You are on page 1of 111

COLLEGE OF SCIENCE AND TECHNOLOGY

Module Title: Animal Physiology


Code: BIO 2164

Academic Year: 2019

Year: Y2 Biotechnology

Lecturer: Mrs. Mukamugema Jane


1

COURSE CONTENT content

Chapter 1. INTRODUCTION TO ANIMAL PHYSIOLOGY.............................................................3


Chapter 2: ENERGY AND CELLULAR RESPIRATION..........................................................................4
2.1. Catalysts and energy flow mechanism..............................................................................................4
2.2. Overview of cellular respiration.......................................................................................................5
2.3. Stages of cellular respiration.............................................................................................................6
2.4. Anaerobic respiration......................................................................................................................14
Chapter 3. NERVOUS SYSTEM AND SENSORY RECEPTION...........................................................16
3.1. Nervous system and nerve cells......................................................................................................16
3.2. Divisions of the nervous system.....................................................................................................19
3.3. Sensory reception...........................................................................................................................25
Chapter 4. ENDOCRINE GLANDS AND FUNCTIONS.........................................................................27
4.1. Body metabolism regulation...........................................................................................................27
4.2. Mode of chemical regulatory signals..............................................................................................28
4.3. Hormone chemical forms and solubility.........................................................................................31
4.4. Major endocrine glands and hormones...........................................................................................32
4.5. Endocrine and nervous coordination by hypothalamus...................................................................35
4.6. Hormone function mechanisms......................................................................................................38
Chapter 5. REPRODUCTIVE PROCESS AND REGULATION.............................................................42
5.1. Sexual reproduction........................................................................................................................42
5.2. The origin and maturation of germ cells.........................................................................................42
5.3. Male reproductive system...............................................................................................................46
5.4. Female reproductive system...........................................................................................................47
5.5. Endocrine regulation in reproduction..............................................................................................49
Chapter 6. SKELETON, MUSCLE, AND MOTION................................................................................55
6.1. Animal skeletons and their roles.....................................................................................................55
6.2. Skeletal muscle structure and function...........................................................................................59
6.3. Interaction of muscles and skeleton for movement.........................................................................63
Chapter 7. DIGESTIVE PROCESS AND NUTRITION...........................................................................65
7.1. Food capture and mastication.........................................................................................................65
7.2. Mechanisms of digestion................................................................................................................66
7.3. Organization of the alimentary canal..............................................................................................67
2

7.4. Regulation of food intake and digestion.........................................................................................76


7.5. Nutritional requirements.................................................................................................................77
Chapter 8. BLOOD CIRCULATION AND FUNCTION..........................................................................82
8.1. Blood composition..........................................................................................................................82
8.2. Open and closed circulations..........................................................................................................84
8.3. Mammalian heart............................................................................................................................85
8.4. Prevention of blood loss.................................................................................................................87
8.5. Blood groups and transfer...............................................................................................................89
8.6. Lymphatic system...........................................................................................................................90
Chapter 9. RESPIRATORY SYSTEM AND BREATHING.....................................................................91
9.1. Oxygen transport in animal respiration...........................................................................................91
9.2. Respiratory structures or organs.....................................................................................................91
9.3. Mammalian respiratory system.......................................................................................................92
9.4. Transport of respiratory gases.........................................................................................................95
Chapter 10. EXCRETION AND HOMEOSTASIS...................................................................................97
10.1. Water and osmotic regulation.......................................................................................................97
10.2. Salt and water balance in terrestrial animals.................................................................................97
10.3. Vertebrate kidney structure and function......................................................................................98
10.4. Physiological adaptations to temperature....................................................................................103
3

Chapter 1. INTRODUCTION TO ANIMAL PHYSIOLOGY

Physiology : is the branch of biological sciences dealing with understanding life functions
including all physical and chemical processes. Animal physiology is the study of how animals
work, or more specifically the physical and chemical processes that occur within animals.
Examples of these processes include gas exchange, blood circulation, osmoregulation, digestion,
nervous and muscle systems and endocrinology. The bodies of animal organisms are built in
different ways and various functions are performed according to their levels of complexity.
Unicellular protozoa groups, including organisms such as malarial and trypanosome parasites,
are the simplest animal-like organisms; however they are complete organisms that perform all of
the basic functions of life as seen in the more complex animals. Such functions are mainly seven:
nutrition, respiration, growth, locomotion, excretion, sensitivity, and reproduction.

