to neuronal function. Reactive oxygen species
(ROS) may be involved in neuronal lipid metabolism and activation. Suppression of
reactive oxygen species (ROS) activates neuropeptide Y (NPY)/Agouti-related protein
(AgRP) neurons to promote feeding, whereas activation of reactive oxygen species
(ROS) activates prooptiomelanocortin (POMC) neurons to reduce feeding. It is likely
that neuronal reactive oxygen species (ROS) accumulation may intrinsically link
neuronal substrate metabolism to feeding behavior and systemic energy balance. AMP-
activated protein kinase may also act as cellular energy sensor within neurons to link
neuronal lipid metabolism to systemic lipid metabolism and energy balance. Adenosine
monophosphate kinase (AMPK) is widely expressed in the arcuate nucleus (ARC),
paraventricular nucleus (PVN), and ventromedial nucleus (VMH) of the hypothalamus
and is able to sense intracellular energy status by the adenosine monophosphate
(AMP)/adenosine triphosphate (ATP) ratio and the level of adipokines (e.g., leptin and
ghrelin). Activated adenosine monophosphate kinase (AMPK) responds to the cellular
energy status and can switch cellular metabolism toward catabolic processes that
produce adenosine triphosphate (ATP) and away from anabolic processes that consume
adenosine triphosphate (ATP). Hypothalamic adenosine monophosphate kinase
(AMPK) may also regulate downstream lipid metabolism through brown adipose tissue
(BAT) thermogenesis. Hypothalamic adenosine monophosphate kinase (AMPK) is
involved in the autonomic regulation of brown adipose tissue (BAT) thermogenesis,
specifically in the sympathetic nervous system (SNS).

AMP-activated protein kinase also plays a key role in the hypothalamic response to
leptin in the context of high-fat feeding. Leptin results in reduced hypothalamic
adenosine monophosphate kinase (AMPK) activity. Hypothalamic phosphor-AMPK is
also modulated by the fatty acid (FA) composition of the diet and is dependent on brain
region and metabolic status. adenosine monophosphate kinase (AMPK) regulates the
activity of a number of enzymes involved in the synthesis of complex lipids that are
critical for optimal brain function and metabolism.

There is growing evidence to suggest that the accumulation of fatty acid (FA)
metabolites may signal nutrient status and thus may be critical to central lipid sensing
and the modulation of systemic metabolism. For example, upon entry into the neuron,
long-chain fatty acids (LCFAs) are esterified to LCFA-CoA by Acyl-CoA synthetase
(ACS). It is thought that this accumulation of fatty acid (FA) derived from LCFA-CoA
triggers a lipid sensing mechanism to inhibit hepatic glucose production (HGP) and to
maintain systemic glucose homeostasis. In metabolic tissues, intracellular LCFA-CoAs
enter the mitochondria via Carnitine palmitoyltransferase I (CPT-I), where they are then
subject to fatty acid β-oxidation. Importantly, the liver isoform of Carnitine
palmitoyltransferase I (CPT-I), CPT-Iα, is prevalent in the hypothalamus, and inhibition
105
of hypothalamic Carnitine palmitoyltransferase-Iα (CPT-1α) causes an increase in
intracellular LCFA-CoA, which triggers a satiation signal, leading to reduced systemic
glucose production and food intake. The brain also expresses a neuron-specific isoform
of Carnitine palmitoyltransferase I (CPT-I), Carnitine palmitoyltransferase Ic (CPT-Ic),
which is found in the endoplasmic reticulum (ER) of key energy-sensing nuclei of the
hypothalamus [arcuate nucleus (ARC)] and has also been repeatedly implicated in the
modulation of systemic metabolism (paper53). Carnitine palmitoyltransferase I (CPT-I)
has been suggested to act downstream of malonyl CoA in the hypothalamic control of
feeding. This is particularly pertinent to fatty acid (FA) metabolism in hypothalamic
neurons since Carnitine palmitoyltransferase I (CPT-I) activity is inhibited by malonyl
CoA, and thus elevated malonyl-CoA leads to an accumulation of LCFA-CoA, which
has been previously referred to as a satiety signal. Importance of hypothalamic sensing
of circulating lipids in the maintenance of hepatic metabolism and systemic glucose
homeostasis.

Figure (65): Medial basal hypothalamic nuclei control

Figure (66): Hypothalamic satiety network (www.google.com)


106
4.5 Essential Fatty Acids in Brain
Essential fatty acids include linoleic (LA~18:2n-6) and α-linolenic (LNA~18:3n-3)
acids. The abbreviations n-6 and n-3 for these unsaturated fatty acids represent the
position of the first double bond when counting from the methyl carbon atom at the distal
end of the fatty acid (FA) chain. These fatty acids (FAs) and their respective derivatives
are also commonly referred to as omega-6 and omega-3 fatty acids (FAs) respectively.
Both these fatty acids (FAs) are required for synthesis of long-chained fatty acids, as they
cannot be synthesized de novo and consequently are referred to as essential fatty acids
(EFAs).In addition, n-6 and n-3 poly unsaturated fatty acids (PUFAs) such as dihomo-
gamma-linolenic acid (DGLA020:3n-6), arachidonic acid (AA020:4n-6) and
eicosapentaenoic acid (EPA020:5n-3) are precursors for eicosanoids biosynthesis,
which can exert a wide range of biological actions. Whilst eicosapentaenoic acid (EPA)
content of membranes is a source for eicosanoid synthesis, docosahexaenoic acids
(DHA) content may also play a contributory role, both directly and indirectly. Although
a relatively unexplored field, docosahexaenoic acids (DHA) can function as a source for
eicosapentaenoic acid (EPA) through a retroconversion reaction, which in turn will be
oxygenated to the various eicosanoid metabolites. Direct oxygenated products of
docosahexaenoic acids (DHA), resulting from peroxidative damage have also been
identified and are commonly referred to as docosanoids.

The majority of membrane poly unsaturated fatty acids (PUFAs) are synthesized from
dietary linoleic acid (LA) and linolenic acid (LNA), which act as precursors for the
synthesis of these longer-chained poly unsaturated fatty acids (PUFAs) via a series of
desaturation and elongation reactions.

Figure (67): Chemical structure of linoleic acid and α-linolenic acid


(www.google.com)

107
Figure (68): Polyunsaturated fatty acid metabolic pathways in mammals
(www.google.com)

4.6 Brain Membrane Structure and Function Related to Fatty Acid Profile
Cell membrane provide domains in which very active enzymes, ions, and transporters
can be found. The term "membrane fluidity" refers to the physical state of the fatty acyl
chains comprising the membrane bilayer structure, as well as measure of the different
rates of motion of molecule elements within the membrane. The fatty acyl chains, which
elicit the most influence, such as arachidonic acid (AA), eicosapentaenoic acid (EPA),
and docosahexaenoic acids (DHA), contain double bonds, which are exclusively in the
cis conformation. The replacement of even a single double bond in these poly
unsaturated fatty acids (PUFAs) is sufficient to exert a profound effect on the physical
properties of the membrane.
108
Polyunsaturated fatty acids (FAs) have been shown to regulate neuronal excitability. It
was reported that age-related declines in poly unsaturated fatty acids (PUFAs) resulted
in a decrease of Na+, K+, ATPase activity in brain synaptosomes. It was found a
correlation between docosahexaenoic acids (DHA) and phospholipase A2 (PLA2)
activity, a rate limiting enzyme involved in the hydrolysis of arachidonic acid (AA) from
membrane for metabolism into eicosanoids. The involvement of docosahexaenoic acids
(DHA) in synaptic signal transduction has been reported. A primary role for
docosahexaenoic acids (DHA) in phospholipid mediated signal transduction at the
synapse involving activation of phospholipase A2 (PLA2) and/or C. Changes in
phospholipid unsaturation may have profound effects on signal transduction pathways
in which protein kinase C (PKC) plays a contributory role. Several behavioral aspects of
brain function have been shown to be affected by dietary fatty acids (FAs). Studies have
reported a correlation between changes in behavior and learning parameters with
changes in brain membrane poly unsaturated fatty acids (PUFAs) composition. It has
been suggested that the growing fetus obtains its supply of essential fatty acids (EFAs)
from its mother's essential fatty acids (EFAs) stores. Thus, if these stores are limited or
not balanced with the correct poly unsaturated fatty acids (PUFAs), there is the potential
risk for structural aberrations, which may have deleterious effects on fetal and/or
neonatal brain functions. Neurogenesis peaks around the 14th week of gestation and is
completed by around the 25th week, once adult neuron numbers are attained. Just prior to
completion of neurogenesis, functional connections between target cells in the nervous
system are established, through the growth of neurites (axons and dendrites) and the
formation of synapses, known as synaptogenesis. Following neuron development, glial
cells begin to originate (gliogenesis). Unlike neurons, glial cell production continues
throughout adult life. A necessity for docosahexaenoic acids (DHA), and hence
sufficient dietary consumption by the mother, that can be utilized by the developing fetus
are essential. Reports showed the importance of docosahexaenoic acids (DHA) during
gliogenesis.

4.7 Essential Fatty Acids Precursors of Eicosanoids


Some of the fatty acids (FAs) are precursors for what have been defined as eicosanoids,
which include: prostaglandins (PGs), thromboxanes (TX), leukotrienes (LT), and the
intermediate hydroperoxyeicosatetraenoic (HPETE) and hydroxyeicosatetraenoic
(HETE) acids. Both prostaglandins (PGs) and leukotrienes (LT) act in a paracrine and
autocrine form on cell function through a family of trans membrane receptors regulating
numerous aspects of cell function. Eicosanoid compounds are also involved in a range of
physiological and pathological processes, including vascular, immunological, and
inflammatory responses.

Arachidonic acid (AA) is the most common substrate in humans for the synthesis of
eicosanoids, which yield two series of prostaglandins, the 1 and 3 series. The synthesis of
prostaglandins begins with the rate-determining hydrolysis of arachidonic acid (AA)
and other 20-carbon poly unsaturated fatty acids (PUFAs) from the sn2-position of the
phospholipids. Metabolites of arachidonic acid (AA) are responsible for many of the
manifestations of inflammation. Arachidonic acid (AA) metabolism is initiated either by
cyclooxygenase, generating prostaglandins (PGs), thromboxanes (TX), and
prostacyclines, or by lipooxygenase to generate leukotrienes (LT) and lipoxins.
109
Figure (69): steps of formation of eicosanoids (www.google.com)

Figure (70): Essential fatty acids as precursors for eicosanoid synthesis


(www.google.com)

5.Cholesterol
Cholesterol is the characteristic steroid alcohol of animal tissues. Cholesterol is a
hydrophobic compound, with a single hydroxyl group located at carbon 3 of a ring to
which a fatty acid can be attached, producing an even more hydrophobic cholesterol
ester. It is absent in plant fat. Cholesterol occurs in both the free and ester form of
cholesterol and fatty acids. Free cholesterol is a component of cell membranes. In
plasma, about one third of cholesterol is free and two thirds exist as esters, containing
linoleic and oleic acids. Intracellularly, the stock pool of cholesterol is formed by its
esters with oleic, palmitic, and linoleic acids in some cells. Cholesterol in the organism
originates both from the external environment by absorption from the digestive tract and
by synthesis de novo from acetyl-CoA.
110
Figure (71): Chemical structure of cholesterol (www.google.com)

Figure (72): Cholesterol ester (www.google.com)

Absorption of cholesterol is a complex process performed by specific carriers on the


brush border of enterocytes in the jejunum.

According to guidelines of National Cholesterol Education Program USA (NCEP), total


cholesterol (TC) concentrations below 200mg/dl, have been regarded as desirable,
whereas, concentrations greater than 240mg/dl, are referred to as hyperlipidemic. Total
cholesterol (TC) should be less than 180mg/dl for children.

Cholesterol modulates membrane fluidity over the range of physiological temperatures.


The structure of the tetracyclic ring of cholesterol contribute to the deceased fluidity of
the cell membrane as the molecule is in a trans conformation making all but the side
chain of cholesterol rigid and planar. In this structural role, cholesterol reduces the
permeability of the plasma membrane to neutral solutes, protons (positive hydrogen
ions), and sodium ions. Within the cell membrane, cholesterol also functions in
intracellular transport, cell signaling, and nerve conduction. Cholesterol is essential for
the structure and function of invaginated caveolae and clathrin-coated pits, including
cavedae-dependent and clathrin-dependent endocytosis. Cholesterol is implicated in
cell signaling processes, assisting in the formation of lipid rafts in the plasma membrane.
Lipid raft formation brings receptor proteins in close proximity with high concentration
of second messenger molecules. Within cells, cholesterol is the precursor molecule in
several biochemical pathways in the liver, cholesterol is converted to bile, which is
stored in the gall bladder. Bile contains bile salts, which solubilize fats in the digestive
tract and aid in the intestinal absorption of fat molecules as well as the fat-soluble
vitamins, A, D, E, and K. Cholesterol is an important precursor molecule for the

111
synthesis of vitamin D and the steroid hormones, including the adrenal gland hormones
(cortisol and aldosterone), as well as the sex hormones (progesterone, estrogen, and
testosterone), and their derivatives. Some research indicates cholesterol may act as
antioxidant.

5.1 Cholesterol in Brain


Cholesterol is a major constituent of the human brain (with about 35gms of cholesterol in
an adult brain), and the brain is the most cholesterol-rich organ, containing about 20% of
the body's total cholesterol. Brain lipids consist of glycerophospholipids, sphingolipids,
and cholesterol in roughly equimolar proportions. Cholesterol is tightly regulated
between the major brain cells-neurons and glia, that is, astrocytes, microglia, and
oligodendrocytes- and is essential for normal brain development. Cholesterol is required
for synapse and dendrite formation, and is for axonal guidance. Cholesterol depletion
leads to synaptic and dendritic spine degeneration, failed neurotransmission, and
decreased synaptic plasticity. Clinically, deficits in cholesterol homeostasis in the central
nervous system (CNS) manifests as severe primary neurological disorders such as
Neimann-Pick C disease and Parkinson's disease. Cholesterol is a pivotal constituent of
cell membranes and for the function of the hedgehog protein.

In contrast to the distribution in plasma lipoproteins essentially all (>99.5%) cholesterol


in the central nervous system (CNS) is present in an unesterified free form. There are two
major pools of central nervous system (CNS) cholesterol: one pool (containing up to
70% of the central nervous system cholesterol) consists of the myelin sheaths of
oligodendroglia, the other pool is made up by plasma membranes of astrocytes and
neurons. The lipid-protein composition of myelin differs from that of other cell
membranes; myelin dry weight consists of about 70% lipids and 30% proteins, in other
cell membranes the distribution is about 30% lipids and about 70% proteins. Major lipid
constituents of myelin are cholesterol, phospholipids, and glycosphingolipids in molar
ratios of about 4:4:2. The lipid composition, in particular the cholesterol composition of
myelin, is believed to have a pivotal role in membrane morphology and function such as
the transmission of nerve impulses. In neurons (which are composed of cell cell body
and axon), electrical impulses are transmitted rapidly along the axon. The axon is
wrapped by myelin made up from the membranes of several oligodendrocytes, separated
by periodic gaps in the myelin sheath-called nodes of Ranvier. This discontinuous
insulation allows the salutatory conduction of the action potential. Due to its reduced
permeability to ions, the cholesterol enrichment of the myelin sheath propagates the
transmission of current along the axon rather than across the membranes of
oligodendrocytes.

5.2 Synthesis of Cholesterol in Peripheral Tissues


Under physiological conditions, de novo synthesis in the whole body exceeds absorption
of cholesterol from food several times. All cells of the human body except nucleus-free
erythrocytes are able to synthesize cholesterol de novo and the product is dedicated for
intracellular use. Cholesterol dedicated for plasma lipoproteins is synthesized in the
liver and in the distal part of the small intestine.

Cholesterol is synthesized by virtually all human tissues, although primarily by liver,


112
intestine, adrenal cortex, and reproductive tissue. All the carbon atoms in cholesterol are
provided by acetate and dihydronicotinamide adenine dinucleotide phosphate
(NADPH) provides the reducing equivalents. The pathway is driven by hydrolysis of the
high-energy thioester bond of acetyl Co enzyme A (CoA) and the terminal phosphate
bond of adenosine triphosphate (ATP). Cholesterol is synthesized in the cytoplasm. The
rate-limiting and regulated step in cholesterol synthesis is catalyzed by the endoplasmic
reticulum membrane protein, hydroxyl methyl glutaryl (HMG) CoA reductase, which
produces mevalonic acid from hydroxyl methyl glutaryl CoA (HMGCoA). The enzyme
is regulated by a number of mechanisms:

1- Expression of the hydroxyl methyl glutaryl (HMG) CoA reductase gene is activated
when cholesterol levels are low, resulting in increased enzyme and, therefore, more
cholesterol synthesis.
2- Hydroxyl methyl glutaryl CoA (HMG CoA) reductase activity is controlled
covalently through the actions of an adenosine monophosphate (AMP)-activated
protein kinase (AMPK), which phosphorylates and inactivates Hydroxyl methyl
glutaryl CoA (HMG CoA) reductase and an insulin-activated protein phosphatases
(which activates Hydroxyl methyl glutaryl CoA (HMG CoA) reductase).
3- Statins are competitive inhibitors of Hydroxyl methyl glutaryl CoA (HMG CoA)
reductase.

Figure (73): Biosynthesis of cholesterol (www.google.com)

Cholesterol synthesis takes place initially in the cytoplasm (up to the hydrocarbon
intermediate squalene), and then in the endoplasmic reticulum (squalene cyclisation and
subsequent steps). Complete biosynthesis of cholesterol amounts to nearly 200
enzymatic processes, and it is not surprising that higher organisms at the top of the food
pyramid make use of cholesterol synthesized by lower organisms, thus saving the load

113
connected with the hugely complicated and energy-demanding cholesterol synthesis.
Thus, many tissues prefer cholesterol from plasma lipoproteins rather than their own
intracellular synthesis. Also from this view point it is logical that cholesterol synthesis,
because of its energy demands, takes place mainly in the night between 2 and 4a.m. in the
period physical rest of the organism. The maximal synthesis of cholesterol in a healthy
human being varies in the range of 500-1000mg a day.

The ring structure of cholesterol cannot be degraded in humans. Cholesterol is


eliminated from the liver as unmodified cholesterol in the bile, or it can be converted to
bile salts that are secreted into the intestinal lumen. It can also serve as a component of
plasma lipoproteins sent to the peripheral tissues.

5.3 Reverse Cholesterol Transport


The transport of cholesterol from extrahepatic tissues to the liver is termed reverse
cholesterol transport. Free (non-esterified) cholesterol effluxes to extracellular acceptors
most notably phospholipid/apoA-1 disks(pre-βHDL). This process is directly or
indirectly through phospholipid efflux is dependent on functional ATP-binding cassette
transporter (ABCA1) receptor. Cholesterol that associates with apoA-1/phospholipid
disks is a substrate for lecithin-cholesterol acyl transferase (LCAT). Lecithin-cholesterol
acyl transferase (LCAT) transfers a fatty acyl chain from phosphatidylcholine to
cholesterol, forming cholesterol ester. The cholesteryl ester partitions into the
hydrophobic core of the lipoprotein, thus forming spherical high-density-lipoprotein
(HDL) particles. These particles can then deliver cholesteryl ester to the liver and
steroidogenic tissues.

Figure (74): Direct and indirect routes of reverse cholesterol transport


(www.google.com)

114
5.4 Bile Acids Synthesis
The major excretory pathway for cholesterol is bile formation. Cholesterol is a substrate
for bile acid formation. There are two major pathways of bile acid synthesis-neutral
(classic) and acidic (alternative). In the neutral or classic pathway, synthesis begins with
hydroxylation of the cholesterol molecule in the 7α position by microsomal cholesterol
7α-hydroxylase (CYP 7α1); in the acidic or alternative pathway, bile acids are
hydroxylated in position 27 by cholesterol 27-hydroxylase (CYP-27). Bile acids achieve
multiple physiological functions: they are mandatory for lipid digestion and absorption
in the intestine; they represent the end product of cholesterol catabolism; they constitute
the most important molecules to drive bile formation and flow, a property otherwise
termed cholesteric activity. Cholesterol itself is also secreted in bile, stored in the gall
bladder and expelled in the intestine upon feeding. Thus, intraluinally there is almost
always cholesterol from bile than from the diet. At present, the only recognized disposal
mechanism of body cholesterol is through biliary excretion. The liver is thereby the main
source of both cholesterol synthesis and disposal.

Figure (75): synthesis of bile acids, cholic and chenodeoxycholic acids


(www.google.com)

5.5 Synthesis of Cholesterol in The Brain


Unlike cholesterol in other organs in the periphery, brain cholesterol is primarily derived
by de novo synthesis. The intact blood- brain barrier (BBB) prevents the uptake of
lipoproteins from the circulation in vertebrates. In cells outside the brain, the need for
cholesterol is covered by uptake of lipoprotein cholesterol by cells as well as by de novo
synthesis. Cholesterol s synthesized via the isoprenoid biosynthetic pathway. Isoprenoid
biosynthesis starts with acetyl-CoA as a substrate, which by means of 6 subsequent
enzyme reactions is converted into isopentenyl-pyrophosphate, the basic C5 isoprene
unit used for synthesis of all subsequent isoprenoids. In total, at least 20 enzymes are
involved for the generation of cholesterol. The 3-hydroxy-3-methyl glutaryl-
coenzymeA reductase is the rate limiting enzyme in cholesterol biosynthesis and the

115
target of statin pharmacotherapy. Besides cholesterol, the cholesterologenic pathway
forms other important intermediates such as mevalonate, farnesyl pyrophosphate,
squalene, and lanosterol. The first enzymes of the isoprenoid/cholesterol biosynthetic
pathway, that is, the conversion of acetyl-CoA to farnesyl pyrophosphate, are localized
in the cytosol except for Hmgcr which, together with most enzymes involved in
cholesterol synthesis, is localized in the endoplasmic reticulum.

Figure (76): Colesterol synthesis and metabolism in the brain (www.google.com)

Figure (77): Isoprenoid pathway for cholesterol synthesis (www.google.com)


116
The majority of brain cholesterol accumulates between the peripheral period and
adolescence when neurons are enriched by specialized plasma membranes termed
myelin. After myelination, the metabolism of cholesterol in the adult brain is
characterized by a very low turnover and minimal losses. However, recent results
indicate that both cholesterol synthesis and degradation are active in the adult brain as
well and that alteration in these mechanisms profoundly influences higher-order brain
functions. In the central nervous system of mammals as humans, >95% of cholesterol is
synthesized de novo from acetate and exchange between plasma lipoprotein cholesterol
and brain cholesterol has only very little impact. However, some studies could
demonstrate the transfer of small amounts of cholesterol from the periphery through the
blood-brain barrier (BBB) to the central nervous system (CNS). Cholesterol as well as
the receptors for cholesterol-containing molecules are pivotal signaling molecules for
brain morphology during embryonic development. The highest rate of cholesterol
synthesis in humans occurs during the first postnatal weeks. This time window
corresponds with the peak of myelination process and the myelination process is delayed
when cholesterol biosynthesis is deficient.

Cholesterol in neurons can be synthesized by neurons themselves and also be taken up


from other cells within the central nervous system (CNS), namely, from
oligodendrocytes. During maturation of neurons, the endogenous synthesis of
cholesterol is impaired and the neurons depend on cholesterol provided by astrocytes.
Brain-derived neutrophic factor (BDNF) is an important stimulus for de novo synthesis
of cholesterol in neurons. In central nervous system (CNS), two different pathways for
exporting cholesterol: cells of the central nervous system (CNS), in particular astrocytes,
shed cholesterol associated with apolipoprotein (apo)E into the cerebrospinal fluid
(CSF). Despite lipoproteins in the brain are secreted predominantly by glia cells,
neurons are also capable of synthesizing lipoproteins under certain conditions. The
second mechanism is the export of cholesterol as 24(S)-hydroxycholesterol. Unlike
nonoxidized cholesterol, oxysterols such as 24(S)-hydroxycholesterol can cross
lipophilic membranes such as blood-brain barrier (BBB) at a much faster rate than
cholesterol itself. Cholesterol is synthesized de novo in brain cells (neurons,
astrocytes,microglial cells). Efflux of central nervous system (CNS) cholesterol through
the blood-brain barrier (BBB) occurs as 24(S)-hydroxycholesterol (24S-OH-C) and 27-
hydroxycholesterol (27-OH-C). 24(S)-hydroxycholesterol (24S-OH-C) is produced
exclusively in the central nervous system (CNS), 27-hydroxycholesterol (27-OH-C) is
produced in most organs. Unlike cholesterol, 24(S)-hydroxycholesterol (24S-OH-C)
and 27-hydroxycholesterol (27-OH-C) can cross the blood-brain barrier (BBB) because
of the hydroxylated side chains. Primarily as astrocytes and microglia secrete high-
density lipoprotein-like (HDL-like lipoproteins) composed of cholesterol and
phospholipids and apo E as the major apoprotein. In plasma, , 24(S)-hydroxycholesterol
(24S-OH-C) and 27-hydroxycholesterol (27-OH-C) are transported on lipoproteins
such as low-density-lipoprotein (LDL) and high-density-lipoprotein (HDL). Apo E is
produced within the central nervous system (CNS) and interacts with amyloid beta (Aβ).
The availability of cholesterol and of apo E are thought to affect amyloidogenesis and
apo E promoting the formation of amyloid fibrils from soluble amyloid beta (Aβ) n the
central nervous system (CNS). De novo cholesterol synthesis in central nervous system
(CNS) cells can be regulated by the apo E-mediated uptake of lipoproteins via the low-
117
density-lipoprotein (LDL) receptor family. Plasma oxysterols are associated with
membranes and in plasma are bound to lipoproteins, similar to other lipids present in
trace amounts such as gangliosides.