Though incapable of independent existence, cells of a multicellular organism are specialized for
performing the various tasks accomplished by the body as a whole. Similar cells are grouped
together in tissues in order to perform a common function. Tissues are assembled into still larger
functional units called organs. Different organs operate together as systems. There are eleven
different kinds of organ systems are observed in metazoan animals: skeletal, muscular,
integumentary, digestive, respiratory, circulatory, excretory, nervous, endocrine, immune, and
reproductive. Metazoan animals-are multicellular, mitochondrial eukaryotes. Today Metazoan
encompasses all animals with differentiated tissues, including nerves and muscles. Most of those
systems will be discussed throughout this part of Animal physiology.

Depending on the classification approach, some systems can be grouped together due to their
functional closeness, and sometimes different terms are used. For instance the lymphatic system
is often accounted in circulatory system or as a component of the immune system; reproductive
and urine excretory system can be grouped together into a single urogenital system, while
sometimes urinary system is strictly used for the excretory system.
4

Chapter 2: ENERGY AND CELLULAR RESPIRATION

2.1. Catalysts and energy flow mechanism

Most metabolic reactions are activated or catalyzed by enzymes. An enzyme is a


macromolecule .protein in nature that acts as a catalyst. A catalyst is chemical agent that speeds
up a reaction without being consumed by the reaction. In the absence of regulation by enzymes,
chemical traffic through the pathways of metabolism would become terribly congested because
many chemical reactions would take such a long time.

The special catalytic talent of an enzyme is its power to reduce the amount of activation energy,
the energy required to make molecules of a substance take part in chemical reaction. In fact,
given one or more enzymes, catalyzed reactions proceed through one or more intermediate steps.
Body temperature is not high enough to initiate burning of glucose. Instead, if you swallow some
glucose, enzymes in your cells will lower the barrier of activation energy, allowing the sugar to
be oxidized in a series of steps.

Enzymes affect only the reaction rate, and a specific enzyme catalyzes only one reaction. An
enzyme functions by associating in a highly specific way with its substrate, the molecule whose
reaction it catalyzes. One of the most distinctive attributes of enzymes is their high specificity, a
consequence of the exact molecular fit that is required between enzyme and substrate. The
substrate is held within the active site of an enzyme, and when the reaction is completed the
products is released and the enzyme’s active site is free to take another substrate molecule into
its active site. Below is Illustration of enzyme and substrate

An enzyme can convert one or more reactant


molecules to one or more product molecules.
Active site can lower EA and speed up a reaction
by 1) Active site orients substrates in correct
position for reaction, 2) stressing the substrates
and stabilizing the transition state, 3) providing
a favorable microenvironment, and/or 4)
participating directly in the catalytic reaction.
5

Adenosine triphosphate (ATP) - is a complex organic chemical that provides energy to drive
many processes in living cells, e.g. muscle contraction, nerve impulse propagation, and chemical
synthesis etc. The ATP molecule consists of adenosine (the purine adenine and the 5-carbon
sugar ribose) and a triphosphate group. Most of the free energy in ATP resides in the
triphosphate group, especially in two phosphoanhydride bonds between the three phosphate
groups. These two bonds are called “high-energy bonds” because a great deal of free energy in
the bonds is liberated when ATP is hydrolyzed to adenosine diphosphate (ADP) and inorganic
phosphate.

Structure of ATP

ATP +H2O ADP + Pi


where Pi represents inorganic
phosphate

The high-energy bonds of ATP are actually rather weak, unstable bonds. Because they are
unstable, the energy of ATP is readily released when ATP is hydrolyzed in cellular reactions.
ATP is produced by one set of reactions and almost immediately consumed by another. ATP is
formed as it is needed, primarily by oxidative processes in the mitochondria.

2.2. Overview of cellular respiration

Cellular respiration is the process in which cells produce the energy they need to survive. In


cellular respiration, cells use oxygen to break down the sugar and store its energy in form of
adenosine triphosphate (ATP). Cells use ATP to power virtually all of their activities such as to
grow, divide, replace worn out cell parts, and to execute many other tasks.

Cellular respiration sometimes is referred to as aerobic respiration, meaning that it occurs in the
presence of oxygen. Aerobic respiration is in principle similar to the combustion of gasoline in
an automobile engine after oxygen is mixed with the fuel (hydrocarbons). Food provides the fuel
6

for respiration and the exhaust is carbon dioxide and water. Cellular respiration transfers about
40 percent of the energy of glucose to ATP. The rest of the energy from glucose is released as
heat, which warm-blooded organisms use to maintain body temperature, and cold-blooded
organisms release to the atmosphere.

Glucose is degraded to pyruvic acid by cytoplasmic enzymes (glycolytic pathway). Acetyl


coenzyme A is formed from pyruvic acid and is fed into the Krebs cycle. An acetyl group (two
carbons) is oxidized to two molecules of carbon dioxide with each turn of the cycle. Pairs of
electrons are removed from the carbon skeleton of the substrate at several points in the pathway
and are carried by oxidizing agents NADH or FADH 2 to the electron transport chain where 32
molecules of ATP are generated. Four molecules of ATP are also generated by substrate
phosphorylation in the glycolytic pathway, and two molecules of ATP (initially GTP) are formed
in the Krebs cycle. This yields a total of 38 molecules of ATP (36 molecules net) per glucose
molecule. Molecular oxygen is involved only at the very end of the pathway.

Some organisms cannot produce ATP through this process because they live in environments
that lack oxygen. These organisms, typically bacteria, rely on anaerobic processes such as
fermentation to generate their ATP. Fermentation is a partial degradation of sugars that occurs
without the use of oxygen. Aerobic metabolism is more familiarly known as true cellular
respiration.

2.3. Stages of cellular respiration

The complete oxidation of glucose as the typical food molecule passes through 3 main stages,
namely Glycolysis, the Krebs cycle, and the Electron transport chain.

A. Glycolysis

Glucose, the primary fuel used in glycolysis, is an organic molecule consisting of 6 carbon, 12


hydrogen, and 6 oxygen atoms. Glycolysis is the anaerobic degradation of glucose to yield
pyruvate, forming initial process before the aerobic phase of respiration, which usually occurs in
mitochondria. Cells without mitochondria rely on glycolysis for most of their ATP synthesis, as
do facultatively anaerobic cells (e.g. striated muscle fibres) when there is a shortage of oxygen.
7

In glycolysis, glucose is broken down with the help of enzymes and other molecules found in the
cytoplasm. Two phosphate groups are attached to glucose, then glucose is split in half to produce
two three-carbon molecules, each with one phosphate group attached. Later, one hydrogen atom
and two electrons are removed from each three-carbon molecule to be transferred as a unit to a
large molecule called nicotinamide adenine dinucleotide (NAD+).

In the final steps of glycolysis, two hydrogen atoms are removed from each three-carbon
compound and bond with oxygen atoms in the cytoplasm to form water. Another enzyme
removes the phosphate group which bonds to a single molecule of adenosine diphosphate (ADP).
In this step, two new ATP molecules are produced. When the two three-carbon compounds are
separated from the phosphate groups, the three-carbon compounds are converted to two
molecules of pyruvate, each composed of 3 carbon, 3 oxygen, and 3 hydrogen atoms.

The pathway generates a net gain of two molecules of ATP per molecule of glucose used, plus
reducing power in the form of two NADH, molecules. NADH serves as a carrier molecule to
convey high-energy electrons to the final electron transport chain, where ATP will be produced.

The energy input and output of glycolysis

Glucose 2 Pyruvate
4 ATP formed – 2 ATP used 2 ATP
2 NAD+ + 4 eˉ + 4 H+
2 NADH + 2 H+
All of the carbon originally present in glucose is accounted for in the two molecules of pyruvate;
no CO2 is released during glycolysis.
8

Series of reactions in glycolysis

Glucose Glyceraldehyde
-3-phosphate
ATP
NAD Inorganic
ADP phosphate
NADH2
Glucose-6-
1,3-diphospho
phosphate
glyceric acid

ADP

ATP
Fructose-6-
phosphate 3-phospho
glyceric acid
ATP

ADP
2-phospho
Fructose-1,6-
glyceric acid
diphosphate
ADP

ATP

Phosphoenol
Dihydroxy Glyceraldehyde- pyruvic acid
acetone- 3-phosphate
phosphate NADH2 NAD

Pyruvic acid Lactic acid

Transition stage: formation of acetyl coenzyme A

The pyruvate molecules move from the cytoplasm to special structures in the cell called
mitochondria, where the remaining steps of cellular respiration are carried out. Each
mitochondrion contains a membrane that is folded back and forth many times and contains
hundreds of thousands of enzymes that direct cellular respiration.

The transition stage is a short biochemical pathway that links glycolysis with the Krebs cycle. In
this brief stage, enzymes called hydrogenases transfer hydrogens and electrons from the two
9

pyruvate molecules to two molecules of NAD+ to form two more molecules of NADH. Another
enzyme breaks off one carbon and two oxygen atoms from each pyruvate molecule. These atoms
combine to form carbon dioxide, the primary waste product of cellular respiration, which
diffuses out of the cell. Each pyruvate molecule is then transformed into a two-carbon compound
called an acetyl group which unites with two molecules of coenzyme A to form two acetyl
coenzyme A molecules that enter the Krebs cycle.