While in the brain the oxidation of the steroid side chain at position 24 is the primary
mechanism for the elimination of cholesterol, outside the brain the oxidation occurs at
position 27. Cytochrome P46 (CYP46) responsible for the 24S-hydroxylation of
cholesterol, is localized in neurons indicating that neurons have distinct role in the
excretion of cholesterol from the brain.

6.Triglycerides
Triglycerides are the most abundant of all lipids. It is found abundantly in adipocytes.
These are major components of storage fats in plant and animal cells. Excess calories,
alcohol, and sugar in the body get converted into triglycerides and stored in fat cells
throughout the body. Chemically triglycerides are esters of glycerol with three fatty acid
molecules. Data obtained from National Institute of Health limits triglycerides value to
200mg/dl as the normal range and 500mg/dl as an abnormal range.

Figure (78): Chemical structure of triglycerides (www.google.com)

Figure (79): Triglyceride range (www.google.com)

6.1 Lipogenesis
In presence of either high fat and/or carbohydrate intake, lipogenesis is stimulated and
excess fat is stored as triglycerides (TGs) (also named triacyl glycerols, TAG).
118
Alterations in lipogenesis and lipolysis are both causes and consequences of insulin
resistance, since the imbalance in lipid metabolism is the primary cause of lipotoxicity.
Triglyceride (TG) synthesis is a crucial and strictly regulated process that occurs
principally in the adipose tissue, but also in the liver, muscle, heart, and pancreas. The
process of fatty acid with Acyl-CoA is through the formation of monoacyl glycerol
(MAG) and diacyl glycerol (DAG) by reacting with glycerol-3-phosphate (G3P).
Several hormones control lipogenesis including insulin that stimulates lipid synthesis
and adipogenesis, while glucagon and catecholeamines promote acetyl-CoA
carboxylase (ACC) phosphorylation and inhibit fatty acids (FAs) synthesis. Sources of
glycerol-3-phosphate (G3P) and Acetyl-CoA are plasma glycerol and free fatty acids
(FFAs), but these substrates may also be synthesized de novo.

The first step of free fatty acids (FFAs) esterification is the reaction glycerol-3-phosphate
(G3P). In adipose tissue the main source of glycerol-3-phosphate (G3P) is glucose via
glycolysis, since the activity of glycerokinase (GK), the enzyme that transforms glycerol
into glycerol-3-phosphate (G3P), the uptake of glucose is low. This process is stimulated
by insulin that promotes the uptake of glucose into the cell but also the transformation of
dihydroxyacetone-3P (DHAP) into glycerol-3-phosphate (G3P) by glycerophosphate
dehydrogenase and finally the reaction with free fatty acids (FFAs) to synthesize triacyl
glycerol (TAG). Glycerol-3-phosphate (G3P) can also be synthesized from non-
carbohydrate substrates such as pyruvate, lactate, or amino acids through
glyceroneogenesis that plays a significant role both in adipose tissue and the liver. Since
the liver expresses glycerokinase (GK), it has been thought that during lipogenesis the
main substrate for triglyceride synthesis (TG) was plasma glycerol. During the synthesis
of triacyl glycerol (TAG), the liver utilizes mainly glycerol derived from
glyceroneogenesis (over54%), while the rest of the glycerol derives either from plasma
glycerol (30%) or from plasma glucose through glycolysis (12%). Hepatic
gluconeogenesis and glyceroneogenesis have the synthesis of glyceraldehyde-3P in
common. Free fatty acids (FFAs) and visceral fat accumulation are both associated with
increased gluconeogenesis, and it is likely that glyceroneogenesis is also increased thus
explaining the positive correlation between hepatic and visceral fat.

Triacyl glycerol (TAG) synthesis requires the activation of free fatty acids (FFAs) into
Acyl-CoA by enzyme acyl-CoA synthetase. Free fatty acid-CoA (FFA-CoA) and
glycerol-3 phosphate (G3P) are transformed via acylation, by glycerol-3-phosphate
acyltransferase (GPAT) and acyl CoA acyl glycerol-3-phosphate acyl transferases
(AGPAT), to phosphatidic acid (PA); then, after a dephossphorylation by
phosphohydrolase (PAP2), diacyl glycerols (DAG) are formed. Diacylglycerol
acyltransferase (DGAT) catalyzes the conversion of diacyl glycerol (DAG) into triacyl
glycerol (TAG). In the adipocyte, glycerol-3 phosphate (G3P) might come either from
glycolysis or from non-carbohydrate substrates via the enzyme phosphoenolpyruvate
carboxykinase (PEPCK), through a process named glyceroneogenesis. In the liver
glycerol-3 phosphate (G3P) can also be synthesized from plasma glycerol. De novo
lipogenesis (DNL) occurs in the cytoplasm of various cells (e.g., adipocytes and
hepatocytes) where citric acid is converted to acetyl-CoA by ATP-citrate lyase (ACL)
and subsequently to malonyl-CoA by acetyl-CoA carboxylase(ACC). De novo
lipogenesis (DNL) occurs mainly in the liver, but it might occur in adipose tissue as well,
119
although with low rates. This process requires the two enzymes ATP-citrate lyase
(ACL), acetyl-CoA carboxylase (ACC), and the multi-enzymatic complex fatty acid
synthase (FAS). Glycerol- 3 phosphate (G3P) can be synthesized from non-
carbohydrate substrates such as pyruvate, lactate, or amino acids in oxaloacetate, that is
converted to Glycerol- 3 phosphate (G3P) either directly from phoenolpyruvate (PEP),
via the key enzyme phosphoenol pyruvate carboxykinase (PEPCK), or through
synthesis of dihydroxyacetone (DHA).

Triglycerides (TGs) are synthesized either from circulating free fatty acids (FFAs)
derived from the diet, peripheral lipolysis or de novo lipogenesis (DNL). De novo
lipogenesis (DNL) occurs primarily in the liver and mostly after a high-carbohydrate
meal when only part of the carbohydrates are stored as hepatic glycogen while the excess
is converted to fatty acids and triacylglycerol (TAG). During glycolysis citric acid is
converted to acetyl-CoA, malonyl-CoA, and palmitate, the first fatty acid synthesized.
Other fatty acids are then produced through different mechanisms, e.g., stearic acid by
elongation of palmitic acid, palmitoleic acid and oleic acid by desaturation of palmitic
acid and stearic acid respectively.

Figure (80): De novo lipogenesis (www.google.com)

6.2 Lipolysis
Lipolysis occurs:
a- In adipocytes (where the majority of lipolysis occurs), releasing fatty acids (FAs)
into the circulation.
b- In other cells such as those in liver and muscle to provide fatty acids for local
oxidation
c- In the intravascular space, using circulating lipids as substrate.
120
The vast majority (>95%) of the body's triglyceride (TG) is found in adipose tissue, with
smaller amounts in other tissues. The rate- limiting step for mobilization of adipose
tissue triglyceride (TG) is hydrolysis by hormone-sensitive lipase (HSL) so named
because of its responsiveness to insulin and catecholamines. In the post-absorptive state
most detectable lipolysis is that mediated by adipose tissue hormone-sensitive lipase
(HSL). It is known that fatty acids (FAs) of muscle triglyceride (TG) are derived from
both circulating lipoprotein-triglyceride (lipoprotein-TG) and free fatty acids (FFAs),
and hence at least partially from adipose tissue lipolysis. Lipolysis of circulating
lipoprotein- triglyceride (TG) occurs in the capillary lumen prior to cellular uptake of
fatty acids (FAs). The fatty acids (FAs) transported into the cell as a result of this process
may then either be re-esterified into triglyceride (TG) or immediately oxidized (more
likely in skeletal muscle or myocardium) lipoprotein lipase (LPL), an enzyme bound to
the luminal side of the capillary endothelium and activated by circulating apolipoprotein
C-II and inhibited by apolipoprotein C-III. Lipolysis must be regulated to ensure
adequate supply of lipid fuel for tissues. Starvation and exercise represent the two main
physiological situations of increased lipolysis. It is apparent that lipolysis is controlled
largely by sympathetic activity and insulin concentrations (insulin is the most potent
antilipolytic hormone).

Lipolysis occurs mainly in the adipose tissue. Since fat accumulates mainly in
subcutaneous adipose tissue (SAT), this is the main contributor to plasma free fatty acids
(FFAs). The amount of visceral adipose tissue (VAT) is small compared to total
subcutaneous adipose tissue (SAT), although it may reach more than 38% of the total fat,
thus its contribution to systemic free fatty acids (FFAs) is minimal. Lipolysis involves
the hydrolysis of triacylglycerol (TAG) that results in the release of fatty acids (FAs) and
glycerol into the circulation. Triacylglycerol (TAG) hydrolysis requires different steps
through the action of lipases. The first step, triacylglycerol (TAG) hydrolysis into
diacylglycerol (DAG), is obtained by adipose triglyceride lipase (ATGL) and results in
the release of one fatty acid (FA). Subsequently, diacylglycerols (DAGs) are converted
by the enzyme monoacyl glycerol lipase (MGL) into monoacyl glycerols (MAG) with
the release of one free fatty acid (FFA) or are completely hydrolyzed by hormone-
sensitive lipase (HSL) with the release of two free fatty acids (FFAs) and one glycerol.

Figure (81): Lipolysis (www.google.com)


121
6.3 Beta-Oxidation The Catabolic Pathway for Triacylglycerol
Beta-oxidation (β-oxidation) is the catabolic pathway that occurs in mitochondria and
produces energy from triacylglycerol (TAG) hydrolysis:

1- Free fatty acids (FFAs) are transformed to Acyl-CoA in cytosol.


2- Protein Carnitine Palmitoyl Transferase-1 (CPT1) catalyzes the transfer of the acyl
group of a long-chain fatty acyl-CoA to carnitine to form acylcarnitines (mainly
Palmitoylcarnitine).
3- Carnitine Acyltransferase (CACT) transfers acylcarnitine across outer
mitochondrial membrane.
4- Carnitine Palmitoyl Transferase-2 (CPT2) reconverts acylcarnitine in acyl CoA
and carnitine.
5- Acyl-CoA enters in β-oxidation cycle and is degraded in several Acetyl-CoA
molecules.
6- Acetyl-CoA enters in Krebs cycle to produce energy as adenosine triphosphate
(ATP).

The most important catabolic pathway for triacylglycerol (TAG) and fatty acid (FA)
degradation is β-oxidation that occurs in mitochondria and produces the energy for
homeostasis of cells and tissues. The oxidation of fatty acids occurs in particular during
fasting state and carbohydrate starvation. In liver mitochondria, the acetyl-CoA
produced during β-oxidation is converted to ketone bodies, i.e., acetoacetate, beta-
hydroxybutyrate (BOH), and acetone. Ketone bodies are released and then taken up by
other tissues such as the brain, muscle or heart where they are converted back to acetyl-
CoA to serve as an energy source. Glucose and hormones like insulin, glucagon, and
catecholamines that control lipolysis and lipogenesis modulate substrate availability for
β-oxidation. Accelerated glucose metabolism could inhibit β-oxidation due to increased
production of pyruvate that is transformed to malonyl-CoA, reducing the fatty acid
catabolic pathway. Also excessive free fatty acids (FFAs) can impair mitochondrial
function leading to abnormal fatty acid (FA) oxidation. In this condition, the
mitochondria tend to oxidize more glucose than lipids, even in resting condition. During
a stress condition, when there is an increased energy demand, the mitochondria is unable
to switch from fatty acid (FA) to carbohydrate oxidation. This determines the depletion
of Krebs cycle intermediates and accumulation of acetyl-CoA carboxylase (ACC), thus
contributing to insulin resistance.

7.Lipoproteins
The plasma lipoproteins include chylomicrons, very-low-density lipoproteins (VLDL),
low-density lipoproteins (LDL), and high-density lipoproteins (HDL). They function to
keep lipids (primarily triacylglycerol and cholesteryl esters) soluble as they transport
them between tissues. Lipoproteins are composed of natural lipid core (containing
triacylglycerol, cholesteryl esters, or both) surrounded by a shell of amphipathic
apolipoproteins, phospholipid, and non-esterified cholesterol. To keep triglycerides and
cholesterol esters in a water solution, the surface of lipoproteins consists of molecules
that are apolar toward the lipoprotein core, but are polar toward the water phase.
Apolipoproteins act as cofactors for enzymes in lipid metabolism and as ligands when
lipoproteins are recognized at cell surface. Most important are apolipoproteins B (apo B)
122
and E, which facilitate removal of low-density lipoprotein (LDL) and chylomicron
remnant/ intermediate-density lipoprotein (IDL) from plasma, respectively.
Lipoproteins increase in size and decrease in density from high-density lipoprotein
(HDL) to low-density lipoprotein (LDL) to lipoprotein (a) [Lp (a)] to intermediate-
density lipoprotein (IDL) to very-low-density lipoprotein (VLDL) to chylomicrons.
Chylomicron remnants have size and density like intermediate-density lipoprotein
(IDL) and very-low –density lipoprotein (VLDL).

It is thought that astrocytes are major site of lipoprotein synthesis and assembly in the
brain.

Figure (82): Structure of lipoprotein (www.google.com)

Figure (83): Functions of lipoproteins (www.google.com)


123
Figure (84): Lipoprotein metabolism (www.google.com)

Figure (85): Endogenous and exogenous pathways for lipoproteins metabolism


(www.google.com)

7.1 Chylomicrons
Chylomicrons are assembled in intestinal mucosal cells from dietary lipids (primarily
triacylglycerol) plus additional lipids synthesized in these cells. Chylomicrons are
produced in the gut by enterocytes and contain apolipoprotein B48. Chylomicrons are
the largest particles both in size as well as density, and its concentration is directly
correlated with dietary triglyceride contents. They transport dietary triglycerides and
cholesterol to the liver, muscle, and adipose tissue. Free fatty acids (FFAs) are released
124
from chylomicrons in the capillaries of peripheral tissue after apoprotein (apo) C-II
activates lipoprotein lipase (LPL). Chylomicron remnants rich in cholesterol, are then
formed. Chylomicron remnants are taken up by liver cells. After a 12-14 hours fast,
chylomicrons are absent from the blood stream. Thus individuals who are having a lipid
profile done should fast overnight to ensure that chylomicrons have been cleared.

7.2 Very-Low-Density Lipoproteins


Nascent very-low-density lipoprotein (VLDL) are produced in the liver and are
composed predominantly of triacyl glycerol. The function of very-low-density
lipoprotein (VLDL) is to carry triacylglycerol from the liver to the peripheral tissues
where apolipoprotein lipase degrades the lipid. As triacylglycerol is removed from the
very-low-density lipoprotein (VLDL), the particle receives cholesteryl esters from high-
density lipoprotein. This process is accomplished by cholesteryl ester transfer protein.
Triglycerides (TG) in very-low-density lipoprotein (VLDL) will be degraded in fat and
muscle tissue by the enzyme lipoprotein lipase, and the cholesterol-rich intermediate-
density lipoprotein particle is formed.

7.3 Intermediate-Density Lipoproteins


Very-low-density lipoprotein (VLDL) particles after degradation by lipase enzyme in
the capillaries of adipose tissue and muscle give rise to intermediate-density lipoprotein
(IDL). Some intermediate-density lipoprotein (IDL) particles are cleared by liver cells,
whereas others are converted to low-density lipoprotein (LDL) particles through the
action of the triglyceride-degrading enzyme hepatic lipase.

7.4 Low-Density Lipoproteins


Low-density lipoproteins are synthesized partly in intestinal chyle and partly after
lipolysis of very-low-density lipoproteins (VLDL). Low-density lipoproteins are taken
up via low-density-lipoprotein (LDL) receptor in the liver and other tissues. Low-
density-lipoprotein (LDL) undergo receptor-mediated endocytosis, and their contents
are degraded in the lysosomes.

Figure (86): Receptor-mediated endocytosis for low-density lipoproteins


(www.google.com)
125
High levels of low-density lipoprotein indicate much more cholesterol in the blood
stream than necessary. According to VCEP guidelines, low-density lipoprotein
cholesterol concentrations below 100mg/dl are considered optimal, whereas
concentrations in the range of 160-189mg/dl are considered to be on the high side.
Normal human low-density lipoprotein cholesterol concentration can be as low as 50 to
70 mg/dl.

7.5 High-Density Lipoproteins


High-density lipoproteins (HDL) are created by lipidation of apoA-1 synthesized in the
liver and intestine. High-density lipoprotein (HDL) is commonly referred to as good
cholesterol. High-density lipoproteins (HDL) carry cholesterol and other lipids from
tissues back to the liver for degradation.

Figure (87): High-density-lipoprotein metabolism (www.google.com)

High-density lipoprote

ins have number of functions including:


1- They serve as a circulating reservoir of apo C-II and apo E for chylomicrons and
very-low-density lipoprotein (VLDL).
2- They remove unesterified cholesterol from cell surfaces and other lipoproteins and
esterifying it using phosphatidylcholine-cholesterol acyl transferase, a liver-
synthesized plasma enzyme that is activated by apoA-1.
3- They deliver cholesteryl esters to the liver (reverse cholesterol transport).

High levels of high-density lipoprotein (HDL) cholesterol have been considered as a


good indicator of a healthy heart. The concentrations of 60mg/dl or higher have been
considered as optimal, whereas high-density lipoprotein (HDL) concentrations below
40mg/dl are considered major risk for cardiovascular disease (CVDs). However, high-
density lipoprotein (HDL) is often interpreted in the context of total cholesterol (TC) and
126
low-density lipoprotein (LDL) concentrations and hence may be regarded as less
significant when low-density lipoprotein (LDL) is low.

7.6 Lipoprotein (a)


Lipoprotein (a) [Lp (a) ] is secreted from the liver. Berg defined lipoprotein (a) [Lp(a)] as
a cholesterol-rich plasma lipoprotein, which is directly correlated with atherosclerosis.
Lipoprotein (a) [Lp(a) ] particles are low-density lipoprotein (LDL) particles with an
extra apolipoprotein, apo(a), apo (a) has homology with plasminogen and therefore may
interfere with fibrinolysis and indirectly promote arterial thrombosis.

7.7 Cerebrospinal Fluid Lipoproteins


Essentially all cholesterol within the central nervous system(CNS) is associated with
cell membranes and only tiny amounts are located within the intercellular space and in
cerebrospinal fluid (CSF) under physiological conditions. Cholesterol within the
intercellular space and in cerebrospinal fluid is associated with apos, in particular with
apo E. The role of apo E-containing lipoproteins is postulated to redistribute lipids and to
regulate cholesterol homeostasis within the brain. The concentration of lipoproteins in
the cerebrospinal fluid (CSF) is low. In cerebrospinal fluid (CSF) the concentration of
apo E is about 1%, and of phospholipids about 2% only. Triglycerides (TG) are present in
trace amounts and apo B is absent. Apo E and apo B cannot cross the blood-brain barrier
(BBB).

7.8 Apolipoproteins within The Central Nervous System


Apo E, a 39KD protein, is the major apo in the central nervous system. The functions of
apo E in the central nervous system (CNS) are heterogeneous and range from
participation in cholesterol homeostasis and in nonlipid activities such as protein
chaperoning and signal transduction. Apo E is expressed in the brain in high
concentrations, such that the brain is the organ with the second highest apo E expression
after the liver. Astrocytes are the major source of apo E followed by oligodendrocytes,
microglia, and ependymal layer cells. The stability of apo E in the brain requires the
association with lipids. Apo E is the major apo of cerebrospinal fluid (CSF) Similar to
lipoprotein cholesterol, apo E does not cross the blood-brain barrier (BBB). It has been
suggested that there is a dynamic exchange of apo E among brain cells, that apo E is the
major transport protein for extracellular cholesterol and other lipids, and that apo E-
mediated cholesterol exchange occurs between neuronal and non-neuronal cells of the
central nervous system (CNS).

Apo E-containing high-density lipoprotein (HDL)-like lipoproteins are secreted by


astrocytes and are taken up into neurons via low-density lipoprotein receptors. This
transfer of key lipids and cholesterol facilitates axonal extension and neuronal survival
and requires the presence of sphingomyelin in the apo E.-containing lipoprotein particle.
Apo E does not only play a major role in glia-neuronal lipid metabolism but also acts as a
ligand for multiple receptors in neurons, which interact with a number of downstream
physiological processes. Brain apo E is an important regulator of peripheral energy
homeostasis. Apolipoprotein E has three major isoforms (apoE2, apoE3, and apoE4),
which have varying effects on lipid homeostasis and neuronal function. While both
apoE2 and apoE3 preferentially associate with phospholipid rich HDL-like particles,
apoE4 prefers large triglyceride rich very-low-density lipoprotein (VLDL) particles.
ApoE4 has been implicated in the pathogenesis of Alzheimer's disease (AD).
127
In vivo studies related that apo E retards amyloid β (Aβ) clearance, possibly via an effect
at the blood-brain barrier (BBB). Apo E overexpression might be able to decrease the
progression of amyloid β (Aβ) deposition. The proposed mechanism consists in the
absorption of amyloid β (Aβ) by the CNS-derived lipoproteins followed by the clearance
of these particles from the brain.

Figure (88): Interrelationship among apo E and amyloid β clearance


(www.google.com)

Apolipoprotein J (also known as clusterin) is also present in lipoprotein particles and


regulates cholesterol and brain lipid metabolism. The high expression of clusterin in the
brain and its role in cholesterol trafficking has been established. Of particular interest is
the role of clusterin in the clearance of amyloid beta (Aβ) peptides. Secreted amyloid
beta peptides bind avidly to apoE and to clusterin/apoJ, which prevents the
oligomerization of these amyloid beta peptides. A minor part of this complex is
endocytosed by microglia and astrocytes and degraded by insulin degrading enzyme and
neprilysin but the major part is endocytosed without degradation. The major receptors
for the uptake at the blood-brain barrier are low-density-lipoprotein receptor-related
protein 1 (LRP1) and low-density-lipoprotein receptor-related protein 2 (LRP2). The
third clearance pathway, besides the proteolytic degradation and the endocytic uptake by
microglia and astrocytes, is the efflux of amyloid beta (Aβ) peptides through the blood-
brain barrier (BBB). This efflux is mediated by low-density-lipoprotein receptor-related
protein 1 (LRP1) and low-density-lipoprotein receptor-related protein 2 (LRP2).

7.8.1 Amyloid Beta


Amyloid beta (Aβ) is a cleavage product of amyloid precursor protein (APP). Amyloid
128
beta (Aβ) is an intrinsically disorderd protein, which can self-aggregate and form an
array of supramolecular assemblies with different morphology including oligomers,
amorphous aggregates, and amyloid like fibrils. Cholesterol is involved in several steps
of amyloid beta (Aβ) processing. The elimination products, the oxysterols such as the
brain-specific 24(S)-hydroxycholesterol, as well as 27-hydroxycholesterol increase
soluble amyloid precursor protein (sAPP) production and contribute to amyloidogenesis
with 24(S)-hydroxycholesterol having more pronounced effects than 27-
hydroxycholesterol. Plasma 27-hydroxycholesterol concentration was associated with
soluble amyloid precursor protein (sAPP) levels. Amyloid beta (Aβ) is cleared from the
brain by two pathways. First, through receptor-mediated endocytosis by cells in brain
parenchyma and along the interstitial fluid or through the blood-brain barrier (BBB) and
second, through endopeptidase-mediated proteolytic degradation. The receptor-
mediated endocytosis by members of the LDL-receptor family on astrocytes and
microglia can be an efficient way to reduce brain amyloid beta (Aβ). Most amyloid beta
(Aβ) internalized by apoE receptors is degraded in lysosomes or in transcytosed into the
plasma. However, amyloid beta (Aβ) can escape from degradation or transcytosis,
accumulates in neurons, and exerts its toxicity on the neuronal function. amyloid beta
(Aβ) generation is localized in detergent-resistant membrane fractions (DRMs) [(lipid
rafts) (DRM: special cholesterol-rich membrane subdomains)]. The enzymes
responsible for the enzymatic cleavage of amyloid precursor protein (APP) called β-
secretase, or β-site amyloid precursor protein (APP) cleaving enzyme (βACE), is found
in detergent-resistant membrane fractions (DRMs) in primary hippocampal neurons.

Figure (89): Role of clusterin in amyloid beta clearance (www.google.com)

8.Hypothalamic Fat Sensing


Hypothalamic nutrient sensing may be a key regulator of hepatic lipid homeostasis. The
liver maintains lipid homeostasis through tightly coordinated synthesis and secretion of
triglyceride-rich lipoproteins (VLDL-TG) lipoproteins, and fatty acid (FA) oxidation.
Very-low-density lipoprotein-triglyceride (VLDL-TG) secretion may be a response by
the autonomic nervous system (ANS) to mobilize lipids during a period of relative
nutrient deficiency. In addition, sympathetic denervation prevented the increase in the
very-low-density lipoprotein-triglyceride (VLDL-TG) secretion in the fasted state,

129
whereas total denervation, or parasympathetic denervation did not, suggesting that the
central regulation of lipid mobilization during fasting may be largely mediated through
the sympathetic nervous system. Sympathetic hepatic denervation prevented the
stimulatory effect of neuropeptide-Y (NPY) on very-low-density lipoprotein-
triglyceride (VLDL-TG) secretion. In addition to neuropeptide-Y (NPY), it has been
shown that MC-expressing neurons of the hypothalamus may also regulate hepatic
lipogenesis and triglyceride (TG) metabolism. Central lipid sensing may be implicated
in the regulation of hepatic lipid homeostasis.

Within the hypothalamus, the arcuate nucleus (ARC), ventromedial nucleus of


hypothalamus (VMH), dorsomedial nucleus of hypothalamus (DMH), lateral nucleus of
hypothalamus (LH), and paraventricular nucleus of hypothalamus (PVN) are well
recognized for their involvement in the regulation of a variety of metabolic functions of
the body, and in particular the control of hepatic metabolism. The ventromedial nucleus
of hypothalamus (VMH) is involved in the sympathetic control of the liver, whereas the
lateral nucleus of hypothalamus (LH) plays role in the parasympathetic control of the
liver. The paraventricular nucleus of the hypothalamus was shown as an important
integrative center for the regulation of sympathetic and parasympathetic pathways to the
liver. Hypothalamic action of metabolic signals including insulin, leptin, and fatty acids
(FAs) has been shown to control glucose homeostasis. Hypothalamic nuclei are
heterogeneous containing different types of neurons, which are able to control multiple
organ systems.

Figure (90): hypothalamic lipid sensing in hepatic metabolism (www.google.com)

9.Phospholipids
Phospholipids are polar, ionic compounds composed of an alcohol that is attached by a
phosphodiester bridge to either diacylglycerol or sphingosine. Like fatty acids,
phospholipids are amphipathic in nature, that is, each has a hydrophilic head (the
phosphate group plus whatever alcohol is attached to it, for example, serine,

130
ethanolamine, and choline) and a long hydrophobic tail containing fatty acids or fatty
acid-derived hydrocarbons.

Figure (91): Phospholipid structure (www.google.com)

Phospholipids are the predominant lipids of cell membranes. In membranes, the


hydrophobic portion of a phospholipid molecule is associated with the nonpolar portions
of other membrane constituents, such as glycolipids, proteins, and cholesterol.
Membrane phospholipids also function as a reservoir for intracellular messengers and,
for some proteins, phospholipids serve as anchors to cell membranes.

Phospholipids that contain glycerol are called glycerophospholipids (or


phosphoglycerides). Glycerophospholipids are formed from phosphatidic acid (PA) and
an alcohol.

Figure (92): Glycerophospholipids structure (www.google.com)


131
Sphingophospholipids are a class of lipids containing a backbone of sphingoid bases, a
set of aliphatic aminoalcohols that includes sphingosine. Sphingomyelin is a type of
sphingolipid found in animal cell membranes, especially in the membranous myelin
sheath that surrounds some nerve cell axons. It usually consists of phosphocholine and
ceramide, or a phosphoethanolamine head group; therefore, sphingomyelins can also be
classified as sphingophospholipids.

Figure (93): Sphingolipids structure (www.google.com)

9.1 Synthesis of Glycerophospholipids


Glycerophospholipid synthesis involves either the donation of phosphatidic acid from
CDP-diacylglycerol to an alcohol, or the donation of the phosphomonoester of the
alcohol to 1,2-diacyl glycerol. CDP is the nucleotide cytidine diphosphate.
Phospholipids are synthesized in the smooth endoplasmic reticulum. From there, they
are transported to the Golgi apparatus and then to membranes of organelles or the plasma
membrane, or are secreted from the cell by exocytosis.

Figure (94): Synthesis of glycerophospholipids (www.google.com)

132
9.2 Phosphatidylserine

Figure (95): Phosphatidylserine structure (www.google.com)

Phosphatidylserine (PS) is the major acidic phospholipid in human membranes and


constitutes 2% to 20% of the total phospholipid mass of adult human plasma and
intracellular membranes. Within the healthy human brain, myelin is enriched in and the
phosphatidylserine (PS) content of gray matter doubles from birth to age 80 y.
Throughout the human body, phosphatidylserine (PS) is a structural component of
endoplasmic reticulum (ER), nuclear envelopes, Golgi apparati, inner (cytosolic)
leaflets of plasma membranes, outer mitochondrial membranes, and myelin. About 20%
to 30% of the phosphatidylserine (PS) in human gray matter is in the form of 1-stearoyl-
2-docosahexaenoyl-sn-glycero-3-phosphoserine. The docosahexaenoic acid (DHA)
content of neuronal phosphatidylserine (PS) is of functional importance; in the cortex of
the brain, a reduction in the docosahexaenoic acid (DHA) content of phosphatidylserine
(PS) is associated with the progression of mild cognitive impairment to Alzheimer's
disease (AD). Consequently, the incorporation of phosphatidylserine (PS) into human
membranes is sensitive to the availability of both phosphatidylserine (PS) and
docosahexaenoic acid (DHA). Additionally, fatty-acid recycling at the sn-1 and sn-2
positions of phosphatidylserine (PS) is frequent, rapid and energy-consuming, allowing
co-accumulation of docosahexaenoic acid (DHA) and phosphatidylserine (PS) and
facilitating docosahexaenoic acid (DHA) enrichment of phosphatidylserine (PS)
molecules within membranes.

9.2.1 Synthesis of Phosphatidylserine and Incorporation into Membranes


Most phosphatidylserine (PS) that is synthesized de novo, including that synthesized
within the central nervous system (CNS), results from the PS synthase 1- (PSS1-)
catalyzed substitution of serine for choline on phosphatidylserine (PS) within
mitochondria-associated membrane (MAM) domains of the endoplasmic
reticulum(ER). Some newly synthesized phosphatidylserine (PS) is transported from the
endoplasmic reticulum (ER) to the inner (cytosolic) leaflet of the plasma membrane,
where thermodynamic barriers minimize its movement to the outer (extracellular) leaflet
of the plasma membrane; all healthy human cells exhibit PS-rich cytosolic plasma
membrane leaflets and PS-poor extracellular leaflets. Maintenance of transmembrane
phosphatidylserine (PS) asymmetry is critical to cell survival; active translocation of
phosphatidylserine (PS) to the extracellular leaflet is a required and irreversible signal
for the initiation of phagocytic engulfment of apoptotic cells. To avoid inappropriate
133
engulfment, healthy cells devote up to 4% of all adenosine triphosphate (ATP)
consumption to maintaining transmembrane phosphatidylserine (PS) asymmetry. Most
newly synthesized phosphatidylserine (PS) is actively transported from mitochondria-
associated membrane (MAM) domains of the endoplasmic reticulum (ER) to the outer
leaflet of the mitochondrial inner membrane. Phosphatidylserine-synthesizing
mitochondria-associated membrane (MAM) domains of the endoplasmic reticulum
(ER) tether transiently to the cytosolic leaflet of the mitochondrial outer membrane via
interactions involving mitochondria-associated membrane (MAM) domains, the
mitochondrial outer membrane, and the ER-mitochondria encounter structure
(ERMES), a complex of 5 proteins (Mmm1, Mdm10, Mdm12, Mdm34, and mitofusin2)
that forms a molecular bridge between the endoplasmic reticulum (ER) and
mitochondrion. Movement from the cytosolic leaflet of the outer mitochondrial
membrane to the outer leaflet of the inner mitochondrial membrane requires metabolic
energy, is very rapid, and typically depletes the outer mitochondrial membrane of
phosphatidylserine (PS), generating a requirement for nearly continuous replenishment
from the endoplasmic reticulum (ER). Once within the inner mitochondrial membrane,
phosphatidylserine (PS) is converted rapidly to another major membrane phospholipid,
phosphatidylethanolamine (PE), by PS decarboxylase-1 in a reaction that produces most
of a cell's phosphatidylethanolamine (PE). As intracellular phosphatidylethanolamine
(PE) content reaches a steady-state, a small amount is transported across ER-
mitochondria encounter structure (ERMES) into mitochondria-associated membrane
(MAM) domains of the endoplasmic reticulum (ER) for reconversion into
phosphatidylserine (PS) by PS synthase 2 (PSS2). The expression of PS synthase 2
(PSS2) is greatest in the PS-enriched brain and testes. Oral phosphatidylserine (PS) is
highly bioavailable in humans and readily crosses the blood–brain barrier. The amount
of exogenous phosphatidylserine (PS) that is incorporated into human cell membranes
and is transported from the plasma membrane's outer leaflet to its inner leaflet by a PS-
specific ATP-dependent aminophospholipid translocase (“flippase”) increases as
phosphatidylserine (PS) intake increases. As intracellular phosphatidylserine (PS)
content increases, the activities of PS synthase 1 (PSS1) and PS synthase 2 (PSS2)
decrease, conserving phosphatidylcholine and phosphatidylethanolamine (PE).

Figure (96): Phosphatidylserine synthesis pathway (www.google.com)

134
Figure(97): Phosphatidylserine synthesis and metabolism in the brain
(www.google.com)

9.2.2Phosphatidylserine and Neurotransmission


The incorporation of phosphatidylserine (PS) into neuronal cell membranes influences
the metabolism of the neurotransmitters acetylcholine (ACh), norepinephrine,
serotonin, and dopamine. Adequate amounts of DHA-enriched phosphatidylserine (PS)
are required for the fusion of intraneuronal secretory granules with the presynaptic
membrane, the subsequent release of neurotransmitter molecules into the synaptic cleft
during the intracellular transmission of action potentials and proper postsynaptic
neurotransmitter-receptor interactions. Additionally, exogenous phosphatidylserine
(PS) stimulates electroencephalographic (EEG) evidence of increased cholinergic
neurotransmission in healthy men and women. The neurotransmitter-driven
postsynaptic activation of the signal transducer, calcium/calmodulin-dependent protein
kinase C (PKC), requires an interaction between postsynaptic membrane-associated sn-
1,2-diacylglycerol (DAG)/PKC complexes and postsynaptic membrane-bound
phosphatidylserine (PS). The binding of diacylglycerol (DAG) (originating from the
ACh-triggered catabolism of either phosphatidylinositol or phosphatidylcholine) to the
membrane targeting domain of protein kinase C (PKC) increases the affinity of protein
kinase C (PKC) for the negatively charged serine-rich head groups of postsynaptic
membrane-bound phosphatidylserine (PS) (but not for other phospholipids). The ionic
attraction of these PS-specific clusters of negative charge is required for the attraction of
cytosolic calmodulin-associated Ca2þ ions to protein kinase C (PKC). The formation of
a DAG/Ca2þ ion/PS/PKC complex induces a de-inhibiting conformational change in
the catalytic site of protein kinase C (PKC) that activates the enzyme; subsequent
downstream phosphorylations of intracellular proteins by activated protein kinase C
(PKC) and the biochemical consequences of those phosphorylations “translates” the
presynaptic message into specific responses within the postsynaptic cell.

9.3 Dergradation of Phospholipids


The degradation of phosphoglycerides is performed by phospholipases found in all
tissues and pancreatic juice.

135
Figure (98): Degradation of phospholipids (www.google.com)

Sphingomyelin is degraded to ceramide plus phosphorylcholine by the lysosomal


enzyme sphingomyelinase.

Figure (99): Degradation of sphingomyelin (www.google.com)

10.Glycolipids
Glycolipids are glycosyl derivatives of lipids. They are a part of a large family of
substances known as glycoconjugates. The term glycolipid designates any compound
containing one or more monosaccharide residue bound by a glycosidic linkage to a
hydrophobic moiety such as acylglycerol, a sphingoid, a ceramide (N-acylsphingoid) or
a prenyl phosphate.

136
Figure (100): General structure of glycolipids (www.google.com)

Glycolipids are classified as follows:

1-Glycoglycerolipids
The term glycoglycerolipid is used to designate glycolipids containing mono, di or
trisaccharides linked glycosidically to the hydroxyl group of diglycerides (e.g.
monogalacosyldiglycerides).

Figure (101): General structure of glycoglycerolipids (www.google.com)

2-Glycosphingolipids
The term glycosphingolipid designates lipids containing at least one monosaccharide
residue linked to ceramide moiety. Ceramides are amides of fatty acids with long chain
di or trihydroxy bases. The acyl group of ceramides is generally a long chain saturated or
monounsaturated fatty acids.

Figure (102): Ceramide structure (www.google.com)


137
The glycosphingolipids can be subdivided as follows:
a- Neutral glycosphingolipids which contain one or more glycosyl moieties linked to
ceramides, e.g. cerebrosides.

Figure (103): Glycosphingolipid (www.google.com)

Cerebrosides are monoglycosyl ceramides in which glucose or galactose sugar residue is


attached by O-ester linkage to the primary alcohol of the ceramide. Galactosylceramides
are found in all nervous tissues, but they can amount to 2% of the dry weight of grey
matter and 12% of white matter. Glucosylceramide is found at low levels in animal
tissues, such as spleen and erythrocytes, as well as in nervous tissues.

b- Oligoglycosylceramides contain more than one sugar moiety. They are vital
components of cellular membranes of most eukaryotic organisms and some
bacteria. An example is lactosylceramide (LacCer).
c- Acidic glycosphingolipids are divided into two groups:
1- Sulfoglycosphingolipids are sometimes called sulfatides or
sulfatoglycosphingolipids also. They carry a sulfate ester group attached to the
carbohydrate moiety. They are mainly found in tissues that are very active in
sodium transport such as kidneys, salt glands, and gills.
2- Gangliosides is a group of glycosphingolipids consists of molecules composed of
ceramide linked by a glycosidic bond to an oligosaccharide chain containing
hexose and sialic acid units. These lipids can amount to 6% of the weight of lipids
from brain. One of the common monosialo-gangliosides is ganglioside Gm1.

Figure (104): Structure of GM1 ganglioside (www.google.com)


138
Glycosphingolipids (GSLs) functions fall into two major categories: mediating cell-cell
interactions via binding to complementary molecules on opposing plasma membranes
(trans recognition) and modulating activities of proteins in the same plasma membrane
(cis regulation).Additional roles of glycosphingolipids (GSLs), including ceramide,
sphingosine,sphingosine-1-phosphate, lsoglycolipids, and lysogangliosides, are their
inhibition or activation of apoptosis, proliferation, and stress responses.

Galactosyl ceramide (GalCer), sulfatide, and sphingomyelin are structural to myelin


sheaths and are major lipids in oligodendrocytes and Schwan cells.

10.1 Glycosphingolipid Metabolic Pathways


Glycosphingolipid (GSL) biosynthesis begins with condensation of serine and
palmitoyl-CoA catalyzed by serine-palmitoyl transferase (SPT) on the cytoplasmic face
of the endoplasmic reticulum (ER), leading to de novo biosynthesis of ceramide, the core
of glycosphingolipids (GSLs). Ceramide consists of a fatty acid acyl chain that varies in
length and saturation, and a sphingoid base that differs in the number and position of
double bonds and hydroxyl groups. The fatty acid chain length of ceramide is controlled
by tissue- and cell-specific ceramide synthases (also called longevity assurance genes).
In addition, ceramide can be generated by acid sphingomylinase (aSMase) hydrolysis of
sphingomyelin in the lysosome or at the plasma membrane and by activities of secreted
aSMase se at the plasma membrane or associated with lipoproteins. Neural
sphingomyelinase (nSMase) also cleaves plasma membrane sphingomyelin to
ceramide. In the salvage pathway, lysosomally derived sphingosine can be reacylated.
Once formed, ceramide is sorted to three pathways:

1- Galactosyl ceramide (GalCer) synthesis in the endoplasmic reticulum (ER) that is


followed by 3-sulfo-GalCer (sulfatide) synthesis in the Golgi.
2- Glucosyl ceramide (GlcCer) synthesis on the cytoplasmic face of the Golgi as the
precursor of most glycosphingolipids (GSLs).
3- Ceramide transfer protein (CERT) delivery to the md-Golgi for sphingomyelin
synthesis. In the trans Golgi lumen, the transfer of a β-galactose onto glucosyl
ceramide (GlcCer) by lactosylceramide (LacCer) synthase forms lactosylceramide
(LacCer). Several galactosyl-N-acetylgalactosaminyl-, N-acetylglucosaminyl-,
and sialyl transferases can elongate the oligosaccharide chain of
glycosphingolipids (GSLs) along the luminal side of Golgi, thereby defining the
different series of glycosphingolipids (GSLs). Addition of sialic acid to
lactosylceramide (LacCer) forms GM3 ganglioside by LacCer α-2,3-sialyl
transferase (GM3 synthase) that initiates synthesis of the ganglioside or sialo-GSL
series. The type of sugars in the oligosaccharide backbone depends on the activity
of glycosyl transferases in the Golgi, the cell type, and the developmental or disease
stage.

Ceramide can also be generated through degradation of sphingomyelin in the lysosome


or at the plasma membrane by aSMase and nSMase. Newly synthesized
glycosphingolipids (GSLs) can exit the cell by exocytic vesicles while membrane and
extracellular glycosphingolipids (GSLs) can be transported intracellularly via
endocytosis with subsequent degradation sphingosine and free fatty acids by hydrolases
in the lysosome.
139
The catabolism of complex glycosphingolipids (GSLs) also proceeds by stepwise,
sequential removal of sugars by lysosomal exohydrolases to the final common products,
sphingosine and fatty acids. Individual defects in glycosphingolipids (GSLs) hydrolases
result in excessive accumulation of specific glycosphingolipids (GSLs) in lysosomes
leading to the various lysosomal storage diseases (LSDs). Nonenzymatic proteins are
essential to glycosphingolipids (GSLs) degradation either by presenting lipid substrates
to their cognate enzymes or by interacting with their specific enzymes. Two genes, PSAP
(prosaposin) and GM2A (GM2 activator protein), encode five such proteins. Four
saposins (A, B, C, and D) or sphingolipid activator proteins (Sap) are derived from
proteolytic cleavage of a single precursor protein, prosaposin, in the late endosome and
lysosome. Each of these saposins has specificity for a particular glycosphingolipid
(GSL) hydrolase. The synthesis/degradation pathways of glycoasphingolipid (GSL)
metabolism form a network with the product of one enzyme serving as a substrate for
other enzymes e.g., ceramide formed from sphingoyelin may act directly or serve as a
substrate for ceramidase, for sphingomyelin synthase, or for glycosyl ceramide synthase
(GCS), thereby being converted to sphingosine, sphingomyelin, and a byproduct,
diacylglycerol, or glucosyl ceramide (GlcCer), respectively.

Figure (105): Metabolism of glycosphingolipids (www.google.com)

10.2 Distribution of Glycolipids in Cell


Most of glycolipids are distributed in membranous structures in the cell. Two-thirds of
the total glycolipids are distributed in intracellular membranes such as Golgi apparatus,
endosomes, lysosomes, nuclear membrane, endoplasmic reticulum, and mitochondria.
The tight interactions between cholesterol and glycolipids in the membrane are the
driving force that segregates them from phospholipids that remain fluid in nature.

It was observed that in plasma membranes, glycosphingolipids along with cholesterol


form clusters, called rafts. This region has relatively less phospholipids than other areas

140
of plasma membrane. Approximately 70% of total cellular glycolipids are found in rafts.
Glycosphingolipids (GSLs) cluster in lipid rafts, small lateral microdomains of self-
associating membrane molecules. Besides sphingolipids, lipid rafts are enriched in
cholesterol and selected proteins, including GPI-anchored proteins and some
transmembrane signaling proteins such as receptor tyrosine kinases. On the cytoplasmic
side, acylated proteins, such as Src family protein tyrosine kinases and Gα subunits of G
proteins associate with lipid rafts. Lipid rafts are apparently small (10-50nm in
diameter), each containing perhaps hundreds of lipid molecules along with a few protein
molecules. It has been argued that external clustering of lipid rafts into larger structures
might bring signaling molecules such as kinases and their substrates together to initiate
intracellular signaling. Thus, glycosphingolipids (GSLs) may act as intermediaries in
the flow of information from the outside to the inside of cells.

For Further Readings :


1. Bruce, K.; Zsombok, A.; Eckel, R. (2017). Lipid processing in the brain: a key
regulator of systemic metabolism. Frontiers in Endocrinology, 8.
2. Chalson, S.; Delion-Vancassel, S.; Belzung, C.; Guilloteau, D.; Leguisquet, A.;
Besnard, J.; Durand, G. (1998). Dietary fish oil affects monoaminergic
neurotransmission and behavior in rats. J. Nutr, 128(12):2512-2519.
3. Champe, P.; Harvey, R.; Ferrier, D. (2008). Biochemistry. 4th edition. Lippincott
Williams & Wilkins. Philadelphia.
4. Coppack, S.; Jensen, M.; Miles, J. (1994). In vivo regulation of lipolysis in humans.
Journal of Lipid Research, 35:177-193.
5. Donald, J.;. Voet, V.; Pratt, C. (2008). Lipids, Bilayers and Membranes. Principles
of Biochemistry, Third edition. Wiley. p. 252. ISBN 978-0470-23396-2.
6. Frayar, CD.; Hirsch, R.; Eberhardt, MS.; Yoon, SS.; Wright, JD. (2010).
Hypertension, high serum total cholesterol, and diabetes racial and ethnic
prevalence differences in U.S. adults NCHS Data Brief, 36:1-8.
7. Ginsberg, HN.; Goldberg, IJ. (2001). Disorders of lipoprotein metabolism. In:
Harrison's Principles of Internal Medicine. 15th edition. McGraw Hill. New York.9
8. Glade, M.; Smith, K. (2015). Phosphatidylserine and the human brain. Nutrition,
31:781-786.
9. Hirabayashi, Y. (2012). A world of sphingolipids and glycolipids in the brain-Novel
functions of simple lipids midified with glucose-.Proc. Jpn. Acad. Ser., 88:129-143.
10. Jones, C.; Arai, T.; Rapoport, S. (1997). Evidence for the involvement of
docosahexaenoic acid in cholinergic stimulated signal transduction at the synapse.
Neurochem. Res., 22(6):663-670.
11. Kleess, L.; Janicic, N. (2015). Severe hypertriglyceridemia in pregnancy: a case
report and review of the literature.
12. Landinin, M. (1999). Brain development and assessing the supply of
polyunsaturated fatty acid. Lipids, 34(2):131-137.
13. Malhotra, R. (2012). Membrane glycolipids: functional heterogeneity: a review.
Biochemistry & Analytical Biochemistry, 1(2).
14. Martin, M.; Dotti, C.; Ledesma, M. (2010). Brain cholesterol in normal and
pathological aging. Biochimica et Biophysica Acta, 1801(8):934-944.
15. Martin, R. (1998). Docosahexaenoic acid decreases phospholipase A2 activity in
the neuritis/ nerve growth cones of PC12 cells. J. Neurosci. Res., 15(54):805-813.