C. Krebs Cycle

The Krebs cycle (citric acid cycle, Tricarboxylic acid cycle, TCA cycle) is a series of cyclical
chemical reactions carried out within cells to convert molecules from food into carbon dioxide,
water, and energy. The reactions themselves contribute little energy to the cell, but
dehydrogenations involved are a source of electrons for Electron Transport Systems via which
ATP is produced from ADP and inorganic phosphate. In addition, some GTP (Guanosine
triphosphate) is produced directly during the cycle. Cells with mitochondria perform the cycle by
means of enzymes within the matrix bounded by the inner mitochondrial membrane.

At the start of the cycle, an acetyl group combines with a four-carbon molecule called
oxaloacetate to yield a six-carbon compound, citric acid. In the remaining steps of the cycle, the
citric acid molecule is gradually rearranged and stripped of two of its carbon atoms, which are
given off in the form of carbon dioxide. Four electrons are also released. These electrons travel
along a nearby series of carrier molecules in the cell, the electron transport chain.
10

Series of reactions in the Krebs cycle

Two
molecules of CO2 are created from the oxidation of one pyruvate. At one stage in the cycle,
hydrogens and electrons are transferred to a molecule of flavin adenine dinucleotide (FAD++) to
form FADH2, a molecule like NADH that temporarily stores hydrogen and electrons for later
use. By the end of the Krebs cycle, most of the usable energy from the original glucose molecule
has been transferred to ten molecules of NADH (two from glycolysis, two from the transition
stage, and six from the Krebs cycle); two molecules of FADH2; and four molecules of ATP, two
of which were formed in glycolysis.
11

D. Electron Transport Chain

The electron transport chain is a collection of molecules embedded in the inner membrane of the
mitochondrion in eukaryotic cells (in prokaryotes, they reside in the plasma membrane). Most
components of the chain are proteins, which exist in multiprotein complexes numbered I through
IV. Two of the commonly known electron carriers, ubiquinone (Q) and cytochrome C (Cyt C),
move rapidly, transferring electrons between the large protein complexes.

The energy released at each step of the chain is stored in a form which the mitochondrion (or
prokaryotic cell) can use to make ATP. This mode of ATP synthesis is called oxidative
phosphorylation because it is powered by the redox reactions of the electron transport chain.
NADH and FADH2 molecules release their load of electrons and hydrogen ions near these
molecules. The electrons are passed down the chain until they reach oxygen, the final molecule
in the chain. The combination of the electrons, hydrogen ions, and oxygen forms water, used by
the cell in other biochemical reactions.

At the end, NADH and FADH2 are converted back to NAD+ and FAD++ respectively, providing
the cell with a steady supply of these molecules so that cellular respiration can be carried out
over and over again. In most cells, the electron transport chain produces 32 molecules of ATP.
Together with the two ATP molecules gained in glycolysis and the two generated in the Krebs
cycle, cellular respiration produces a grand total of 36 molecules of ATP for every molecule of
glucose.

In eukaryotic cells, the inner membrane of the mitochondrion is the site of electron transport and
chemiosmosis, the processes that together constitute oxidative phosphorylation. Chemiosmosis is
that process in which energy stored in the form of a hydrogen ion gradient across a membrane is
used to drive cellular work such as the synthesis of ATP, and it refers to the flow of H + across a
membrane rather than water or solute transport. Small amount of ATP is formed by substrate-
level phosphorylation, a mode of ATP synthesis by which an enzyme transfers a phosphate
group from a substrate molecule to ADP, rather than adding an inorganic phosphate to ADP.
12

Oxidative phosphorylation and ATP synthesis

Energy metabolism from glycerol, fats, and proteins

Although glucose is the primary fuel for cellular respiration, cells can rely on other molecules to
produce ATP. The cellular respiration pathway is connected to other metabolic pathways that can
donate molecules to cellular respiration at different steps along the way. For example, glycerol, a
breakdown product of fat, can enter the cellular respiration pathway in the middle of glycolysis.
Another product of fat digestion, fatty acids, can enter at the transition stage. Glycerol is
modified in glycolysis to pyruvate, and fatty acids are modified to acetyl coenzyme A in the
transition stage. The pyruvate and acetyl coenzyme A are processed through the remaining steps
of cellular respiration to yield ATP. The ability to use alternate molecules enables cells to keep
ATP production going even if they run out of glucose. Below are the illustrations (Diagrams)
that helps to understand.
13

Pathway for oxidation of glucose and other carbohydrates


14
15
16

2.4. Anaerobic respiration

There are two general mechanisms by which certain cells can oxidize organic fuel and generate
ATP without the use of oxygen and the distinction between these two is based on whether an
electron transport chain is present.