141
16. Mogensen, K. (2017). Essential fatty acid deficiency. Nutrient Issues in
Gastroenterology.
17. Nordestgaard, B. (2017). A test in context: lipid profile, fasting versus nonfasting.
JACC, 70(13):1637-1646.
18. Nuutinen, T.; Suuronen, T.; Kauppinen, A.; Salminen, A. (2009). Clusterin: a
forgotten player in Alzheimer's disease. Brain Research Reviews, 61(2):89-104.
19. Onwe, PE.; Folawiyo, MA.; Okike, PI.; Balogun, ME.; Umahi, G.; Besong, EE.;
Okorocha, AE.; Afoke, AO. (2015). Lipid profile and the growing concern on lipid
related diseases. IOSR-JPBS, 10(5):22-27.
20. Orth, M.; Bellosta, S. (2012). Cholesterol: its regulation and role in central nervous
system disorders. Cholesterol, Volume 2012.
21. Saponaro, C.; Gaggini, M.; Carli, F.; Gastaldelli, A. (2015). The subtle balance
between lipolysis and lipogenesis: a critical point in metabolic homeostasis.
Nutrients, 7:9453-9474.
22. Schnaar, R.; Kinoshita, T. (2017). Essentials of glycobiology. 3rd edition. Cold
Spring Harbor. California.
23. Verma, N. (2017). Introduction to hyperlipidemia and its treatment:a review. Int J
Curr Pharm Res, 9(1):6-14.
24. Vyroubal, P.; Chiarla, C.; Giovannini, I.; Hyspler, R.; Ticha, A.; Hrnciarikova, D.;
Zadak, Z. (2008). Hypocholestorolemia in clinically serious conditions-review.
Biomed Pap Med Fac Univ Palacky Olomouc Czech Repub, 152(2):181-189.
25. Xu, YH.; Barnes, S.; Sun, Y.; Grabowski, G. (2010). Multi-system disorders of
glycosphingolipid and ganglioside metabolism. J Lipid Res, 51(7):1643-1675.
26. Yang, W.; Shi, H.; Zhang, J.; Shen, Z.; Zhou, G.; Hu, M. (2017). Effects of the
duration of hyperlipidemia on cerebral lipids, vessels and neurons in rats. Lipids
Health Dis, 16(26).
27. Youdim, K.; Martin, A.; Joseph, J. (2000). Essential fatty acids and the brain:
possible health implications. Int. J. Devl Neuroscience, 18:383-399.
28. Zlokovic, B. (2008). The blood-brain barrier in health and chronic
neurodegenerative disorders. Neuron, 57(2):178-201.

142
CHAPTER FIVE
LIPID DISORDERS AND BRAIN DISEASES

1.Lysosomal Lipid Storage Diseases


Lipid storage diseases, or the lipidoses, are a group of inherited metabolic disorders in
which harmful amounts of fatty materials (lipids) accumulate in various cells and tissues
in the body. People with these disorders either do not produce enough of one of the
enzymes needed to break down (metabolize) lipids or they produce enzymes that do not
work properly. Over time, this excessive storage of fats can cause permanent cellular and
tissue damage, particularly in the brain, peripheral nervous system, liver, spleen, and
bone marrow. Disorders in which intracellular material that cannot be metabolized is
stored in the lysosomes are called lysosomal storage diseases. The lysosome is an
intracellular organelle, often termed the recycling center of the cell. It has an acidic
interior containing hydrolytic enzymes (hydrolases). These hydrolases, together with
the integral transporter proteins (such as NPC1 and NPC2), traffic, break down and
recycle cellular products. A defect in these results in the accumulation of partially
metabolized substrates and a shortage of other lysosomal products. This 'traffic jam'
leads to a complex chain of events, resulting in cell dysfunction and death and the
consequent disease phenotype.

Lipid storage diseases are inherited from one or both parents who carry a defective gene
that regulates a particular lipid-metabolizing enzyme in a class of the body's cells. They
can be inherited two ways:

- Autosomal recessive inheritance occurs when both parents carry and pass on a copy
of the faulty gene, but neither parent is affected by the disorder. Children of either
gender can be affected by an autosomal recessive pattern of inheritance.

- X-linked (or sex-linked) recessive inheritance occurs when the mother carries the
affected gene on the X chromosome.

Sphingolipidoses (singular "sphingolipidosis") are a class of lipid storage disorders


relating to sphingolipid metabolism. The main members of this group are Niemann–Pick
disease, Fabry disease, Krabbe disease, Gaucher disease, Tay–Sachs disease and
metachromatic leukodystrophy. They are generally inherited in an autosomal recessive
fashion, but notably Fabry disease is X-linked recessive.

143
Figure (106): Gaucher disease and related lysosomal storage diseases

Figure (107): Sphingolipidosis (www.google.com)

1.1Gaucher Disease
The most common glycosphingolipid (GSL) storage disease is Gaucher disease (GD),
which is caused by mutations in the enzyme β-glucocerebrosidase (acid 13-glucosidase,
144
or GBA), resulting in the accumulation of glucosylceramide (GlcCer) in the liver and
spleen (and other tissues in more severe cases). Gaucher disease (GD) is an inborn error
of metabolism due to a deficiency of the lysosomal enzyme glucocerebrosidase.

Gaucher disease (GD) is inherited in an autosomal recessive fashion. Genetic studies


have shown that the glucocerebrosidase gene (GBA) is located on chromosome 1q21,
and more than 200 mutations of this gene have been reported as being associated with
Gaucher disease (GD) .

Gaucher disease (GD) is the most common of approximately 40 hereditary lysosomal


storage disorders. The disease is characterized by progressive lysosomal storage of
glucocerebroside (glucosylcerarnide) in macrophages predominantly in bone, bone
marrow, liver, and spleen. Glucocerebrosidase is a 67-kDa glycoprotein 7%
carbohydrate. In Western countries, type 1disease occurs in 1175,000 births. However,
in Jews of Ashkenazi (north-central and eastern European) descent, the only ethnic
group in whom the prevalence is markedly increased, it is seen in approximately 1/600
individuals and the carrier rate is 1/15. In the United States, it is estimated that there are
approximately 12,000 individuals with the disease, and approximately 66% are of
Ashkenazi Jewish ethnicity;' However, due to a highly variable penetrance, the disease
appears to be symptomatic only in approximately 50% of the Ashkenazi Jews who are
genetically affected. Type 2 Gaucher disease (GD) has its onset in infancy so-called
neuronopathic infantile, cerebral, or perinatal lethal Gaucher disease (GD) and has an
estimated prevalence of 1/150,000. Accounting for 1% of all Gaucher diseases, affected
infants have a very short life expectancy of 2-3 years or less due to severe neurological
consequences of the disease. Type 3 Gaucher disease (GD) is more prevalent than type 2
Gaucher disease (GD) because of the considerably longer survival, and accounts for
approximately 5% of all patients with Gaucher disease (GD). Although it is panethnic,
clusters have been well studied in Northern Europe, Egypt, and East Asia.

Table (1): Clinical classification of Gaucher disease


Clinical Classification of Gaucher Disease
Type1:Nonneuropathic Type 2:Acute Neuropathic Type3:Chronic/Subacute
(Adult) (Infantile) Neuropathic (Juvenile)
Whom it strikes Young adults/adults; most common Infants rarely, with no
Children/Young adults, with no in Askenaz:Jewish population ethnicity ethnicity; 1 in
50000 live births (1 in 450)
Distinguishing Liver, spleen, and bone, nervous Early nervous system Later onset of
nervous system symptom system problems problems, brainstem problems:
incoordination, abnormalities mental deterioration, myoclonic seizures
Effects of disease Varies from mild to severe Death in infancy (age<2Y) Slowly
progressive; becomes severe later in childhood Glucocerebrosidase Some activity, but
much less Very little activity Little activity activity than normal

The phenotype has been divided into 3 major types based on the clinical signs/
symptoms. Type 1, the most common type of Gaucher disease (GD), often presents with
abdominal pain and/or enlargement due to hepatosplenomegaly as well as a combination

145
of anemia, leukocytopenia, and thrombocytopenia as the most common clinical
manifestations. Extensive skeletal disease, such as Erlenmeyer flask deformity of the
distal femur, is the typical radiologic finding of Gaucher disease (GD). Pathologic
fracture after falling or minor injury may be an initial presentation in some patients.
Neurologic symptoms (in types 2 and 3), such as convulsions, dementia, ocular muscle
apraxia, mental retardation, and myoclonus, can be seen, as are osteoporosis, hypertonia,
apnea, and yellowish brown skin pigmentation. The demonstrable cardiac, renal, or
pulmonary symptoms are usually absent.

The peripheral blood usually shows a moderate to marked anisopoikilocytosis, mild


microcytosis, hypochromia, and moderate thrombocytopenia. Bone marrow aspirate
and biopsy reveal cellular marrow admixed with individual as well as clusters and
aggregates of large Gaucher cells. The Gaucher cells have abundant granular or fibrillary
blue-gray cytoplasm with a wrinkled tissue paper–like appearance with abundant lightly
periodic acid-Schiff–positive fibrillary material in the cytoplasm. A CD68
immunohistochemical stain usually highlights the Gaucher cells, as does an iron stain.

An enzyme panel demonstrates decreased or absent activity of the characteristic


lysosomal-glucosidase (glucocerebrosidase) in leukocytes, whereas enzyme activity for
other lysosomal enzymes is usually within normal range.

The infantile form (type 2) of Gaucher disease (GD) may lead to early death. Most
affected children die before the age of 5 years. In adult form (type 1) of Gaucher disease
(GD), the clinical features are extremely variable in each patient, and even within a
family various members can exhibit very different clinical problems and course.

Gaucher disease (GD) is caused by a deficiency of lysosomal glucocerebrosidase, a


glycoprotein enzyme necessary for the metabolism of glucosylceramide. This
ubiquitous glycosphingolipid is the basic building block for the complex globosides and
gangliosides that are important components of cell membranes, imbedded receptors, and
lipid rafts. A co-factor protein, saposin C, is an essential activator for
glucocerebrosidase, and along with a second transport protein, lysosomal integral
membrane, is necessary for the successful transfer of glucocerebrosidase (GBA) from
the endoplasmic reticulum to the lysosome.' Rare cases of Gaucher disease (GD), mostly
neuronopathic in presentation, are attributable to hereditary deficiency of saposin C. In
the usual patients with Gaucher disease (GD), deficiency of glucocerebrosidase (GBA)
leads to intracellular accumulation of glucosylceramide as well as other
glycosphingolipids, including the ganglioside GM3, a significant signaling molecule,
and glucosylsphingosine, a potentially toxic lysolipid. Although elevated levels of
glucosylceramide occur in many cell types, lysosomal storage in macrophage lysosomes
is particularly prominent because of the heavy substrate burden inherent in macrophage
phagocytic function. Because macrophages are more commonly distributed in bone
marrow, liver, spleen, and lung than in other body sites, in patients with Gaucher disease
(GD) these organs can accumulate 20-100 times the normal levels of glucosylceramide.
The major manifestations of Gaucher disease (GD) result from the progressive
accumulation of the glycolipid-laden macrophages (Gaucher cells) in the various organs
and tissues noted previously. However, the abnormal macrophages account for only a
146
small fraction of the increased volume of the liver and spleen, in which much of the
increased cell mass is represented by an inflammatory hyperplastic cellular component.
All of the pathogenic pathways leading to organ damage have not yet been elucidated,
but certain findings are notable. Gaucher cells and other immune-modulating cells
secrete lysosomal enzymes and cathepsins. In addition, inflammatory mediators,
including interleukins 6, 8, and 10 and macrophage inflammatory proteins 1 alpha and 1
beta, are increased in the surrounding tissue and in the peripheral blood. Gaucher cells
themselves, at some point in their evolution, phenotypically correspond most closely to
"alternately activated" macrophages, which are associated with chronic inflammation,
healing processes, and fibrosia. In patients with neuronopathic types of Gaucher disease
(GD), the deacylated form of glucosylceramide, namely glucosylsphingosine, is
elevated in neuronal cells and may play a key role as a neurotoxic agent by deregulating
calcium homeostasis. With respect to bone disease, both mineralized bone and bone
marrow are affected. Osteonecrosis, medullary infarction, osteoporosis, and defective
bone remodeling are common in untreated Gaucher disease (GD) patients, and the bone
marrow is infiltrated with Gaucher disease (GD) cells, leading to osteosclerosis and
fibrosis. Patients often do not achieve peak bone mass, which results in subsequent
osteopenia, osteoporosis, and increased fracture risk. Children may have growth
retardation and delayed puberty. Aseptic necrosis of the shoulders and hips are relatively
common leading to subsequent joint deformities and functional disability that may not
always be correctible with joint replacement surgery. Gaucher disease (GD) patients
may experience severe acute bone pain episodes due to infarction (bone crises) as well as
intractable chronic pain. Although thrombocytopenia and anemia are in large part a
function of splenomegaly and hypersplenism, bone marrow infiltration itself appears to
contribute to impaired hematopoiesis, a process that may become irreversible once
fibrosis and osteosclerosis develop. There may also be an intrinsic defect in the bone
marrow that can impact cellular maturation. Bleeding in Gaucher disease (GD) is most
directly related to thrombocytopenia, but patients may have platelet defects and other
coagulation factor deficiencies, including coincidental factor XI deficiency in Ashkenazi
Jews in whom this coagulopathy is relatively common. Type 1 Gaucher disease (GD) is
present in more than 90% of all known patients with Gaucher disease (GD). Despite the
increased prevalence among Ashkenazi Jews, because they are such a small minority
population worldwide, most type 1 Gaucher disease (GD) patients are in fact non-
jewish." Types 2 and 3 are panethnic and can be distinguished from type 1 by the early
onset of neurological involvement. From the ICGG Gaucher Registry data, the incidence
of clinical characteristics for all Gaucher disease (GD) types include the following:
splenomegaly 85%, hepatomegaly 63%, anemia 34%, thrombocytopenia (with or
without bleeding manifestations) 68%, osteopenia and growth retardation 36%.44. In
type 2 disease (acute neuropathic), which accounts for 1% of cases, in addition to the
findings noted in type 1 disease, the following clinical characteristics have been
reported: developmental delay, hydrops fetalis, congenital ichthyosis, strabismus,
developmental delay, supranuclear gaze palsy, bulbar palsy, paresis, hypertonia, rigidity,
opisthotonus, dysphagia, and seizures. The pathology shows neuronal loss, gliosis,
periadventitial Gaucher cells, neuronophagia, with involvement of the frontal cortex,
thalamus, caudate, globus pallidus, pons, and cerebellum. This disease variant, when not
associated with neonatal death, is rapidly progressive over a period of 2-3 years and is
invariably fatal despite any treatment known to date. In type 3 disease, the following has
147
been observed: developmental delay, strabismus and supranuclear gaze palsy,
progressive dementia, myoclonus, corneal opacities, and cardiovascular calcification.
The latter is associated with a unique phenotype found in Arab, Iberian, and Japanese
patients associated with the D409H mutation. Another unique and as yet unexplained
finding in some type 3 patients (Norrbottnian variant) is the development of a severe
gibbus deformity that is not associated with vertebral skeletal involvement.

Figure (108): Macrophage cell in Gaucher disease (www.google.com)

1.2 Niemann-Pick Disease


Niemann-Pick disease is a heterogeneous group of autosomal recessive disorders. Type
A and type B result from deficiency of the enzyme sphingomyelinase, whereas type C
results from mutations in the NPC1 or NPC2 gene, which appears to be involved in
cholesterol trafficking and resulting in accumulation of cholesterol as well as
sphingomyelin. Type A is a lethal infantile form with marked progressive neurologic
involvement. Type B is a later-onset form with no neurologic involvement but
hepatosplenomegaly in many patients. Patients with type C disease manifest progressive
neurologic involvement and hepatosplenomegaly, but may survive into adulthood. The
marrow of these patients contains typical foam cells with small droplets in the cytoplasm
and sea-blue histiocytes. Niemann-Pick disease is caused by a deficiency in
sphingomyelinase, which leads to accumulation of sphingomyelin in the cytoplasm of
macrophages.

Niemann–Pick type C (NP-C) is rare, with a 'classical' clinical incidence of


approximately 1:100 000. It follows an autosomal inheritance pattern, the result of
mutations in one of two genes: NPC1 (chromosome 18q 11-12) or NPC2 (chromosome
14q 24.3), with NPC1 accounting for 95% of cases. However, unusually, neither of these
genes encodes an enzyme; they encode intracellular transporter proteins: NPC1, a late
endosome/lysosomal transmembrane-bound protein and NPC2, a soluble protein.

148
NPC1, a large transmembrane protein of the late endosome/lysosome and NPC2, a
soluble lysosomal protein, work cooperatively to traffic intracellular lipids. Loss of
function in either protein leads to the accumulation of cholesterol and a range of
sphingolipids in the late endosomal/lysosomal intracellular compartment. This disrupts
lysosomal calcium homeostasis, resulting in a host of secondary cellular trafficking
defects. The neuropathological sequelae of these defects include Alzheimer's-like
neurofibrillary tangles, neuronal degeneration, neuroaxonal dystrophy and
demyelination. Also, as endogenously synthesized cholesterol is necessary for axonal
membrane maintenance and repair, white matter tracts are severely affected, with the
corpus callosum showing the most striking axonal loss. Purkinje cells of the cerebellum,
basal ganglia and thalamus are characteristically vulnerable in Niemann–Pick type C
(NP-C), leading to the often pronounced cerebellar dysfunction and ataxia in
Niemann–Pick type C (NP-C) patients.

Niemann–Pick type C (NP-C) can vary widely in both age at onset and symptoms. A
useful classification system subdivides Niemann–Pick type C (NP-C) into four groups
based on the onset of neurological disease:

Ÿ early infantile
Ÿ late infantile
Ÿ juvenile
Ÿ adolescent/adult onset.

Typically, the earlier the onset of neurological disease, the more aggressive the disease
process.

The neurodegenerative disease leads to dementia in almost all Niemann–Pick type C


(NP-C) patients. Niemann–Pick type C (NP-C) is sometimes referred to as 'childhood
Alzheimer's'. Although this is used as an easy identifier rather than for its accuracy, the
two diseases share some neuropathological features.

In patients with adult/adolescent onset, cognitive decline features to a greater or lesser


extent in almost all cases. If found in combination with other disease features, further
investigations should be performed. To highlight a possible diagnosis, the three most
useful clinical features are: cognitive decline, psychosis and progressive ataxia.

The cognitive profile in adult patients with Niemann–Pick type C (NP-C) usually starts
with problems in word fluidity, working memory and executive dysfunction. There may
also be a frontal lobe syndrome with perseveration and loss of interpersonal distance that
manifests as excessive familiarity. Psychiatric symptoms associated with Niemann–
Pick type C (NP-C) can vary. In juvenile- and adolescent-onset patients, intellectual
disability, behavioral problems and attention-deficit hyperactivity disorder (ADHD)
have been reported.

This psychiatric disease may initially be indistinguishable from schizophrenia, with


auditory hallucinations, delusions and disorders of thought and behavior; however,
certain features are suggestive of an organic cause:

149
Ÿ neurological or visceral features
Ÿ cognitive impairment
Ÿ treatment resistance or even a paradoxical worsening of psychosis with drug therapy
Ÿ visual hallucinations, unusual in classical forms of schizophrenia.

Other major psychiatric illnesses described in Niemann–Pick type C (NP-C) include:

depression generally susceptible to selective serotonin reuptake inhibitor (SSRI)


therapy

Ÿ bipolar disorder, often sensitive to mood stabilizers such as sodium valproate


Ÿ obsessive–compulsive behavior
Ÿ catatonia, often in younger patients and sometimes resistant to treatment, although
electroconvulsive therapy (ECT) has been used successfully.

Adolescent- and adult-onset Niemann–Pick type C (NP-C) patients almost always have
some neurological features at presentation, although these may at first be subtle and
eclipsed by psychiatric features. In the more aggressive late infantile/juvenile-onset
group, patients are often first described as being clumsy and struggling at school. This
then progresses to the development of frank neurological disease that may include limb
and gait ataxia, seizures, gelastic cataplexy (the loss of muscle tone with emotional
stimuli), dysarthria, dystonia, dysphagia and dementia. Prognosis in these patients is
poor, with death from the consequences of their advanced neurological disease typically
in their late teenage years or early adulthood. Adolescent and adult patients share some
of these disease features, but in their case the illness is more insidious in its onset and
slower in progression. Cerebellar dysfunction, especially ataxia, is the most commonly
identified neurological feature, although dysarthria and dystonia are also frequently
present. Interestingly, epilepsy, common in infantile and juvenile disease, and cataplexy
(20% of classical Niemann–Pick type C (NP-C) patients), are both rarely seen.

The most important neurological sign in Niemann–Pick type C (NP-C), as it is both


highly prevalent and specific, is a vertical supranuclear gaze palsy (VSGP). Vertical
supranuclear gaze palsy (VSGP) is seen in only a limited number of other
neurodegenerative diseases and rarely so early in their disease process. In Niemann–
Pick type C (NP-C) it nearly always heralds the onset of the neuropsychiatric disease,
regardless of the patient's age.

The gaze palsy, initially in the vertical plane, progresses to also involve horizontal eye
movements as the brainstem pathology advances. Initially, the vertical supranuclear
gaze palsy (VSGP) is subtle and may be missed. It involves vertical voluntary saccadic
movements only, especially of downward gaze, and at this stage slow pursuit eye
movements are preserved. If saccadic eye movements are not tested, the initial vertical
supranuclear gaze palsy (VSGP) will be missed.

Niemann–Pick type C (NP-C) is a neurovisceral disease, but in adolescents and adults


the visceral component is rarely of clinical significance, although splenomegaly with or
without hepatomegaly is usually present.

150
In the perinatal and early juvenile forms, systemic manifestations may be pronounced,
with severe and sometimes fatal liver and pulmonary disease. Interestingly, regardless of
the patient's age, visceral disease, when present, always precedes neuropsychiatric
features, often by years or even decades.

In adolescent- and adult-onset patients, hepatosplenomegaly – although frequently


present – is often unrecognized. A patient with splenomegaly (especially in the absence
of liver disease) with a co-existent neurodegenerative or psychiatric disorder is strongly
suggestive of Niemann–Pick type C (NP-C) and should be appropriately investigated. A
history of paediatric liver disease in such patients should also raise clinical suspicion.

More recently, highly specific and sensitive oxidative cholesterol metabolites for
Niemann–Pick type C (NP-C) have been identified. This 'oxysterol test' can be
performed on a plasma sample and is now used as the first-line diagnostic test with
subsequent genetic confirmation at one of the principal United Kingdom reference
laboratories for lysosomal storage disorders.

Niemann–Pick disease type C (NP-C) clinical manifestations include vertical gaze


palsy, ataxia, dystonia, dementia, cognitive impairment, and seizures with
hepatosplenomegaly in early childhood; progressive neurological function defects are
considered the cause of death that often occurs in teenage years. Pathologic features in
Niemann–Pick type C (NP-C) brain include neuronal loss, especially in the cerebellum,
axonal spheroids, meganeurite formation, and ectopic neurites. The hallmark of
Niemann–Pick type C (NP-C) at the cellular level is accumulation of cholesterol and
other lipids in late endosome/lysosomes.

Niemann–Pick type C (NP-C) and Alzheimer's disease exhibit several similarities,


including endosomal/lysosomal abnormalities, cholesterol imbalance, neurofibrillary
tangle formation, deregulation of the phosphatidylinositol-3 kinase signaling cascade,
and glial-mediated neuroinflammation. In addition, amyloid-beta peptide deposition
was also evident in brains of Niemann–Pick type C (NP-C) patients with ApoE epsilon4
homozygosity.

Recently, it has been reported that brains of some Niemann–Pick type C (NP-C) patients
also contain aberrant alpha-synuclein accumulation and Lewy bodies, which inspires the
proposal to include Niemann–Pick type C (NP-C) as a subclass of Lewy body diseases.

Cholesterol homeostasis is critical for normal function of the central nervous system
(CNS), which is particularly rich in cholesterol. Although the human brain comprises
only 2% of the body mass, it contains about 25% of the total body unesterified
cholesterol.

In contrast to other peripheral tissues that obtain cholesterol from both de novo synthesis
within the cells and uptake of cholesterol-containing lipoprotein particles from serum,
nearly all cholesterol supply in the central nervous system (CNS) comes from in situ
synthesis. Previous studies have shown that during early development neurons rely
heavily on de novo cholesterol synthesis, whereas uptake of exogenous cholesterol

151
provided by glia may be critical for mature neurons later. Dysfunction of either de novo
synthesis or uptake of exogenous cholesterol can lead to disruption of cholesterol
homeostasis in neurons.

Cholesterol esterification impairment in Niemann–Pick type C (NP-C) disease was first


revealed in fibroblasts cultured from Niemann–Pick type C (NP-C) patients, which
distinguished NPC disease from other lysosomal storage diseases. Subsequent research
found that not only unesterified cholesterol, but also gangliosides GM2 and GM3, and
bis-monoacylglycerol phosphate accumulated in late endosomes/lysosomes. The
discovery that lipids other than cholesterol also accumulated in late endosomes/
lysosomes has led to the debate over whether aberrant trafficking of cholesterol or of
other lipids is the primary cause of the Niemann–Pick type C (NP-C) phenotype. It was
suggested that cholesterol accumulation was ganglioside-dependent since depletion of
the ganglioside-related enzyme GM2/GD2 synthase in NPC-deficient neurons
diminished cholesterol accumulation.

Brain cholesterol homeostasis is achieved through different mechanisms from those in


other organs. In vivo, direct measurement of the uptake of low density lipoproteins
(LDL) in different brain regions has indicated that cholesterol carried in low density
lipoproteins (LDL) circulating in serum plays little or no role in the process of sterol
acquisition during brain development or in cholesterol turnover in the mature central
nervous system. In contrast, lipoproteins in brain transport exogenous cholesterol
generated in glia to neurons. Several members of the low density lipoproteins (LDL)
receptor family, including apolipoprotein E, A1, D, and J, are expressed in brains with
apolipoprotein E and apolipoprotein J being the major apolipoproteins in central nervous
system (CNS).

Recent studies have indicated that disruption in cholesterol homeostasis plays an


important role in several neurodegenerative diseases, including Alzheimer's disease and
Niemann–Pick type C (NP-C) disease. In Niemann–Pick type C (NP-C), cholesterol
accumulation occurs early and is closely associated with neurodegeneration. Although
loss-of-function mutations in NPC1 and NPC2 genes have been identified as the genetic
cause of this disorder, the precise mechanism by which NPC deficit leads to neuronal
death remains elusive. Recent research has led to a better understanding of the roles of
NPC1 and NPC2 in cholesterol flux through late endosomes/lysosomes, which may
reveal new therapeutic strategies. Other pathological features such as
neuroinflammation and autophagy are also linked to the development of the disease.

Niemann–Pick type C (NP-C) is characterized by the accumulation of unesterified


cholesterol and other lipids in endosomal/lysosomal compartments that stem from
inherited deficiencies of the NPC1 or NPC2 proteins. NPC1 is a transmembrane protein
and NPC2 is a soluble protein. Both are involved in intralysosomal cholesterol
trafficking. Recently, NPC2 was proposed to act as a donor of cholesterol to NPC1 that
facilitates cholesterol transport across the lysosomal membrane. Loss-of-function
mutations in the NPC1 gene also lead to failure of the calcium-mediated fusion of
endosomes with lysosomes, resulting in the accumulation of Glucosylceramide
(GlcCer) , lactosylceramide (LacCer), and GM2 in late endosomes and lysosomes. The
152
exact mechanism of disease causality by either cholesterol (primarily) or
glycosphingolipids (GSLs) (secondary or primary) accumulation has not been resolved.
Elevation of sphingosine was found before cholesterol and glycosphingolipid (GSL)
accumulation. The importance of sphingosine in NPC1 has been addressed recently).

Although NPC1 gene is expressed in all tissues, the nervous system manifestations of the
disease are predominant and lethal. The reason why neurons are most vulnerable to
NPC1 deficiency remains unknown. Apoptosis was found in cortical neurons treated
with a blocker of cholesterol transport, U18666A, in liver cells of Npc1−/− mice, and in
brains of NPC patients and Npc1−/− mice. However, additional results support the
notion that another type of programmed cell death, autophagic cell death, plays a critical
role in neuronal death in Niemann–Pick type C (NP-C) disease.

Autophagy or “self-eating” is an adaptation process conserved in cells from yeasts to


mice and humans. As a house-keeping mechanism, autophagy engulfs fragments of
damaged organelles and long-lived membrane proteins and transfers packaged cargos to
lysosomes for degradation. Recent evidence indicates that autophagy is associated with
neurodegeneration in Alzheimer's disease (AD), Parkinson's disease, and Huntington
disease. Research from our laboratory and others have also shown that autophagy
activity is increased in Npc1−/− mice. Levels of LC3 (microtubule-associated protein 1
light chain 3 protein)-II, a marker of autophagic activation, are increased in brain of
Npc1−/− mice and in fibroblasts with NPC1 deficiency. LC3-immunopositive granules
were also labelled with filipin-stained cholesterol, suggesting that autophagy in
Niemann–Pick type C (NP-C) is closely associated with cholesterol accumulation. This
notion is further supported by our recent finding that suppression of autophagy by
treatment of mice with allopregnanolone, a neurosteroid that is deficient in brain of
Npc1−/− mice, was associated with reduction in cholesterol accumulation.
Ultrastructural analysis with electron microscopy revealed the existence of classic
double membrane vacuole-like structures in 6-weeks old Npc1−/− mice . These results
suggest an increase in autophagosomes in Niemann–Pick type C (NP-C). However, as
the volume of autophagosomes depends on the dynamics of influx and efflux, whether
this increase represents a net increase in autophagic activity or a efflux jam because of
lysosomal dysfunction remains an open question.

The mechanism by which autophagic activity is elevated is largely unknown. It is


generally agreed that amino acid starvation induces autophagic activity; whether
lipid/cholesterol starvation also results in enhanced autophagic activity is not as certain.
Depletion of cholesterol in human fibroblasts, by either acute chemical treatment or
metabolic suppression of cholesterol synthesis, increased levels of LC3-II and LC3-II-
immunopositive granules suggesting an increase in autophagic activity. Electron
microscopy examination revealed that autophagic vacuoles induced by cholesterol
depletion were indistinguishable from that induced by amino acid starvation, which
further supports the idea that cholesterol starvation can also initiate autophagy. More
convincing evidence suggesting an increase in autophagic induction in Niemann–Pick
type C (NP-C) was obtained where recently reported that cholesterol depletion by
U18666A inhibited the formation of filipin-labeled LC3-immunopositive granules but
promoted the formation of ring-shaped filipin-negative LC3-immunopositive
153
structures. However, the molecular basis for cholesterol depletion-induced autophagy
remains elusive. Blocking intracellular cholesterol trafficking by U18666A in wild-type
fibroblasts increased the expression of LC3 and the conversion of LC3-I to LC3-II, a
process that was dependent on the Beclin-1 rather than the mTOR (mammalian target of
rapamycin) signalling pathway, which may imply that cholesterol depletion-induced
autophagy uses different molecular mechanisms from those induced by amino acid
starvation. It is clear that an increase in autophagosomes, possibly by enhanced initiation
rather than decreased efflux, is associated with cholesterol accumulation in NPC disease.

Figure (109): Generation of Niemann Pick type C syndrome (www.google.com)

Figure (110): Impaired autophagy in Niemann Pick type C disease


(www.google.com)
154
Figure (111): Progression of Niemann-Pick type C disease in mice and humans
(www.google.com)

1.3 Fabry Disease


Fabry disease, Wolman/cholesteryl ester storage disease (CESD), and GM 1 -
gangliosidoses are other lipid storage diseases characterized by hepatosplenomegaly;
GM2-gangliosidosis by hepatomegaly only. Wolman/cholesteryl ester storage disease
(CESD) patients may result in anemia and have sea-blue histiocytes.

Fabry disease, also known as alpha-galactosidase-A deficiency, causes a buildup of fatty


material in the autonomic nervous system (the part of the nervous system that controls
involuntary functions such as breathing and heart beat), eyes, kidneys, and
cardiovascular system. Fabry disease is the only X-linked lipid storage disease. Males
are primarily affected, although a milder and more variable form is common in females.
Occasionally, affected females have severe manifestations similar to those seen in males
with the disorder. Onset of symptoms is usually during childhood or adolescence.
Neurological signs include burning pain in the arms and legs, which worsens in hot
weather or following exercise, and the buildup of excess material in the clear layers of
the cornea (resulting in clouding but no change in vision). Fatty storage in blood vessel
walls may impair circulation, putting the person at risk for stroke or heart attack. Other
symptoms include heart enlargement, progressive kidney impairment leading to renal
failure, gastrointestinal difficulties, decreased sweating, and fever. Angiokeratomas
(small, non-cancerous, reddish-purple elevated spots on the skin) may develop on the
lower part of the trunk of the body and become more numerous with age. People with
Fabry disease often die prematurely of complications from heart disease, renal failure, or
stroke.

155
Fabry disease (OMIM 301500), an X-linked genetic condition caused by alpha-
galactosidase (EC 3.2.1.22) deficiency, is associated with increased accumulation of
glycosphingolipids in cardiovascular tissues and leads to organ failure and premature
death. Affected male patients display clinical features of the disease but female carriers
manifest with symptoms later in their life. The clinical manifestations consist of
vasculature associated complications, but the pathophysiology is unclear. It was shown
that the Fabry disease specific vascular lesions occur as a result of vascular dysfunction
with major components being endothelial dysfunction, alterations in cerebral perfusion
and athero-thrombogenesis. Although some patients with Fabry disease may suffer from
stroke by involvement of larger arteries, small-vessel disease causes cerebral
complications and probably contributes to complications of the kidney and the heart.

Undoubtedly, other cardiovascular risk factors contribute to enhanced worsening of


arterial performance. Hypercholesterolaemia with a markedly raised high-density-
lipoprotein-cholesterol (HDL-C) was observed in patients with Fabry disease and was
previously associated with the occurrence of cardiovascular disease. Atherosclerosis
was previously described in several case studies.

The pathophysiology of lipoproteins in Fabry disease is thought to be affected by cellular


glycosphingolipids accumulation and subsequent inhibition of apoA-I-mediated
cholesterol efflux. The cellular defect in cholesterol trafficking results from substrate
accumulation and thereby suggests that the vascular disease observed in Fabry patients
is caused by an alteration in the lipid homeostasis of the endothelial cells, accumulation
of substrate in the endothelial cells and upregulation of proinflammatory endothelial
markers. As a consequence, the arteriosclerosis in patients with Fabry disease is
probably due to damage to vessel walls occurring as a result of defective
glycosphingolipids metabolism and accumulation of glycosphingolipids in the tissue,
rather than to abnormal low density lipoprotein (LDL) metabolism.

Additionally, low density lipoprotein (LDL) and high density lipoproteins (HDL)
transport glycosphingolipids from circulation to vascular cells through the low-density
lipoprotein receptor that in turn leads to glycosphingolipids accumulation. The defect in
glycosphingolipids metabolism caused by the enzyme deficiency present in Fabry
disease causes an even distribution of excessive plasma glycosphingolipids among
several lipoproteins.

Patients with Fabry disease have raised high-density-lipoprotein-cholesterol (HDL-C)


and, as a result, raised total cholesterol irrespective of the treatment option. The
incidence of cardiovascular disease among these patients is known to be increased but it
remains unclear whether it is related to the accumulation of glycosphingolipids or
cholesterol in the arteries or both.

156
Figure (112): Enzymatic defect in Fabry disease (www.google.com)

Figure (113): Clinical genomics of Fabry disease (www.google.com)

1.4 Krabbe Disease


Krabbe disease (also known as globoid cell leukodystrophy and galactosylceramide
lipidosis) is an autosomal recessive disorder caused by deficiency of the enzyme
galactocerebrosidase. The disease most often affects infants, with onset before age 6
months, but can occur in adolescence or adulthood. The buildup of undigested fats
affects the growth of the nerve's protective insulating sheath (myelin sheath) and causes
severe deterioration of mental and motor skills. Other symptoms include muscle
weakness, reduced ability of a muscle to stretch (hypertonia), muscle stiffening
(spasticity), sudden shock-like or jerking of the limbs (myoclonic seizures), irritability,
unexplained fever, deafness, blindness, paralysis, and difficulty when swallowing.
Prolonged weight loss may also occur. The disease may be diagnosed by enzyme testing
and by identification of its characteristic grouping of cells into globoid bodies in the
white matter of the brain, demyelination of nerves and degeneration, and destruction of
brain cells. In infants, the disease is generally fatal before age 2. Individuals with a later
onset form of the disease have a milder course of the disease and live significantly longer.
No specific treatment for Krabbe disease has been developed, although early bone
marrow transplantation may help some people.

157
Krabbe disease also called globoid cell leukodystrophy (GLD, OMIM #245200) is an
autosomal recessive lysosomal storage disease resulting from a deficiency of the
lysosomal enzyme galactocerebrosidase (galactosylceramidase, GALC). The
deficiency of galactosylceramidase (GALC) impairs the degradation of a major myelin
lipid, galactocerebroside and that of a parent cytotoxic compound,
galactosylsphingosine also called psychosine. The excess of galactosylceramide elicits
the formation of multinucleated macrophages, the globoid cells. Progressive
accumulation of psychosine has been well established in infantile Krabbe disease and in
mouse, dog and monkey animal models; it can explain the prominent death of
oligodendrocytes and myelination arrest, and contributes to progressive demyelination.
Four different forms of Krabbe disease are usually distinguished based on age at onset of
neurological symptoms:

1- the early infantile form starts between 3 and 6 months and is characterized by
hyperirritability, stagnation of development, blindness, hypertonicity, and
decerebrate rigidity. Death occurs between the 1st and the 3rd years of age;
2- the late infantile form (onset between 7 months- 12 months) and
3- the juvenile form (onset between 1 and 10 years) are characterized by spastic
tetraparesis, cerebellar ataxia, optic atrophy, mental retardation and cognitive
decline. 4-the adult form has been rarely reported. It is a more insidious disease with
heterogeneous phenotypes, mostly described as isolated case reports. Here, from a
series of 11 patients diagnosed in French hospitals and 30 cases previously reported
in literature, there was attempt to describe the clinical, radiological,
electrophysiological, biochemical and genetic features of adult Krabbe disease. In
contrast to what can be observed in the infantile form, cerebellar white matter and
deep gray matter changes were only observed in patient #16 who had a childhood
onset. Only 59 % of the adult patients displayed signs of peripheral neuropathy
which is in contrast with findings in the early infantile form, in which a peripheral
demyelinating neuropathy is nearly constant.

Patients described by Bajaj et al (2002) showed similar heterogeneity: the first one (#17)
with a juvenile-onset, had difficulty with sports since adolescence, while the second one
(#28) lived normally until his 30s. On the other hand, all compound heterozygotes for the
30Kb deletion and the G270D mutation (genotype found in three different families and
six patients) had an adult-onset and showed a peripheral neuropathy. Such a variability
was also reported from analysis of the international registry. Nevertheless, while the
concomitant occurrence among siblings of a late infantile form with an essentially adult
onset form has been described, a multiplex family with an infantile form and a late onset
form has to our knowledge never been reported. As reported for the early onset forms of
Krabbe disease, the 30 Kb deletion associated with a 502C>T polymorphism is also the
most common mutation described in the adult form, albeit always in a compound
heterozygous state. From published studies, the presence of G270D or the missplicing
G49G mutations appear predictive of a slowly progressive disease, whatever the nature
of the associated allele. The mutations Y303C, G622S, M617T, E215K, L629R, R63H,
R515C, L618S also seem associated with late onset forms. However, none of these
mutations can be predictive of an adult-onset form.
158
Figure (114): Krabbe disease defective enzyme (www.google.com)

Figure (115): Krabbe disease (www.google.com)

1.5 Farber Disease


Ceramidase deficiency (Farber lipogranulomatosis or Farber Disease) leads to tissue
accumulation of ceramide due to deficient activity of lysosomal ceramidase.

The clinical presentation of Farber Disease (FD) is characterized by the appearance of


subcutaneous skin nodules, ordinarily near the joints, most often interphalangeal, wrist,
elbow and ankle joints, or over points of mechanical pressure. These manifestations are
very painful and lead to progressive joint stiffness, limitation of motion by contractures
and finally to immobilization and deformation of joints. Also, a characteristic sign of
Farber Disease (FD) is the development of a progressive hoarseness due to laryngeal
involvement.

Beside these major manifestations seven phenotypes have been described which differ
in severity and additional organ involvement, like the lungs, nervous system, heart and
lymph nodes. Dependent on residual lysosomal ceramidase turnover, patients have a
159
variable degree of central nervous system disease, leading to progressive neurologic
deterioration. In most cases the neuronal dysfunction rather than the general physical
dystrophy seems to limit the duration of Farber Disease (FD). As well, patients with
Farber Disease (FD) may die due to pulmonary disease with interstitial pneumonia.

First symptoms usually appear between the newborn period and the first birthday. Milder
forms of type 3 were described with onset at 20 months of age. Clinical manifestation in
type 5 of Farber Disease (FD), dominated by neurologic deterioration, begins at 1 to 2
1/2 years of life. Patients mainly die within the first years of life, but prolonged courses in
patients without severe central nervous system (CNS) disease may also be observed.

-Phenotypes of Farber Disease:


Type 1 is the classic form of the disease with early subcutaneous nodules, joint
involvement and hoarseness in all cases. Progressive neurologic involvement and lung
disease are reported in many cases.

In contrast, type 2 and 3 patients show only slight or no symptoms of central nervous
system (CNS) disease. However, they still have a severe disease as a result of
granulomatous inflammation leading to subcutaneous nodules, joint pain and
contractures, hoarseness, failure to thrive and respiratory involvement.

Patients with type 4 Farber Disease (FD) present with severe neurologic deterioration
and large hepatosplenomegaly already in the neonatal period. Histopathology shows
massive granulomatous infiltrations by accumulating macrophages in liver, spleen,
lymphoid tissue, thymus and lungs.

The major clinical presentation in type 5 patients is a progressive central nervous system
(CNS) dysfunction, beginning at 1 to 2 1/2 years of life and manifestating in tetraplegia,
loss of speech, myoclonia, seizures and mental retardation.

Type 6 is a combination of type 1 Farber Disease (FD) and Sandhoff disease, another
lysosomal storage disorder caused by hexosaminidase A and B enzyme defects. Both
acid ceramidase and hexosaminidase A and B are involved in the catabolism of
glycosphingolipids.

A study revealed that one patient is classified as type 7, showing a combined deficiency
of glucocerebrosidase, galactocerebrosidase and ceramidase due to a mutation of
prosaposin, the precursor protein for two sphingolipid activator proteins.

In typical cases of type 1 Farber Disease (FD) the clinical triad of subcutaneous nodules,
joint and laryngeal involvement verifies the disease. When typical features are missing,
diagnosis is confirmed by determination of acid ceramidase activity, which is less than 6
percent of control values, measured in cultured skin fibroblasts, white blood cells or
amniocytes. Another diagnostic approach is the demonstration of typical
histopathologic features on biopsy, showing granulomas with macrophages containing
lipid cytoplasmic inclusions in subcutaneous nodules or other tissues. Determination of

160
ceramide accumulation in tissues by chromatography or mass spectrometry is also an
established diagnostic test for Farber Disease (FD).

The curative effect of hematopoietic stem cell transplantation (HSCT) on the


inflammatory symptoms in Farber Disease (FD) indicates that the granulomatous
inflammation in these patients is not a consequence of mere ceramide storage but reflects
a dysregulation of leukocyte functions, probably due to the intracellular role of ceramide
in intracellular signal transduction. However, the sequence of molecular mechanisms
leading from a defect in ceramide metabolism to chronic granulomatous inflammation
still needs elucidation. Alterations of receptor-mediated apoptosis by ceramide
accumulation in inflammatory cells may be one of the mechanisms underlying abnormal
granuloma formation.

Recent evidence suggests that the sphingolipid ceramide plays an important role as a
second messenger in a number of signal transduction pathways. Many cellular responses
to extracellular stimuli have been linked to the intracellular generation of ceramide,
especially the induction of apoptotic cell death triggered by various stress agents.
Ceramide has been proposed to mediate apoptosis, because many pro-apoptotic stimuli
induce the production of ceramide and treatment with exogenously added ceramides can
induce apoptotic cell death.

In a previous study we have shown that cells obtained from a Farber Disease (FD) patient
showed remarkable changes in their sensitivity to different apoptotic stimuli. When
treated with staurosporine, chemotherapeutic drugs or ionizing radiation, Farber disease
(FD) cells underwent apoptosis and activated caspases comparable to control cells.
However, due to the lack of ceramidase, cell-permeable ceramides had a stronger pro-
apoptotic activity in Farber cells than in controls. Interestingly, it is consistently
observed an accelerated rate of lymphocyte death in Farber Disease (FD) upon
activation of the death receptor molecule fas (CD95). These data suggest that ceramide
does not play an essential role as a general trigger or second messenger in all kinds of
apoptosis, but may rather act as a specific amplifier of receptor-induced cell death. There
is growing evidence that generation of ceramide increases membrane fluidity and raft
formation in the plasma membrane, thereby facilitating receptor mediated signaling.
Indeed, recent studies showed that ceramide may be critically involved in cap formation,
clustering, and activation of the CD95 receptor.

As a major symptom Farber Disease (FD) patients exhibit chronic destructive joint
inflammation resembling rheumatoid arthritis. Indeed, increased CD95 receptor/ligand
interaction has been implicated in the pathogenesis of inflammatory arthritis. It may
therefore be speculated that increased CD95 signaling mediated by elevated ceramide
levels is involved in the inflammatory arthritis of Farber Disease (FD).

161
Figure (116): Enzyme deficiency in Farber disease (www.google.com)

Figure (117): Acid ceramidase in healthy and Farber diseased cell membranes
(www.google.com)

162
Figure (118): Farber disease (www.google.com)

1.6 Metachromatic Leukodystrophy


Leukodystrophy is defined as a group of genetic disorders which are characterized by the
imperfect growth/development of the myelin sheath that covers nerve fibers in the brain.
Myelin, which lends its color to the white matter of the brain is a complex substance
made up of at least ten different chemicals. Leukodystrophies are a group of disorders
that are caused by genetic defects in how myelin produces or metabolizes these
chemicals. Mutation in gene is the main reason for the development of leukodystrophy.
The most common symptom of a leukodystrophy disease is a gradual decline in an infant
or child who previously appeared well. The clinical features consist of progressive
intellectual deterioration with varying degrees of pyramidal and cerebellar dysfunction.
The course of the disease is usually progressive. Seizures, though infrequent, may occur.
Enlargement of the head is a feature in some varieties. The disorder may be inherited in a
recessive, dominant or X-linked manner, depending on the type of leukodystrophy.
Specific leukodystrophies include metachromatic leukodystrophy, Krabbe disease,
adrenoleukodystrophy, Pelizaeus-Merzbacher disease, Canavan disease, Childhood
Ataxia with Central Nervous System Hypomyelination or CACH (also known as
Vanishing White Matter Disease), Alexander disease, Refsum disease, and
cerebrotendinous xanthomatosis. Metachromatic Leukodystrophy(MLD) also called
Arylsulfatase A deficiency is one such type of leukodystrophy which is usually caused
by the lack of an important enzyme called arylsulfatase A. Metachromatic
Leukodystrophy (MLD) occurs in about 1 in 40000 people. Due to arylsulfatase A
enzyme deficiency, chemicals called sulfatides build up in and damage the nervous
system, kidneys, gallbladder, and other organs. In particular, the chemicals damage the
protective sheaths that surround nerve cells. There are three forms of the disease. They
are based on when symptoms begin:

- Late infantile Metachromatic Leukodystrophy(MLD) symptoms usually begin by


ages 1-2 years.
- Juvenile Metachromatic Leukodystrophy(MLD) symptoms usually begin between
3-10 years
- Adult [Late-stage juvenile Metachromatic Leukodystrophy(MLD)] symptoms
may occur over age 16 years.
163
Metachromatic leukodystrophy (MLD) belongs to a family of disorders identified as
lysosomal storage diseases. This disorder is characterized by the lysosomal
accumulation of sulfated glycolipids, specifically 3-O-sulfogalactosyl-containing
glycolipids, as a consequence of defects in the lysosomal hydrolase, arylsulfatase A
(ARSA). The major site of 3-O-sulfogalactosyl-containing glycolipids is the myelin
sheaths of central and peripheral neurons. Because of this location the clinical
manifestations of Metachromatic Leukodystrophy (MLD) are predominately
neurological in nature. Histopathologically, Metachromatic Leukodystrophy (MLD) is
characterized by demyelination of central and peripheral nerves. The accumulation of
sulfated glycolipids in the lysosomes results in the characteristic metachromatic staining
of the tissues, hence the derivation of the name of this disease. Metachromatic
Leukodystrophy (MLD) has an autosomal recessive inheritance pattern. The disease is
passed down through families (inherited). A copy of the defective gene from both parents
should be transmitted to the offspring to have the disease. Parents can each have the
defective gene and remain as carriers but not have Metachromatic Leukodystrophy
(MLD). Children who inherit only one defective gene from one parent will be a carrier,
but usually will not develop Metachromatic Leukodystrophy (MLD). When two carriers
have a child, there is a 25% chance that the child will get both genes and have
Metachromatic Leukodystrophy (MLD).

Clinical features of Metachromatic Leukodystrophy (MLD) include mental


deterioration, hypotonia (low muscle tone), developmental delay, speech abnormalities,
loss of mental abilities, blindness, rigidity, convulsions, impaired swallowing, paralysis,
dementia, impaired school performance, ataxia, tremors, seizures, dementia.

Figure (119): Diseases related to metachromatic leukodystrophy

1.7 Tay-Sachs Disease


Tay–Sachs disease, is caused by mutations in a β-hexosaminidase and results in the
buildup of GM2, culminating in irreversible fatal deterioration of brain function.
Unfortunately, enzyme replacement delivery to the brain has not yet been successfully
developed. In Tay-Sachs disease (hemoaminidase A deficiency), there is accumulation
of GM2 ganglioside in the heart, liver, and spleen. Involvement of the central nervous
system with vacuolated neurons is predominant. Patients present at 6 months of age, and
the disease is fatal by age 2 to 3 years.

Tay-Sachs disease (also known as GM2 gangliosidosis-variant B) and its variant forms
are caused by a deficiency in the enzyme hexosaminidase A. The incidence has been
particularly high among Eastern European and Ashkenazi Jewish populations, as well as
certain French Canadians and Louisianan Cajuns. Affected children appear to develop
164
normally for the first few months of life. Symptoms begin by 6 months of age and include
progressive loss of mental ability, dementia, decreased eye contact, increased startle
response to noise, progressive loss of hearing leading to deafness, difficulty in
swallowing, blindness, cherry-red spots in the retina, and some paralysis. Seizures may
begin in the child's second year. Children may eventually need a feeding tube and they
often die by age 4 from recurring infection. No specific treatment is available.

Tay-Sachs disease (TSD), also known as GM2 gangliosidosis is an autosomal recessive,


progressive neurodegenerative disorder. This results from a deficiency of β-
hexosaminidase A enzyme that leads to an accumulation of GM2 gangliosides. In the
infantile form, progressive loss of neurological function ensues and is usually fatal by
age 4 or 5 years. The 'cherry-red spot' is the hallmark of Tay–Sachs disease (TSD). It is
the result of GM2 accumulation in the retinal ganglion cells, giving the white fundus
appearance surrounding the normal tint of the fovea. Tay–Sachs disease (TSD) is an
autosomal recessive genetic disorder. The most common variant becomes apparent in
infancy. Infants with this disorder typically appear normal until the age 3 to 6 months,
when their development slows and muscles used for movement weaken. Early
milestones are lost and loss of visual attentiveness occurs early. As the disease
progresses, progressive neurological deterioration occurs with seizures, intellectual
disability and paralysis. It is usually fatal by age 4 or 5 years. Other forms of Tay–Sachs
disease (TSD) are very rare. Neurological signs and symptoms can appear in childhood,
adolescence and adulthood and are usually milder than those seen in the infantile form.
While Tay–Sachs disease (TSD) is very rare in the general population, the genetic
mutations that cause this disease are more common in people of Ashkenazi (eastern and
central European communities) Jewish heritage than in those with other backgrounds.
Approximately 1 in 30 people of Ashkenazi Jewish ancestry is a recessive carrier. About
1 in 360 000 people in the general population develops Tay–Sachs disease (TSD)
compared to 1 in 3 600 people of Ashkenazi Jewish ancestry. The pathogenesis of
Tay–Sachs disease (TSD) is attributable to the accumulation of GM2 gangliosides
resulting from deficiency of β-hexosaminidase A enzyme caused by a mutation in the
alpha subunit of the HEXA gene on chromosome 15q. As a result, GM2 gangliosides
accumulate to toxic levels, particularly in neurons in the central nervous system.
Progressive damage caused by the buildup of GM2 gangliosides lead to destruction of
these cells. A 'cherry-red spot' at an early stage is one of the dramatic diagnostic clues
found on fundus examination. It is due to GM2 gangliosides accumulation in the retinal
ganglion cells leading to thickening and loss of transparency of the posterior pole of the
retina. The absence of ganglion cells at the fovea gives rise to the red spot surrounded by
white diseased cells. Its color is due to the pigment epithelium and choroid, and therefore
may demonstrate color variability according to the race. As the ganglion cells die, the
'cherry red spot' fades and optic atrophy becomes apparent. Although experimental
work, such as gene therapy research is underway, there is currently no cure or treatment
to slow the progression of Tay–Sachs disease (TSD). Cherry-red spot' is a useful clinical
indicator in Tay–Sachs disease (TSD) and several other lysosomal storage disorders. A
useful sign, when associated with key clinical features and a good history, it often guides
one to a diagnosis of the disease. It serves as an ideal illustration of the eye as a window to
inborn errors of metabolism.

165
Figure (120): Non-digestion of ganglioside GM2 by lysosomes in Tay-Sachs
disease (www.google.com)

Figure (121): Pathophysiology of Tay-Sachs and Sandhoff diseases


(www.google.com)

1.8 Sandhoff Disease


Sandhoff disease is a rare but severe lysosomal storage disorder caused by a deficiency
of both hexosaminidases A and B resulting in accumulation of glycosphingolipids and
oligosaccharides in the brain. It represents 7% of cases of GM2 gangliosidosis. Bilateral
thalamic involvement has been suggested as a diagnostic marker of Sandhoff disease
white matter and cerebellum can also be involved. Sandhoff disease is a rare autosomal
recessive inherited metabolic disorder caused by mutations in the HEXB gene on
chromosome 5q13. There are three different forms of Sandhoff disease: classic infantile,
juvenile and adult late onset. Classic infantile form is the most common and severe form,
clinical onset is generally between 3 and 9 months of age after a normal development and
is characterized by muscular hypotonia and tonico-clonic or myoclonic seizures,
blindness, psychomotor retardation and paralysis. A cherry-red spot in macular areas is
characteristic but not specific. Juvenile and adult onset forms of Sandhoff disease are
166
very rare. Signs and symptoms can begin in childhood, adolescence or adulthood and are
usually milder than those seen with the infantile form of Sandhoff disease. No treatment
is currently available and patients with infantile types usually die before 3 years of age.
Cerebral images of GM2 gangliosidosis have been published in a few reports and
bilateral homogeneous thalamic hyperdensity on computed tomography (CT) was
reported to be an early sign. Gliosis and intralysosomal storage in addition to calcium
accumulation might be responsible for this appearance. Magnetic resonance imaging
(MRI) typically reveals hypo-intensity in T2-weighted images of the thalamus and high
signal abnormalities have been reported in T2-weighted sequences in the caudate
nucleus, globus pallidus and putamen. Also the cerebral white matter shows
homogenous or patchy high signal intensity which suggests a combination of disturbed
and abnormal myelination and loss. An unusual presentation of Sandhoff disease was
reported by Nassogne et al. (2003) with diffuse signal changes exclusively in the
brainstem of a 3-year-old girl. Brainstem involvement appeared as a construction from
the internal capsule involvement and descended downward along the long fiber tracts of
the pons through the crus cerebri.

Sandhoff disease (variant AB) is a severe form of Tay-Sachs disease. Onset usually
occurs at the age of 6 months and is not limited to any ethnic group. Neurological signs
may include progressive deterioration of the central nervous system, motor weakness,
early blindness, marked startle response to sound, spasticity, shock-like or jerking of a
muscle (myoclonus), seizures, abnormally enlarged head (macrocephaly), and cherry-
red spots in the eye. Other symptoms may include frequent respiratory infections, heart
murmurs, doll-like facial features, and an enlarged liver and spleen.

Figure (122): Sphingolipid metabolism diseases (www.google.com)

2.Hyperlipidemia
2.1Classification of Hyperlipidemia.
- On the basis of lipid type:
1- Hypercholesterolemia: in this the level of cholesterol is elevated.
2- Hypertriglyceridemia: it is defined as an elevated level of triglycerides.
167
- On the basis of causing factor:
On the basis of causing factors hyperlipidemia can be designated as either primary or
secondary. According to Fredrickson primary (familial) hyperlipidemia is classified into
five types on the basis of electrophoresis or ultracentrifugation pattern of lipoproteins.

1- Type I–Raised cholesterol with high triglyceride levels.


2- Type II–High cholesterol with normal triglyceride levels.
3- Type III–Raised cholesterol and triglycerides.
4- Type IV–Raised triglycerides, atheroma and uric acid.
5- Type V–Raised triglycerides.

This classification was later adopted by World Health Organization (WHO). This
method does not directly account for high-density-lipoprotein (HDL) and also does not
distinguish among the different genes that may be partially responsible for some of these
conditions. It remains a popular system of classification but is considered dated by many.

Acquired (Secondary) hyperlipidemia: acquired hyperlipidemia (secondary


dyslipoproteinemias) results from underlying disorders and lead to alterations in plasma
lipid and lipoprotein metabolism. This type of hyperlipidemia may mimic primary forms
of hyperlipidemia and can have similar consequences. They may result in increased risk
of premature atherosclerosis, pancreatitis and other complications of the
chylomicronemia syndrome. The most common causes of acquired hyperlipidemia are
given below.

1- Diabetes Mellitus
2- Use of drugs such as diuretics, β-blockers and estrogens.
3- Alcohol consumption.
4- Some rare endocrine disorders and metabolic disorders.
5- Hypothyroidism
6- Renal failure
7- Nephrotic syndrome

Figure (123): Hyperlipidemia classification (www.google.com)


168
Figure (124): Fredrickson classification of hyperlipoproteinemias

2.2 Ischemic Stroke


Ischemic stroke or Cerebrovascular Accident (CVA) occurs when blood circulation in
part of the brain is blocked or diminished. When blood supply, which carries oxygen,
glucose, and other nutrients, is disrupted, brain cells die and become dysfunctional.
Usually, strokes occur due to blockage of an artery by a blood clot or a piece of
atherosclerotic plaque that breaks loose in a small vessel within the brain. Clinical trials
revealed that lowering of low-density-lipoprotein (LDL) and total cholesterol (TC) by
15% significantly reduced the risk of first stroke.

2.3 Impact of Hyperlipidemia on Cerebral Lipids


Hyperlipidemia is related to neurodegenerative diseases such as Alzheimer's Disease
and Niemann-Pick disease. Human epidemiological studies have revealed that high
cholesterol intake increases the risk of impaired cognitive function, and neuroimaging
has shown inverse associations between triglyceride levels and cerebral blood flow. The
significant negative effects of a high-fat/high-cholesterol diet on hippocampal
morphology have been demonstrated in animal research. It is widely accepted that
hyperlipidemia may lead to atherosclerosis, and hyperlipidemia has been identified as an
independent risk factor for coronary heart disease and ischemic stroke. Extensive
epidemiologic and experimental studies have supported the conception that
hyperlipidemia may cause nervous system-related diseases. Free fatty acids (FFA) in the
serum reflects the level of lipid metabolism and mediates cell damage caused by
oxidative stress. Brain free fatty acids (FFA) content is extremely low under normal
conditions, but it increases when circulation disorders occur in the brain. An increase in
brain triglycerides (TG) could lead to the increase of free fatty acids (FFA) levels in the
brain, which would result in damage to the neurons. Free fatty acids (FFA) may induce
the production of oxidized low density lipoprotein (ox-LDL), while total cholesterol
(TC) stimulates the precipitation of amyloid beta (Aβ) peptide and the expression of
vascular endothelial growth factor (VEGF) in the brain. Oxidized low density
lipoprotein (ox-LDL), amyloid beta (Aβ) peptide and vascular endothelial growth factor
(VEGF) in the brain may destroy the permeability of the blood–brain barrier by
169
damaging the vascular endothelial cells, aggravating the injury and apoptosis of neurons
as a result. The elevated triglycerides (TG) levels produce a large number of free fatty
acids (FFA) in the process of its metabolism, and this aggravate the injury to cerebral
vessels and neurons by enhancing the oxidation reaction and promoting the production
of oxidized low density lipoprotein (ox-LDL) modified by low-density-lipoprotein
(LDL).

2.4 Mild Cognitive Impairment in Familial Hypercholesterolemia


Alzheimer's disease is a progressive neurodegenerative disorder characterized by global
deterioration of cognition and behavior. Development of dementia in Alzheimer's
disease is usually preceded by a prodromal stage of abnormal cognitive performance
known as mild cognitive impairment. A major neuropathological feature of Alzheimer's
disease is the increase in insoluble amyloid fibrils composed of 40–42 amino acid
peptides known as the amyloid beta peptide (Aβ).

Recent studies suggest a connection between cholesterol metabolism and the


pathogenesis of Alzheimer's disease. We reported that hypercholesterolemia accelerates
amyloid beta peptide (Aβ) production in the brain of transgenic mice and associates with
higher levels of amyloid beta peptide (Aβ) in the human brain. The studies suggest that
hypercholesterolemia is only an early (not a late) risk factor for Alzheimer's disease.

A study found an association between familial hypercholesterolemia and mild cognitive


impairment. The proportion of familial hypercholesterolemia patients exhibiting
abnormal cognitive function and meeting criteria for mild cognitive impairment (21.3%)
was significantly higher than that observed in the control group (2.9%; p = 0.00) and far
exceeded the age-specific prevalence predicted from either epidemiological studies in
the general population or the prevalence observed in follow-up of large cohorts with
milder sporadic hypercholesterolemia.

The term mild cognitive impairment is generally used to define a transitional stage
between normal cognitive function and dementia. Estimates of its progression rate to
Alzheimer's disease range from 10 to 15% per year compared to 1–2% for cognitively
intact subjects.

Familial hypercholesterolemia may offer a unique window into the role of cholesterol
metabolism in cognition. Two aspects of familial hypercholesterolemia may be of
particular relevance to Alzheimer's disease. The first is that patients afflicted with this
disorder are exposed to higher cholesterol levels from early in life. This is important
because hypercholesterolemia may be an early risk factor for Alzheimer's disease. The
second feature is the involvement of low-density-lipoprotein (LDL) receptors in familial
hypercholesterolemia. Low-density-lipoprotein (LDL) receptors have been implicated
in synaptic maintenance and in Alzheimer's disease pathogenesis. Members of the low-
density-lipoprotein (LDL) receptors family are involved in amyloid beta peptide (Aβ)
clearance and synaptic plasticity from the brain as supported by a growing body of
literature. One study showed that when an Alzheimer's disease mouse model of
amyloidosis was crossed into an low-density-lipoprotein (LDL) receptors-deficient
background, the mice not only developed exacerbated age-dependent cerebral beta-
170
amyloidosis but also, more severe behavioral abnormalities than observed in low-
density-lipoprotein (LDL) receptors-intact Alzheimer's disease mice.

It is proposed that either early exposure to cholesterol or dysfunction of low-density-


lipoprotein (LDL) receptors contribute to cognitive dysfunction in patients with familial
hypercholesterolemia and it is possible that similar mechanisms may be involved in mild
cognitive impairment not associated with familial hypercholesterolemia.

Figure (125): Impact of Hyperlipidemia on cognitive dysfunction

Figure (126): Link between dyslipidemia and Alzheimer's disease


(www.google.com)

2.4.1 Lipid Metabolism and Alzheimer's Disease


Recent evidence suggests a link between lipid metabolism (especially cholesterol) and
susceptibility to Alzheimer's disease (AD). In many studies, elevated plasma cholesterol
levels were associated with increased risk of Alzheimer's disease (AD). In some studies,
administration of statins, a class of cholesterol-lowering drugs, was associated with a
decreased risk for Alzheimer's disease (AD). Moreover, cholesterol accumulates in

171
amyloid plaques and nerve terminals in Alzheimer's disease (AD) brains. Genetic
studies have linked Alzheimer's disease (AD) susceptibility to genes related to
cholesterol metabolism, including Apo E, a major cholesterol transporter in the
circulation and in the brain. Therefore, abnormal lipid metabolism could be an important
early event in the pathogenesis of Alzheimer's disease (AD). At the cellular level, an
increase in cellular cholesterol content stimulates the production and accumulation of
amyloid beta peptide (Aβ), a molecule of central importance in the current model of
Alzheimer's disease (AD) pathogenesis, in both cultured neurons and Alzheimer's
disease (AD) brains. Alternatively, amyloid precursor protein (APP) processing and
amyloid beta peptide (Aβ) production may also affect cellular lipid metabolism, leading
to alterations in the generation or turnover of cholesterol or sphingolipids. It is possible
that abnormal cholesterol metabolism, which could be a consequence of the presence of
ApoE4, is an early event in Alzheimer's disease (AD) pathogenesis, leading to altered
amyloid precursor protein (APP) processing and increased amyloid beta peptide (Aβ)
production and neurotoxicity, which could in turn exacerbate the lipid disorder. A
deleterious feedback loop between abnormal cholesterol metabolism and amyloid beta
(Aβ) peptide production and neurotoxicity could be one of the molecular mechanisms
underlying the link between lipids and Alzheimer's disease (AD) . Many studies of the
relationship between plasma lipid and lipoprotein levels and the risk of Alzheimer's
disease (AD) have shown a positive correlation. In 1998, Notkola and colleagues found
that a high total cholesterol level increased the risk of developing Alzheimer's disease
(AD). It was reported a similar tendency in Alzheimer's disease (AD) patients and further
showed that the plasma levels of total and low-density-lipoprotein cholesterol (LDL-C)
correlated with the amount of amyloid beta (Aβ) peptide in Alzheimer's disease (AD)
brains. This finding was confirmed in another study, which also showed that the
association of plasma cholesterol levels with Alzheimer's disease (AD) risk was
progressively stronger with increasing pathological certainty of Alzheimer's disease
(AD) diagnosis. The cholesterol–AD association was also observed in a population-
based study in African–Americans. Interestingly, elevated cholesterol levels have also
been linked to vascular dementia, suggesting a general role for abnormal lipid
metabolism in neurodegeneration. However, unlike total or low-density-lipoprotein-
cholesterol (LDL-C), elevated plasma levels of high-density- lipoprotein- cholesterol
(HDL-C) are associated with a significantly decreased risk of dementia. Although high
plasma cholesterol levels are a risk factor for early amyloidogenesis and Alzheimer's
disease (AD) in midlife, cholesterol levels are not associated with Alzheimer's disease
(AD) onset in old age. For example, in a study of 1449 subjects aged 65–79 years who
were followed for an average of 21 years, elevated plasma cholesterol levels in midlife
were a significant risk factor for mild cognitive impairment, which has been considered
to be a predictor of Alzheimer's disease (AD). In the same study, elevated systolic blood
pressure or high plasma cholesterol levels in midlife significantly increased the risk of
Alzheimer's disease (AD) in later life. In a study of the relationship between Alzheimer's
disease (AD) pathology and hypercholesterolemia, hypercholesterolemia correlated
with amyloid deposition in only the youngest subjects (40–55 years of age). These
findings suggest a role for midlife vascular risk factors, such as hypercholesterolemia
and hypertension, in the development of Alzheimer's disease (AD) in late life. Thus, it
might not be surprising that the association between cholesterol levels and Alzheimer's

172
disease (AD) is weak in elderly people. It is not clear how plasma cholesterol levels
affect Alzheimer's disease (AD) risk. It is generally accepted that brain cholesterol is
synthesized in situ and, owing to the blood–brain barrier, plasma cholesterol has little
effect on brain cholesterol levels. One possibility is that elevated plasma cholesterol
levels cause cerebrovascular disease, such as atherosclerosis, leading to decreased brain
metabolism, neuronal dysfunction and, finally, dementia. Alternatively, both increased
plasma cholesterol levels and Alzheimer's disease (AD) could be due to other pathogenic
factors, such as aging and ApoE4. In fact, plasma cholesterol levels increase with aging
and in the presence of ApoE4.

2.4.2 Effect of Amyloid-β on Lipid Metabolism:


It has been suggested that amyloid precursor protein (APP) processing affects cellular
lipid metabolism. In cultured neurons and in transgenic mice, amyloid beta (Aβ) peptide
with 42 amino acids (Aβ42) can activate neutral sphingomyelinases and downregulate
sphingomyelin levels, whereas Aβ40 reduces de novo cholesterol synthesis by inhibiting
the activity of hydroxymethylglutaryl-CoA (HMG-CoA) reductase. Therefore,
maintaining lipid homeostasis could be a biological function of amyloid precursor
protein (APP) processing, and the pathological accumulation of amyloid beta (Aβ)
peptide could lead to abnormal lipid metabolism. Furthermore, both studies in vitro and
in Alzheimer's disease (AD) patients suggest that amyloid beta (Aβ) peptide causes
oxidative stress, leading to lipid oxidation that might contribute directly to
neurodegeneration. Studies also suggest that amyloid beta (Aβ) peptide induces
ozonolysis of cholesterol, leading to the formation of peroxi-derivatives that accelerate
aggregation of amyloid beta (Aβ) peptide monomers and that amyloid beta (Aβ) peptide
oxidizes cholesterol at positions of 7-β and 3-β, thus leading to H2O2 production.
Therefore, a deleterious feedback loop between amyloid beta (Aβ) peptide accumulation
and altered lipid metabolism could be one of the molecular mechanisms underlying the
link between lipid disorders and Alzheimer's disease (AD).

Figure (127): Linking lipids to Alzheimer's disease (www.google.com)

173
3. Parkinson's Disease
Parkinson's disease (PD) is a chronic, progressive neurodegenerative movement
disorder caused by the degeneration of dopamine-producing nerve cells in the brain.
Familial Parkinson's disease (PD) has been linked to mutations in the α-synuclein gene.
Since α-synuclein is the primary component of Lewy bodies, it is thought to have a
central role in the pathogenesis of Parkinson's disease (PD). However, the mechanisms
of its toxicity and aggregate formation remain unclear. Several studies using different
approaches have concluded that lipids are an important modifier of α-synuclein toxicity.
Synaptosomal fractions of brain lystates are enriched in α-synuclein. Interestingly, the
amino terminus of α-synuclein resembles the lipid-binding domains of some
apolipoproteins, suggesting a potential interaction of α-synuclein with lipids. In vitro, α-
synuclein does interact with small unilamellar vesicles containing sphingomyelin and
cholesterol, affecting lipid packing in the vesicles. Furthermore, the formation of α-
synuclein multimers correlated well with the length and degree of saturation of fatty
acids added to cells. This finding suggests that lipids can modulate α-synuclein
oligomerization, a key nucleation step in the formation of prefibrillar and fibrillar
aggregates of α-synuclein. Thus, manipulation of the fatty acid composition in brains
could be a way to reduce the formation of toxic α-synuclein species. In fact, dietary lipids
can affect Parkinson's disease (PD) progression. For example, dietary unsaturated fatty
acids might have a protective role in Parkinson's disease (PD), although a previous study
does not support this notion. A genome-wide screening in yeast also supports the notion
that lipid metabolism modulates α-synuclein toxicity. In yeast, expression of α-
synuclein alone caused only modest reduction in viability. In the screening, 86 of 4850
mutant yeast colonies were highly sensitive to α-synuclein toxicity. Remarkably, of 57
toxicity-modifier genes with known biological functions, 18 (32%) were related to lipid
metabolism and vesicle-mediated transport. In contrast, when a mutant huntingtin
fragment was used to identify modifier genes for Huntington's disease, only a few genes
fell into these categories.

Figure(128): Lipid disorders and Parkinson's disease (www.google.com)


174
4.Prion Diseases
Transmissible spongiform encephalopathies (TSEs) or prion diseases affect both
humans and animals. Human prion diseases are highly heterogeneous and include
sporadic [sporadic Creutzfeldt-Jakob disease (CJD)], genetic [genetic Creutzfeldt-
Jakob disease (CJD), Gerstmann-Sträussler-Scheinker syndrome, Fatal Familial
Insomnia] and infectiously acquired forms [Kuru, variant Creutzfeldt-Jakob disease
(CJD)]. Typically, Transmissible spongiform encephalopathies (TSEs) are
characterized by vacuolation (spongiosis) of the neuropil, gliosis, and neuronal loss.
Furthermore, the pathogenesis of prion diseases is usually associated with an abnormal
and progressive accumulation of scrapie isoform of the prion protein (PrPSc), a misfolded
isoform of the cellular prion protein PrPC . Though cellular prion protein PrPC is essential
for the development and the progression of prion disease, its function is still unknown.
Nevertheless, evidence is provided that cellular prion protein PrPC is involved in neuro-
protective and anti-apoptotic functions, and has a role in structure and maintenance of
synaptic plasticity, cell survival, proliferation and neurite outgrowth. In addition, recent
data suggest that cellular prion protein PrPC is involved in amyloid β (Aβ) peptide
neurotoxicity by acting as a receptor for amyloid beta (Aβ) peptide oligomers. cellular
prion protein PrPC is a ubiquitously expressed glycoprotein, with highest levels found in
the central nervous system (CNS). It is linked to the plasma membrane by a glycosyl-
phosphatidyl-inositol- (GPI-) anchor. Like many other glycosyl-phosphatidyl-inositol
(GPI)-anchored proteins, both cellular prion protein PrPC and scrapie isoform of the
prion protein (PrPSc) are found associated with lipid rafts, detergent resistant membrane
domains (DRMs) enriched in cholesterol and glycosphingolipids. Cellular prion protein
PrPC undergoes internalization, however, the mechanism is still debated because several
pathways, such as rafts/caveolae, caveolae-like and clathrin-dependent endocytosis
were reported to be involved. Once internalized, only 10% of cellular prion protein PrPC
is degraded, whereas the majority is recycled to the plasma membrane. The functional
significance of the recycling process is still unknown, however, endocytosis and
recycling processes appear to be essential for scrapie isoform of the prion protein (PrPSc)
formation. The subcellular compartment of prion conversion is still not entirely
elucidated. Several sites of conversion were proposed, including plasma membrane,
endolysosomal vesicles and endoplasmic reticulum (ER). More recently, two
independent studies performed demonstrated that recycling endosomes were a potential
site of prion conversion. However, a huge body of evidence suggests that cell surface
localization of cellular prion protein PrPC is required for conversion into scrapie isoform
of the prion protein (PrPSc), and that the first contact between cellular prion protein PrPC
and scrapie isoform of the prion protein (PrPSc) occurs at the plasma membrane.
Cholesterol has the key role in the lateral organization of lipid membranes, in the form of
specialized microdomains known as rafts. In the outer leaflet of the membrane, the long
and saturated acyl chains of sphingolipids strongly intercalate with cholesterol resulting
in the dense-organization of lipid ordered phases in the membrane. Cholesterol- and
sphingolipid-rich microdomains are bordered by a lightly packed lipid disordered phase
of unsaturated phospholipids. Distinct proteins can selectively partition into lipid rafts,
thus it is thought that lipid rafts serve as specific protein sorting sites. Lipid rafts are
assumed to be involved in intracellular trafficking of proteins and lipids, secretory and
endocytotic pathways as well as signal transduction pathways. Rafts are detergent-
resistant membrane domains which can be isolated when membranes are treated with
mild detergents. Cholesterol is structurally important for lipid raft formation, and both
175
cellular prion protein PrPC and scrapie isoform of the prion protein (PrPSc) are associated
with these membrane domains. Consequently, many studies in prion research focus on
the relationship between cholesterol metabolism and PrPSc/prion propagation. For
development of clinical disease, scrapie isoform of the prion protein (PrPSc) has to be
close to neuronal cellular prion protein PrPC in a contiguous compartment. Here, it is
likely that lipid raft localization of prion protein (PrP) isoforms in particular plays a key
role. Furthermore, cellular prion protein PrPC was shown to be related to signaling
platforms, including several intracellular effectors, such as Fyn, Lyn, PI3K/Akt, PKA,
ERK1/2, GSK3β and TACE among others. These effectors and the signaling pathways
in which they are involved may be affected when cellular prion protein PrPC is converted
into scrapie isoform of the prion protein (PrPSc). The disruption of PrPC-signaling events
may lead to the loss of its neuroprotective function and/or to the gain of a neurotoxic
function. While many studies characterized the involvement of cholesterol in prion
propagation, more recent studies pointed out that there is an inextricable relationship
between prion infection and cholesterol metabolism. Using microarray analysis on
neuronal cells infected with 22 L prions an up- regulation of genes involved in the
cholesterol pathway was demonstrated. Sterol regulatory element-binding protein
(SREBP2), a transcription factor closely involved in the transcription of genes
associated with cholesterol biosynthesis and cholesterol up-take, was found to be
activated during prion infection. This resulted in increased levels of both total and free
cholesterol, suggesting a link between prion propagation and cholesterol balance. This
response to prion infection appears to be specific for neurons and is also apparent in
prion-infected mice at a preclinical stage of the disease. When the effects of prion
infection on cholesterogenic gene expression of primary neurons, astrocytes and a
microglial cell line were analyzed, only neurons responded with an up-regulation. In
astrocytes, no changes in cholesterogenic gene expression were found, whereas in a
microglial cell line prion infection resulted in a down-regulation of cholesterol
synthesis. Of note, the membrane fluidity in prion-infected cultured cells was shown to
be reduced, and it was demonstrated that prion-infected cells sequester cholesterol in
cellular membranes. These findings may be linked to the described up-regulation of
cholesterol synthesis. Surprisingly, ATP-binding cassette transporter (ABCA1) was
found to be increased during prion infection in both cell culture and mouse models of
prion infection. ATP-binding cassette transporter (ABCA1) is a membrane-associated
protein involved in cellular cholesterol export. Its localization under normal conditions
is known to be mainly at the cell membrane. Upon prion infection, ATP-binding cassette
transporter (ABCA1) was overexpressed even though the level of free cholesterol in
infected cells was increased. However, ATP-binding cassette transporter (ABCA1) was
found internalized during prion infection. The latter explained the increased level of free
cholesterol in infected cells. This effect was reversed when prion infection was cured.
When ATP-binding cassette transporter (ABCA1) was overexpressed, scrapie isoform
of the prion protein (PrPSc ) formation was reduced, probably through the cholesterol
pathway since the levels of prion protein (PrP) gene expression or protein amounts were
unchanged. One possible explanation for the ATP-binding cassette transporter (ABCA1)
enhancement is the cross-talk between cholesterol and the abnormal prion protein.
During prion infection cholesterol synthesis is increased, and this imbalance in
cholesterol homeostasis could lead to an overexpression of different genes involved in
the cholesterol pathway in order to regulate this impairment. These results clearly
demonstrate that there is a mutual relationship between cholesterol balance and prion
176
infection. However, it is unclear whether an increase in cholesterol levels is necessary to
enable persistent prion propagation. It is striking that in prion infection, cholesterol plays
a critical role in the formation of the associated protein aggregates. In addition, prion
infection and amyloid beta (Aβ) peptide accumulation interfere in neuronal cholesterol
metabolism, and it appears that in this condition cellular cholesterol levels are increased,
which can be toxic for cells. Cholesterol is an important regulator of membrane fluidity
and function. It is involved in neurotransmission processes, and it promotes
synaptogenesis, in which it could be a limiting factor. Cholesterol is critical for the
formation of synaptic vesicles containing neurotransmitters; it forms lipid rafts which
are important on the presynaptic membrane for exocytosis of neurotransmitter, and on
the post synaptic membrane, where the receptors of neurotransmitters are embedded.
With all these functions, it is obvious that cholesterol is an important regulator of
neuronal activity and any dysregulation in its metabolism might have serious
consequences on brain function. Therefore, it is reasonable to argue that an imbalance in
cholesterol homeostasis may be one factor accountable for synaptic dysfunction and
neurodegeneration in prion diseases.

Figure (129): Replication cycle of prion (www.google.com)

Figure(130): Prions disrupt cellular function (www.google.com)


177
Figure(131): Pathological processes culminating in Prion and Alzheimer's
disease (www.google.com)

5.Smith-Lemli-Opitz Syndrome
Smith–Lemli–Opitz syndrome (SLOS; OMIM 270400) is intellectual disability and
behavioral problems. In 1993, increased levels of 7–dehydrocholesterol (7DHC) and
decreased levels of cholesterol were found in Smith–Lemli–Opitz syndrome (SLOS)
patients. This abnormal sterol profile was consistent with a deficiency of
7dehydrocholesterol reductase (DHCR7) activity. Subsequently, several groups
identified mutations of 7dehydrocholesterol reductase (DHCR7) in Smith–Lemli–Opitz
syndrome (SLOS) patients. Although present in other ethnic groups,
Smith–Lemli–Opitz syndrome (SLOS) appears to be most frequent in Caucasians of
northern European heritage. The incidence of Smith–Lemli–Opitz syndrome (SLOS)
has been estimated to be on the order of 1/20 000–1/70 000. Severely affected infants
have multiple major congenital anomalies and typically die in the perinatal period. In
contrast, a milder variant of Smith–Lemli–Opitz syndrome (SLOS) combines minor
physical anomalies with distinct behavioral and learning problems. Poor feeding and
postnatal growth failure are frequent early manifestations of Smith–Lemli–Opitz
syndrome (SLOS), and many infants require placement of a gastrostomy tube for
adequate nutritional support. Typical craniofacial features include microcephaly, a small
upturned nose, ptosis and micrognathia. Although cleft lip is not common, many patients
have cleft palate or bifid uvula. In male patients, genital abnormalities are frequently
observed. These range from small penis through various degrees of hypospadius in mild
and classical cases to ambiguous genitalia or gender reversal in more severely affected
infants. Limb findings are common. These include short thumbs, single palmar creases,
postaxial polydactyly and soft-tissue syndactyly of the second and third toes. Syndactyly
of the second and third toes has been described in over 95% of Smith–Lemli–Opitz
syndrome (SLOS) patients. More severely affected patients often have major
malformations of the brain (holoprosencephaly, agenesis/dysgenesis of the corpus
callosum), heart (atrial and ventricular septal defects, patent ductus arteriosus and
atrioventricular canal defect), lungs (abnormal segmentation) or gastrointestinal
anomalies (pyloric stenosis and colonic aganglionosis). In addition to the physical
manifestations, Smith–Lemli–Opitz syndrome (SLOS) patients have a distinct
behavioral phenotype. As infants, they can be irritable, lack interest in feeding and prefer
178
not to be held. Older children demonstrate various degrees of hyperactivity, self-
injurious behavior, temperament deregulation and sleep disturbances. Most
Smith–Lemli–Opitz syndrome (SLOS) children demonstrate autistic characteristics and
many meet the diagnostic criteria for autism. Although syndactyly of the second and
third toes is a normal variant, its presence in a child with other minor anomalies, failure
to thrive or poor growth, developmental delay or autistic behavior should prompt
consideration of Smith–Lemli–Opitz syndrome (SLOS). A clinical suspicion of
Smith–Lemli–Opitz syndrome (SLOS) is confirmed by demonstrating elevated
7dehydrocholesterol (7DHC) in plasma or tissues. Although ultraviolet (UV)
spectroscopy can detect 7dehydrocholesterol (7DHC), 7dehydrocholesterol (7DHC) is
best assayed using Gas Chromatography/Mass Spectroscopy (GC/MS). Sterol analysis
by Chromatography/Mass Spectroscopy (GC/MS) allows for the identification of
lathosterol and desmosterol, cholesterol precursors that accumulate in the rare 'SLOS-
like' syndromes of lathosterolosis and desmosterolosis, respectively. Although
frequently low, plasma cholesterol levels can be within normal limits in
Smith–Lemli–Opitz syndrome (SLOS) patients. In addition, because the standard
laboratory cholesterol test does not distinguish between cholesterol and
7dehydrocholesterol (7DHC), the measured 'cholesterol' value may fall within the
normal range due to the presence of significant amounts of dehydrocholesterol (DHC).
In short, a normal cholesterol level does not exclude Smith–Lemli–Opitz syndrome
(SLOS). Elevated 7dehydrocholesterol (7DHC) levels are relatively specific to
Smith–Lemli–Opitz syndrome (SLOS). Occasionally, mild elevations of
7dehydrocholesterol (7DHC) were observed in patients treated with psychiatric drugs
such as haloperidol or in patients with increased rates of cholesterol synthesis. Mild
elevations of 7dehydrocholesterol (7DHC) have also been reported in patients with
cerebrotendinous xanthomatomatosis. In these atypical cases, Smith–Lemli–Opitz
syndrome (SLOS) can be excluded by sterol analysis of fibroblasts or lymphoblasts
grown under conditions that induce endogenous cholesterol synthesis. Similar testing
can also be used in cases in which blood 7dehydrocholesterol (7DHC) levels are
equivocal. 7dehydrocholesterol reductase (DHCR7) mutation analysis can also be
performed to confirm a diagnosis of Smith–Lemli–Opitz syndrome (SLOS) or in cases
where biochemical testing is equivocal. A staged approach can be used to reduce the cost
of mutation analysis. Sequencing of exons 6–9 identifies approximately 85% of
7dehydrocholesterol reductase (DHCR7) mutations. If both mutations are not identified,
exons 3–5 can then be sequenced. Exons 1 and 2 are noncoding. In a small number of
biochemically positive patients, only a single heterozygous coding mutation has been
identified. In some of these cases, the second allele is not expressed. These nonexpressed
alleles likely represent uncharacterized promoter mutations.

In Smith–Lemli–Opitz syndrome (SLOS), problems may be caused by deficient


cholesterol, deficient total sterols, toxic effects of either 7dehydrocholesterol (7DHC) or
compounds derived from 7dehydrocholesterol (7DHC) or a combination of these
factors. Given the multiple biological functions of cholesterol, it is unlikely that a single
pathological mechanism underlies the varied malformations and clinical problems
found in these patients. Some of the malformations associated with Smith–Lemli–Opitz
syndrome (SLOS) are consistent with impaired sonic hedgehog (SHH) functioning.
Sonic hedgehog (SHH) plays an important role in pattern formation of the central

179
nervous system (CNS), facial structures and limbs. Mutations of sonic hedgehog (SHH)
can cause holoprosencephaly, a brain malformation found in some Smith–Lemli–Opitz
syndrome (SLOS) patients. Sonic hedgehog (SHH) is cholesterol modified, secreted
from a signaling cell, and binds to a receptor called Patched (PTCH). Patched (PTCH)
regulates transmembrane signaling in the responding cell by modulating the function of
a protein called Smoothened (SMO). A number of mechanisms by which sonic
hedgehog (SHH) signaling might be impaired in Smith–Lemli–Opitz syndrome (SLOS)
have been proposed. It was demonstrated that reduced total sterol levels in fibroblasts
derived from Smith–Lemli–Opitz syndrome (SLOS) mutant mice impair sonic
hedgehog (SHH) signal transduction in the responding cell due to inhibition of
Smoothened (SMO). Another work suggests that the amino terminus of
7dehydrocholesterol reductase (DHCR7) may interact directly with Smoothened
(SMO) to regulate sonic hedgehog (SHH) signaling and studies suggest that PTCH-
mediated transport of vitamin D3 [a metabolic product of 7dehydrocholesterol (7DHC)]
modulates Smoothened (SMO) function. The altered sterol composition in
Smith–Lemli–Opitz syndrome (SLOS) affects the physiochemical properties and
function of cellular membranes. Lipid rafts are ordered lipid domains that function in
signal transduction. Substitution of 7dehydrocholesterol (7DHC) for cholesterol alters
both lipid raft stability and protein composition. Substitution of 7dehydrocholesterol
(7DHC) for cholesterol also decreases membrane bending rigidity, a physiochemical
change that may explain abnormal secretory granule formation. Tulenko et al (2006)
showed that Smith–Lemli–Opitz syndrome (SLOS) membranes had altered increased
membrane fluidity and that synthetic membranes containing 7dehydrocholesterol
(7DHC) studied by X-ray diffraction had an atypical membrane organization. These
physical perturbations of membrane structure likely underlie functional defects in
immunoglobulin-E (IgE) receptor-mediated mast cell degranulation and cytokine
production, N-methyl-D-aspartate receptor (NMDA) receptor function and serotonin1A
receptor ligand binding. 7dehydrocholesterol (7DHC) and metabolic products of
7dehydrocholesterol (7DHC) may have toxic effects. Accumulation of
7dehydrocholesterol (7DHC) in fibroblasts impairs intracellular cholesterol transport
similar to that seen in Niemann-Pick Disease, type C, and 7dehydrocholesterol (7DHC)
appears to increase the degradation rate of hydroxymethyl glutaryl coenzyme A
reductase (HMG-CoA reductase). hydroxymethyl glutaryl coenzyme A reductase
(HMG-CoA reductase) catalyzes the rate-limiting step in cholesterol biosynthesis, and
increased degradation of this enzyme may contribute to decreased sterol synthesis in
Smith–Lemli–Opitz syndrome (SLOS) patients. Dehydrocholesterol (DHC) analogs of
pregnenolone, pregnanetriol, dehydroepiandrosterone (DHEA) and androstenediol
have been identified in Smith–Lemli–Opitz syndrome (SLOS) patients. 7-
dehydroallopregnanolone, a dehydrocholesterol (DHC) analog of the neuroactive
steroid allopregnanolone, has been identified in Smith–Lemli–Opitz syndrome (SLOS)
patients. A study identified a novel oxysterol, 27-hydroxy-7-DHC, in serum from
Smith–Lemli–Opitz syndrome (SLOS) patients. 27-hydroxy-7-dehydrocholesterol, in
contrast to 27-hydroxycholesterol, differentially activates liver X receptors (LXR).
These are nuclear receptors, which regulate lipid metabolism. It is not yet known
whether these 7dehydrocholesterol (7DHC)-derived steroids and oxysterols have
unique biological functions that contribute to the Smith–Lemli–Opitz syndrome (SLOS)
phenotype.
180
Figure (132): Smith-Lemli-Opitz syndrome enzyme deficiency (www.google.com)

Figure (133): Pathogenesis of Smith-Lemli-Opitz syndrome (www.google.com)

Figure (134): Biochemistry of Smith-Lemli-Opitz syndrome (www.google.com).


181
6.Hereditary Sensory and Autonomic Neuropathy Type1
Inherited peripheral neuropathies are common neurodegenerative disorders of the
peripheral nervous system. They are genetically and clinically heterogeneous, but based
on the predominant involvement of motor or sensory neurons, they fall into three main
subclasses: hereditary motor and sensory neuropathies (HMSN), hereditary motor
neuropathies (HMN) and hereditary sensory and autonomic neuropathies (HSAN). The
clinical spectrum of hereditary sensory and autonomic neuropathies (HSAN) is wide and
led to the further subdivision into six subtypes. Patients with hereditary sensory and
autonomic neuropathy type I (HSAN-I; MIM# 162400) present with progressive loss of
pain and temperature sensation in the limbs, variable distal muscle weakness and limited
autonomic dysfunction. Typically, first symptoms appear between the second and third
decade of life. The loss of sensation leads to painless injuries complicated by ulcerations
and osteomyelitis, often necessitating amputation. Hereditary sensory and autonomic
neuropathy type I (HSAN-I) has a dominant mode of inheritance and is associated with
mutations in SPTLC1 (MIM# 605712) and SPTLC2 (MIM# 605713), two subunits of
the enzyme serine palmitoyltransferase (SPT). Serine palmitoyltransferase (SPT) is
located at the outer membrane of the endoplasmic reticulum (ER), where it catalyzes the
pyridoxal-5'-phosphate (PLP) dependent condensation of L serine with palmitoyl-CoA.
This is the first and rate-limiting step in the de novo biosynthesis of sphingolipids.
Sphingolipids, having both a structural and a signaling function, are essential for all
eukaryotic cells. Mutations in various enzymes of the sphingolipid metabolism are
associated with neurodegenerative diseases, highlighting the importance of these
ubiquitous components in neuronal functioning. A common feature of the hereditary
sensory and autonomic neuropathy type I (HSAN-I) associated mutations in SPTLC1
and SPTLC2 is a reduction of the canonical enzymatic activity, but the effect on total
sphingolipid levels remains controversial. Furthermore, the mutations cause a shift in
the substrate specificity of serine palmitoyltransferase (SPT), enabling the mutant
enzyme to metabolize, besides serine, also alanine and glycine. This results in the
formation of the two atypical deoxysphingoid bases (DSB) 1-deoxy-sphinganine (1-
deoxy-SA) and 1-deoxymethyl-sphinganine (1-deoxymethyl-SA). Both metabolites
can be converted to 1-deoxy(methyl)-ceramide and 1-deoxy(methyl)-sphingosine [1-
deoxy(methyl)- SO] but because they lack an essential hydroxyl group, conversion to
more complex sphingolipids like glyco- or phosphosphingolipids and phosphorylation
are hampered, resulting in the accumulation of these intermediate metabolites in the cell.
Importantly, it was previously shown that deoxysphingoid bases (DSBs) have
pronounced neurotoxic effects on neurite formation in cultured sensory neurons. It was
performed a systematic screening of the known hereditary sensory and autonomic
neuropathies (HSAN) genes in a large cohort of hereditary sensory and autonomic
neuropathies (HSAN) patients and identified two novel SPTLC1 mutations, p.S331F
and p.A352V, bringing the total number of SPTLC1 mutations to six (the others being
p.C133W, p.C133Y, p.C133R and p.V144D). The p.S331F mutation occurred de novo in
a patient with a severe congenital phenotype, the p.A352V mutation was detected in an
isolated patient with a typical hereditary sensory and autonomic neuropathy type I
(HSAN-I) phenotype. Further evaluation of the pathogenicity of this latter variant is
warranted because of the absence of deoxyribonucleic acid (DNA) of family members
for segregation analysis, the limited evolutionary conservation of the targeted amino
acid and the similar chemical properties of alanine and valine. The biochemical evidence
182
reported so far indicates that all pathogenic mutants in SPTLC1 and SPTLC2 are
consistently associated with a reduction in the canonical serine palmitoyltransferase
(SPT) activity and an increased formation of deoxysphingoid bases (DSB). Results
demonstrate that the SPTLC1 mutation p.S331F, which is associated with a severe
hereditary sensory and autonomic neuropathy type I (HSAN-I) phenotype, has the same
effect on serine palmitoyltransferase (SPT) activity and deoxysphingoid bases (DSB)
formation as the previously characterized SPTLC1 mutations, suggesting that additional
factors influence the disease course. Furthermore, It is shown that the p.A352V
mutation, for which genetic evidence for pathogenicity is poor, is truly disease causing,
as it is associated with a reduced serine palmitoyltransferase (SPT) activity and elevated
deoxysphingoid bases (DSB) levels in patient plasma.

Figure (135):Mechanisms of hereditary sensory and autonomic neuropathy type


I disease

7.Huntington Disease
Huntington disease is an autosomal dominant neurodegenerative disorder characterized
by behavioral abnormalities, cognitive decline, and involuntary movements including
chorea and dystonia, that lead to a progressive decline in function and independence. The
onset of illness is typically in middle-age with a prevalence in most Caucasian
populations is about 10 per 100,000. Its prevalence is much less in African and Japanese
populations where it affects about 0.5 individuals per 100,000. The course of illness is
uniformly fatal and only one symptomatic treatment (tetrabenazine) for the involuntary
choreic movements of Huntington disease has been approved in the United States.
Although the neurologic manifestations are incapacitating and patients are at high risk of
suicide, cardiovascular disease, after pneumonia, is the leading cause of death in
afflicted individuals. More recent and detailed data regarding the exact etiology of
cardiovascular disease deaths are not available. It is hypothesized that disordered lipid
metabolism may contribute to neurological dysfunction and degeneration in Huntington
disease.

The pathophysiology of Huntington disease is linked to an expanded trinucleotide repeat


of cytosine-adenine-guanine (CAG) in the IT-15 gene on chromosome4. Individuals
who have inherited the Huntington disease genetic mutation are healthy on average for

183
the initial two thirds of life before the insidious emergence of motor, cognitive and
behavioral disturbances. Neuronal dysfunction in the clinically pre-manifest stages of
disease evolves eventually into neuropathological changes including prominent cell loss
and atrophy in the putamen and caudate (neostriatum) and the accumulation of
cytoplasmic and nuclear inclusions that contain the mutant protein huntingtin. Despite
the known genetic etiology of the disease, the exact function of the huntingtin protein
and its role in the pathogenesis of Huntington disease is not clear. Mitochondrial
dysfunction and bioenergetic defects may be contributing mechanisms.

Although cholesterol turnover has been shown to be very slow in the adult central
nervous system, abnormalities in cholesterol homeostasis have been associated with
neurodegenerative disorders that include Huntington disease, Alzheimer disease, and
Niemann-Pick type C. The biosynthesis of cholesterol and fatty acids is impaired in cell
cultures that include the Huntington disease genetic mutation and in animal models of
Huntington disease. The hypothesis that lipid dysregulation is pathogenic in Huntington
disease is supported by the observations that the mRNA transcription of key genes in the
cholesterol and fatty acid biosynthetic pathways is downregulated in human postmortem
Huntington disease striatal and cortical tissue as well as in murine models of Huntington
disease. The molecular mechanism that has been linked to impaired lipid biosynthesis is
a mutant huntingtin-dependent reduction in active sterol regulatory element response
protein 2 (SREBP-2). Since fatty acids are precursors of triglyceride and phospholipid
synthesis, the normal synthesis of all these important lipids require regulation by these
sterol regulatory element binding proteins.

Consistent with the reduced synthesis of cholesterol in cells affected by mutant


huntingtin protein, total cholesterol mass is reduced in the central nervous systems of
mice with a model of Huntington disease and in human cells in which the expression of
mutant huntingtin has been activated. Individuals with Huntington disease also have
total serum cholesterol concentrations approximately 40 mg/dL lower than healthy
individuals without Huntington disease. Although the correlation of cholesterol levels in
the central nervous system and blood are not well defined, levels of 24
hydroxycholesterol are approximately 10% of those in human plasma. Twenty-four
hydroxycholesterol is formed from cholesterol in the brain and is important for
cholesterol central nervous system homeostasis.

Fatty acid metabolism also appears disordered in individuals with Huntington disease.
SREBP-regulated genes affect both fatty acid and cholesterol metabolism, influencing
the elongation and desaturation of fatty acids. Fatty acid dysregulation is also supported
by the fact that fibroblasts in Huntington disease patients grow more slowly than those
from healthy individuals when they are in a lipid-deprived medium but their growth
normalizes when a mixture of linoleic and linolenic acids is added to the medium.
Evidence that fatty acid composition has been implicated as a factor affecting the fluidity
of cell membranes and preliminary data suggesting that freshly isolated cells from
diseased patients have altered membrane fluidity also suggest that dysregulated fatty
acid pathways exist.

The compartmentalization of fatty acids may also be altered in patients with Huntington
disease. Palmitoylation is the process by which fatty acids, such as palmitic acid,
184
covalently attach to residues of membrane proteins, such as cysteine. The precise
function of palmitoylation depends on the protein under consideration. However, this
process enhances the hydrophobicity of proteins and contributes to their membrane
association and the subcellular trafficking of proteins between membrane
compartments. The palmitoylation of huntingtin by huntingtin-interacting protein is
crucial for its normal function and transport but this process is impaired in those with
Huntington disease. Huntington-interacting protein is enriched in normal brain and co-
localizes with huntingtin in the striatum and in the medium spiny projection neurons, a
subset of neurons affected in Huntington disease. Reduced interaction between
huntingtin and huntingtin-interacting protein may contribute to the neuronal
dysfunction in Huntington disease by dysregulating normal neuronal intracellular
transport pathways.

A striking feature of Huntington disease is an impressive alteration in nutritional status


characterized by increased appetite and caloric intake. Paradoxically, this increase in
energy intake is usually accompanied by increased sedentary energy expenditure and
weight loss. The metabolic profile of transgenic mice with a model of Huntington
disease and those of humans with the disease appears to be distinct and indicates a
change from normal to one of a catabolic phenotype, particularly early in the disease.
One striking feature of this hypercatabolic state is that fatty acid and amino acid
catabolism has been shown to be abnormal. Of importance is the fact that this hyper-
catabolic state which occurs in Huntington disease patients is associated chronologically
with more rapid progression of the neurologic symptoms. Despite this interesting
phenomenon, the etiology for increased sedentary energy expenditure with subsequent
weight loss and more rapid disease progression in these patients is unknown.

A puzzling feature of Huntington disease patients is that, despite having a lower-than-


normal body mass index, they tend to have insulin resistance in peripheral tissues and an
increased risk for developing type 2 diabetes mellitus.

Figure (136): Brain cholesterol in Huntington's disease (www.google.com)


185
Figure (137): Altered cholesterol metabolism in Huntington's disease

Figure (138):Insulin resistance role in Huntington's disease pathologies


(www.google.com)

8.Phosphatidylserine in Deteriorating Brain


Aging of the human brain is associated with loss of neurons, dendritic atrophy, loss of
synaptic connections, decreased synaptic density, decreased synthesis of acetylcholine

186
(Ach) and other neurotransmitters, abnormal neuronal membrane lipid composition
(especially decreased membrane phosphatidylserine (PS) content and increased
membrane cholesterol content), and reduced sensitivity of postsynaptic membranes to
acetylcholine (Ach). A decrease in the ratio of phosphatidylserine (PS) to cholesterol
within neuronal membranes causes neurochemical changes that can contribute to an
increase in the viscosity of cellular membranes, thus reducing enzymatic activities that
require optimum fluidity. These cell membrane changes can be indirectly responsible for
alterations in enzymatic activities, receptor functions, membrane carriers, and neuronal
electrical characteristics, and can result in functional impairments. Phosphatidylserine
(PS) also may protect cell membranes from oxidative damage. In cell culture studies,
human neurons cultured in the presence of phosphatidylserine (PS) (25 mM) exhibited
significant reductions in electric shock-induced reactive oxygen species (ROS)
production, and phosphatidylserine (PS) supplementation has been reported to inhibit
the oxidation of cell membrane phospholipids by reactive oxygen species (ROS)
generated by xanthine oxidase. Concurrent with inhibition of oxidation of cell
membrane phospholipids, there was reduction in the rate of free radical-induced cell
death. Antioxidant defenses are bolstered by phosphatidylserine (PS); rats fed
phosphatidylserine (PS) up-regulated antioxidant enzyme activities in the brain
(superoxide dismutase and catalase) and liver (superoxide dismutase and glutathione
peroxidase) and the capacity of human high-density-lipoprotein (HDL) particles to
prevent the oxidation of circulating low-density-lipoprotein (LDL) particles is
proportional to the phosphatidylserine (PS) content of the high-density-lipoprotein
(HDL) particles. Aging of the human brain during adulthood is associated with
biochemical alterations and structural deterioration that impair neurotransmission.
Exogenous phosphatidylserine (PS) slows, halts, or reverses biochemical alterations and
structural deterioration in nerve cells and supports human cognitive functions, including
the formation of short-term memory, the consolidation of long-term memory, the ability
to create new memories, the ability to retrieve memories, the ability to learn and recall
information, the ability to focus attention and concentrate, the ability to reason and solve
problems, language skills and the ability to communicate, and locomotor functions,
especially rapid reactions and reflexes.

Figure (139): Phosphatidylserine metabolism in brain (www.google.com)

187
For Further Readings:
1. Bajaj, NP.; Waldman, A.; Orrell, R.; Wood, NW.; Bhatia, KP. (2002). Familial adult
onset of Krabbe's disease resembling hereditary spastic paraplegia with normal
neuroimaging. J Neurol Neurosurg Psychiatry, 72:635–638.
2. Beltroy, EP.; Richardson, JA.; Horton, JD.; Turley, SD.; Dietschy, JM. (2005).
Cholesterol accumulation and liver cell death in mice with Niemann-Pick type C
disease. Hepatology. ,42:886–893.
3. Bijlsma, MF.; Spek, CA.; Zivkovic, D.; van de Water, S.; Rezaee, F.;
Peppelenbosch, MP. (2006). Repression of smoothened by patched-dependent
(pro-) vitamin D3 secretion. Plos Biol, 4:e232.
4. Bi, X.; Liao, G. (2010). Cholesterol in Niemann-Pick type C disease. Subcell
Biochem, 51:319-335.
5. Block, R.; Dorsey, R.; Beck, C.; Brenna, T.; Shoulson, I. (2010). Altered cholesterol
and fatty acid metabolism in Huntington disease. J Clin Lipidol, 4(1):17-23.
6. Burek, C.; Roth, J.; Koch, HG.; Harzer, K.; Los, M.; Schulze-Osthoff, K. (2001).
The role of ceramide in receptor- and stress-induced apoptosis studied in acidic
ceramidase-deficient Farber disease cells. Oncogene, 20:6493–6502.
7. Chan, LY.; Balasubramaniam, S.; Sunder, R.; Jamalia, R.; Karunakar, TVN.;
Alagaratnam, J. (2011). Tay-Sach Disease with "Cherry-Red Spot" - First Reported
Case in Malaysia. Med J Malaysia, 66(5):497-498.
8. Chen, M.; Wang, J. (2008). Gaucher disease review of the literature. Arch Pathol
Lab Med, 132:851-853.
9. Cremesti, A.; Paris, F.; Grassme, H.; Holler, N.; Tschopp, J.; Fuks, Z.; Gulbins, E.;
Kolesnick, R. (2001). Ceramide enables fas to cap and kill. Journal of Biological
Chemistry, 276:23954–23961.
10. Debs, R.; Froissart, R.; Aubourg, P.; Papeix, C.; Douillard, C.; Degos, B.; Fontaine,
B.; Audoin, B.; Lacour, A.; Said, G.; Vanier, M.; Sedel, F. (2013). Krabbe disease in
adults: phenotypic and genotypic update from series of 11 cases and a review. J
Inherit Metab Dis, 36:859-868.
11. Ehlert, K.; Frosch, M.; Fehse, N.; Zander, A.; Roth, J.; Vormoor, J. (2007). Farber
disease: clinical presentation, pathogenesis and a new approach to treatment.
Pediatr Rheumatol Online J, 5(15).
12. Evans, RM.; Emsley, CL.; Gao, S.; Sahota, A.; Hall, KS.; Farlow, MR.; Hendrie, H.
(2000). Serum cholesterol, APOE genotype, and the risk of Alzheimer's disease: a
population-based study of African Americans. Neurology, 54(1):240-242.
13. Evans, W.; Hendriksz, C. (2017). Niemann-Pick type C disease-the tip of the
iceberg? A review of neuropsychiatric presentation, diagnosis and treatment.
BJPsychBull, 41(2):109-114.
14. Glade, M.; Smith, K. (2015). Phosphatidylserine and the human brain. Nutrition,
31:781-786.
15. Hannaoui, S.; Shim, S.; Cheng, Y.; Corda, E.; Gilch, S. (2014). Cholesterol balance
in prion diseases and Alzheimer's disease. Viruses, 6(11):4505-4535.
16. Jick, H.; Zornberg, GL.; Jick, SS.; Seshadri, S.; Drachman, DA. (2000). Statins and
the risk of dementia. Lancet, 356:1627-1631.
17. Jurevics, HA.; Kidwai, FZ.; Morell, P. (1997). Sources of cholesterol during
development of the rat fetus and fetal organs. J. Lipid Res.,38:723–733.
18. Kaushansky, K.; Lichtman, M.; Prchal, J.; Levi, M.; Press, O.; Burns, L.; Caligiuri,
188
M. (2016). Williams Hematology. 9th edition. Editors: Edmonson, K.; Lebowitz, H.
McGraw-Hill Education, USA. ISBN: 978-0-07-183300-4.
19. Kivipelto, M.; Helkala, E-L.; Hanninen, T.; Laakso, MP.; Hallikainen, M.;
Alhainen, K.; Soininen, H.; Tuomilehto, J.; Nissinen, A. (2001). Midlife vascular
risk factors and late-life mild cognitive impairment: a population-based study.
Neurology, 56(12):1683-1689.
20. Klionsky, DJ.; Abeliovich, H.; Agostinis, P.; Agrawal, DK.; Aliev, G.; Askew, DS.;
Baba, M.; Baehrecke, EH.; Bahr, BA.; Ballabio, A.; Bamber, BA.; Bassham, DC.;
Bergamini, E.; Bi, X.; Biard-Piechaczyk, M.; Blum, JS.; Bredesen, DE.; Brodsky,
JL.; Brumell, JH.; Brunk, UT.; Bursch, W.; Camougrand, N.; Cebollero, E.;
Cecconi, F.; Chen, Y.; Chin, LS.; Choi, A.; Chu, CT.; Chung, J.; Clarke, PG.; Clark,
RS.; Clarke, SG.; Clave, C.; Cleveland, JL.; Codogno, P.; Colombo, MI.; Coto-
Montes, A.; Cregg, JM.; Cuervo, AM.; Debnath, J.; Demarchi, F.; Dennis, PB.;
Dennis, PA.; Deretic, V.; Devenish, RJ.; Di Sano, F.; Dice, JF.; Difiglia, M.; Dinesh-
Kumar, S.; Distelhorst, CW.; Djavaheri-Mergny, M.; Dorsey, FC.; Droge, W.;
Dron, M.; Dunn, WA.; Jr Duszenko, M.; Eissa, NT.; Elazar, Z.;Esclatine, A.;
Eskelinen, EL.; Fesus, L.; Finley, KD.; Fuentes, JM.; Fueyo, J.; Fujisaki, K.;
Galliot, B.; Gao, FB.; Gewirtz, DA.; Gibson, SB.; Gohla, A.; Goldberg, AL.;
Gonzalez, R.; Gonzalez- Estevez, C.; Gorski, S.; Gottlieb, RA.;Haussinger, D.; He,
YW.; Heidenreich, K.; Hill, JA.; Hoyer-Hansen, M.; Hu, X.; Huang, WP.; Iwasaki,
A.; Jaattela, M.; Jackson, WT.; Jiang, X.; Jin, S.; Johansen, T.;Jung, JU.; Kadowaki,
M.; Kang, C.; Kelekar, A.; Kessel, DH.; Kiel, JA.; Kim, HP.; Kimchi, A.; Kinsella,
TJ.; Kiselyov, K.; Kitamoto, K.; Knecht, E.; Komatsu, M.; Kominami, E.; Kondo,
S.; Kovacs, AL.; Kroemer, G.; Kuan, CY.; Kumar, R.; Kundu, M.; Landry, J.;
Laporte, M.; Le, W.; Lei, HY.; Lenardo, MJ.; Levine, B.; Lieberman, A.; Lim, KL.;
Lin, FC.; Liou, W.; Liu, LF.; Lopez-Berestein, G.; Lopez-Otin, C.; Lu, B.; Macleod,
KF.;Malorni, W.; Martinet, W.; Matsuoka, K.; Mautner, J.; Meijer, AJ.; Melendez,
A.; Michels, P.; Miotto, G.; Mistiaen, WP.; Mizushima, N.; Mograbi, B.;
Monastyrska, I.; Moore, MN.; Moreira, PI.; Moriyasu, Y.; Motyl, T.; Munz, C.;
Murphy, LO.; Naqvi, NI.; Neufeld, TP.; Nishino, I.; Nixon, RA.; Noda, T.;
Nurnberg, B.; Ogawa, M.; Oleinick, NL.; Olsen, LJ.; Ozpolat, B.; Paglin, S.;
Palmer, GE.; Papassideri, I.; Parkes, M.; Perlmutter, DH.; Perry, G.; Piacentini, M.;
Pinkas-Kramarski, R.; Prescott, M.; Proikas-Cezanne, T.; Raben, N.; Rami, A.;
Reggiori, F.; Rohrer, B.; Rubinsztein, DC.; Ryan, KM.; Sadoshima, J.; Sakagami,
H.; Sakai, Y.; Sandri, M.; Sasakawa, C.; Sass, M.; Schneider, C.; Seglen, PO.;
Seleverstov, O.; Settleman, J.; Shacka, JJ.; Shapiro, IM.; Sibirny, A.; Silva-Zacarin,
EC.; Simon, HU.; Simone, C.; Simonsen, A.; Smith, MA.; Spanel-Borowski, K.;
Srinivas, V.; Steeves, M.; Stenmark, H.; Stromhaug, PE.; Subauste, CS.; Sugimoto,
S.; Sulzer, D.; Suzuki, T.; Swanson, MS.; Tabas, I.; Takeshita, F.; Talbot, NJ.;
Talloczy, Z.; Tanaka, K.; Tanida, I.; Taylor, GS.; Taylor, JP.; Terman, A.; Tettamanti,
G.; Thompson, CB.; Thumm, M.; Tolkovsky, AM.; Tooze, SA.; Truant, R.;
Tumanovska, LV.; Uchiyama, Y.; Ueno, T.; Uzcategui, NL.; van der Klei, I.;
Vaquero, EC.; Vellai, T.; Vogel, MW.; Wang, HG.; Webster, P.; Wiley, JW.; Xi, Z.;
Xiao, G.; Yahalom, J.;Yang, JM.; Yap, G.; Yin, XM.; Yoshimori, T.; Yu, L.; Yue, Z.;
Yuzaki, M.; Zabirnyk, O.; Zheng, X.; Zhu, X.; Deter, RL. (2008). Guidelines for the
use and interpretation of assays for monitoring autophagy in higher eukaryotes.
Autophagy, 4:151–175.
189
21. Koh, CH.; Peng, ZF.; Ou, K.; Melendez, A.; Manikandan, J.; Qi, RZ.; Cheung, NS.
(2007). Neuronal apoptosis mediated by inhibition of intracellular cholesterol
transport: microarray and proteomics analyses in cultured murine cortical neurons.
J. Cell Physiol, 211:63–87.
22. Koh, CH.; Qi, RZ.; Qu, D.; Melendez, A.; Manikandan, J.; Bay, BH.; Duan, W.;
Cheung, NS. (2006). U18666A-mediated apoptosis in cultured murine cortical
neurons: role of caspases, calpains and kinases. Cell Signal, 18:1572–1583.
23. Kwon, HJ.; Abi-Mosleh, L.; Wang, ML.; Deisenhofer, J.; Goldstein, JL.; Brown,
MS.; Infante, RE. (2009). Structure of N-terminal domain of NPC1 reveals distinct
subdomains for binding and transfer of cholesterol. Cell, 137: 1213–1224.
24. Lesser, G.; Kandiah, K.; Libow, LS.; Likourezos, A.; Breuer, B.; Marin, D.; Mohs,
R.; Haroutunian, V.; Neufeld, R. (2001). Elevated serum total and LDL cholesterol
in very old patients with Alzheimer's disease. Dement. Geriatr. Cogn. Disord.,
12(2):138-145.
25. Liao, G.; Yao, Y.; Liu, J.; Yu, Z.; Cheung, S.; Xie, A.;Liang, X.; Bi, X. (2007).
Cholesterol accumulation is associated with lysosomal dysfunction and autophagic
stress in Npc1 −/− mouse brain. Am. J. Pathol.,171:962–975.
26. Mallikarjun, K.; Bhayya, D.; Singh, D.; Shyagali, T. (2011). Metachromatic
leukodystrophy: a rare case report. J. Academy Adv Dental Research, 2(3).
27. Nassogne, M-C.; Commare, M-C.; Lellouch-Tubiana, A.; Emond, S.; Zerah, M.;
Caillaud, C.; Hertz-Pannier, L.; Saudubray, J-M. (2003). Unusual presentation of
GM2 gangliosidosis mimicking a brain stem tumor in a 3-year-old girl. AJNR Am J
Neuroradiol, 24:840–842.
28. Nelson, TJ.; Alkon, DL. (2005). Oxidation of cholesterol by amyloid precursor
protein and β-amyloid peptide. J. Biol. Chem., 280:7377-7387.
29. Notkola, I-L.; Sulkava, R.; Pekkanen, J.; Erkinjuntti, T.; Ehnholm, C.; Kivinen, P.;
Tuomilehto, J.; Nissinen, A. (1998). Serum total cholesterol, apolipoprotein E e4
allele, and Alzheimer's disease. Neuroepidemiology, 17:14-20.
30. Porter, F. (2008). Smith-Lemli-Opitz syndrome: pathogenesis, diagnosis and
management. European Journal of Human Genetics, 16:535-541.
31. Rosenbloom, B.; Weinreb, N. (2013). Gaucher disease: a comprehensive review.
Critical Reviews in Oncogenesis, 18(3):163-175.
32. Rotthier, A.; Auer-Grumbach, M.; Jassens, K.; Baets, J.; Almeida-Souza, L.; Van
Hoof, K.; Jacobs, A.; De Vriendt, E.; Schlotter-Weigel, B.; Loescher, W.;
Vondracek, P.; Seeman, P.; De Jonghe, P.; Van Dijck, P.; Jordanova, A.; Hornemann,
T.; Timmerman, P. (2010). Mutations in the SPTLC2 subunit of serine
palmitoyltransferase cause hereditary sensory and autonomic neuropathy type1.
Am J Hum Genet, 87:513-522.
33. Rotthier, A.; Baets, J.; De Vriendt, E.; Jacobs, A.; Auer-Grumbach, M.; Levy, N.;
Bonello-Palot, N.; Kilic, SS.; Weis, J.; Nascimento, A.; Swinkels, M.; Kruyt, MC.;
Jordanova, A.; De Jonghe, P.; Timmerman, V. (2009). Genes for hereditary sensory
and autonomic neuropathies: a genotype-phenotype correlation. Brain, 132:2699-
2711.
34. Rotthier, A.; Penno,A.; Rautenstrauss, B.; Auer-Grumbach, M.; Stettner, G.;
Asselbergh, B.; Van Hoof, K.; Sticht, H.; Levy,N.; Timmerman, V.; Hornemann, T.;
Janssens, K. (2011). Characterization of two mutations in the SPTLC1 subunit of
serine palmitoyltransferase associated with hereditary sensory and autonomic
190
neuropathy type1. HUMAN MUTATION Mutation in Brief, 32:E2211-E2225.
35. Saouab, R.; Mahi, M.; Abilkacem, R.; Boumdin, H.; Chaouir, S.; Agader, O.; Amil,
T.; Hanine, A. (2011). A Case Report of Sandhoff Disease. Clin Neuroradiol, 21:83-
85.
36. Schnabel, D.; Schröder, M.; Furst, W.; Klein, A.; Hurwitz, R.; Zenk, T.; Weber, J.;
Harzer, K.; Paton, BC.; Poulos, A.; Suzuki, K.; Sandhoff, K. (1992). Simultaneous
deficiency of sphingolipid activator proteins 1 and 2 is caused by a mutation in the
initiation codon of their common gene. Journal of Biochemical Chemistry,
267:3312–3315.
37. Siddiqi, ZA.; Sanders, DB.; Massey, JM. (2006). Peripheral neuropathy in Krabbe
disease: effect of hematopoietic stem cell transplantation. Neurology, 67:268–272.
38. Stepein, K.; Hendriksz, C. (2017). Lipid profile in adult patients with Fabry
disease-ten year follow up. Molecular Genetics and Metabolism Reports, 13:3-6.
39. Suzuki, K. (1998). Twenty five years of the “psychosine hypothesis”: a personal
perspective of its history and present status. Neurochem Res, 23:251–259.
40. te Vruchte, D.; Lloyd-Evans, E.; Veldman, RJ.; Neville, DC.; Dwek, RA.; Platt,
FM.; van Blitterswijk, WJ.; Sillence, DJ. (2004). Accumulation of
glycosphingolipids in Niemann-Pick C disease disrupts endosomal transport. J.
Biol. Chem., 279:26167– 26175.
41. Tulenko, TN.; Boeze-Battaglia, K.; Mason, RP. Tint, GS.; Steiner, RD.; Connor,
WE.; Labelle, EF. (2006). A membrane defect in the pathogenesis of the Smith-
Lemli-Opitz syndrome. J Lipid Res, 47(1):134-143.
42. Varki, A.; Cummings, R.; Esko,J.; Stanley, P.; Hart, G.; Aebi, M.; Darvill,A.;
Kinoshita, T.; Packer, N.; Prestegard, J.; Schnaar, R.; Seeberger, P. (2017).
Essentials of glycobiology. 3rd edition. Cold Spring Harbor Laboratory Press. New
York.
43. Verma, N. (2017). Introduction to hyperlipidemia and its treatment: a review.
International Journal of Current Pharmaceutical Research, 9(1):6-14.
44. Wang, C.; Melberg, A.; Weis, J.; Månsson, JE.; Raininko, R. (2007). The earliest
MR imaging and proton MR spectroscopy abnormalities in adult-onset Krabbe
disease. Acta Neurol Scand, 116:268–272.
45. Wu, YP.; Mizukami, H.; Matsuda, J.; Saito, Y.; Proia, RL.; Suzuki, K. (2005).
Apoptosis accompanied by up-regulation of TNF-alpha death pathway genes in the
brain of Niemann-Pick type C disease. Mol. Genet. Metab.,84:9–17.
46. Xu, Q.; Huang, Y. (2006). Lipid metabolism in Alzheimer's and Parkinson's
disease. Future Lipidology, 1(4):441-453.
47. Xu, YH.; Barnes, S.; Sun, Y.; Grabowski, G. (2010). Multi-system disorders of
glycosphingolipid and ganglioside metabolism. J Lipid Res, 51(7):1643-1675.
48. Yang, W.; Shi, H.; Zhang, J.; Shen, Z.; Zhou, G.; Hu, M. (2017). Effects of the
duration of hyperlipidemia on cerebral lipids, vessels and neurons in rats. Lipids
Health Dis, 16(26).
49. Zambon, D.; Quintana, M.; Mata, P.; Alonso, R.; Benavent, J.; Cruz-Sanchez, F.;
Gich, J.; Pocovi, M.; Civeira, F.; Capurro, S.; Bachman, D.; Sambamurti, K.;
Nicholas, J.; Pappolla, M. (2010). Higher incidence of mild cognitive impairment
in familial hypercholesterolemia. Am J Med, 123(3):267-274.

191

You might also like