2.4.1. Marine bacteria

Anaerobic respiration takes place in certain prokaryotic organisms that live in environments
without oxygen. These organisms have an electron transport chain but do not use oxygen as a
final electron acceptor at the end of the chain. Oxygen performs this function very well because
it is extremely electronegative, but other, less electronegative substances can also serve as final
electron acceptors. Some sulfate-reducing marine bacteria for instance use the sulfate ion (SO 42-)
at the end of their respiratory chain. Operation of the chain builds up a proton-motive force used
to produce ATP, but H2S (hydrogen sulfide) is produced as a byproduct rather than water.

2.4.2. Fermentation

Fermentation is a way of harvesting chemical energy without using either oxygen or any electron
transport chain. As an alternative to respiratory oxidation of organic nutrients, fermentation is an
expansion of glycolysis that allows continuous generation of ATP by the substrate-level
phosphorylation of glycolysis.

Fermentation consists of glycolysis plus reactions that regenerate NAD + by transferring electrons
from NADH to pyruvate or derivatives of pyruvate. The NAD + can then be reused to oxidize
sugar by glycolysis, which nets two molecules of ATP by substrate-level phosphorylation. There
are many types of fermentation, differing in the end products formed from pyruvate. Two
common types are alcohol fermentation and lactic acid fermentation.

a) Alcohol fermentation

In alcohol fermentation, pyruvate is converted to ethanol (ethyl alcohol) in two steps. The first
step releases carbon dioxide from the pyruvate, which is converted to the two-carbon compound
acetaldehyde. In the second step, acetaldehyde is reduced by NADH to ethanol. This regenerates
the supply of NAD+ needed for the continuation of glycolysis.
17

Many bacteria carry out alcohol fermentation under anaerobic conditions. Yeast (a fungus) also
carries out alcohol fermentation. For thousands of years, humans have used yeast in brewing,
winemaking, and baking. The CO2 bubbles generated by baker's yeast during alcohol
fermentation allow bread to rise.

b) Lactic acid fermentation

During lactic acid fermentation, pyruvate is reduced directly by NADH to form lactate as an end
product, with no release of CO2 (Lactate is the ionized form of lactic acid.) Lactic acid
fermentation by certain fungi and bacteria is used in the dairy industry to make cheese and
yogurt.

Human muscle cells make ATP by lactic acid fermentation when oxygen is scarce. This occurs
during the early stages of strenuous exercise, when sugar catabolism for ATP production
outpaces the muscle's supply of oxygen from the blood. Under these conditions, the cells switch
from aerobic respiration to fermentation. The lactate that accumulates was previously thought to
cause muscle fatigue and pain, but recent research suggests instead that increased levels of
potassium ions (K+) may be to blame, while lactate appears to enhance muscle performance. In
any case, the excess lactate is gradually carried away by the blood to the liver. Lactate is
converted back to pyruvate by liver cells.

(a) Alcohol fermentation (b) lactic acid fermentation


16

Chapter 3. NERVOUS SYSTEM AND SENSORY RECEPTION

3.1. Nervous system and nerve cells

3.1.1. Neurons and glial cells

The nervous system is specialized for reception of stimuli and conduction of impulses
(information) from one region of the body to another. Two basic types of cells in nervous tissue
are neurons, the basic functional unit of the nervous system, and neuroglia, a variety of non-
nervous cells that insulate neuron membranes and serve in various supportive functions.

Structure of a motor (efferent) neuron

A neuron, or nerve cell, may assume many shapes, depending on its function and location. From
the nucleated cell body or soma extend one or more dendrites, which receive electrical impulses
from receptors or other nerve cells, and a single axon that carries electrical impulses away from
the cell body to other nerve cells or to an effector organ on synaptic terminals.
17

In vertebrates and some complex invertebrates, the axon is often covered with an insulating
sheath of myelin. Nodes of Ranvier are interruptions in Schwann cell’s insulation that allow
action potentials to leap from node to node. The axon is often called a nerve fiber. In vertebrates,
nerve axons are often bundled together and wrapped with connective tissue to form a nerve. Cell
bodies of these nerve processes are located either in the central nervous system or in ganglia
(discrete bundles of nerve cell bodies located outside the central nervous system).

Neurons are commonly classified as afferent, or sensory; efferent, or motor; and interneurons,
which are neither sensory nor motor but connect neurons with other neurons. Interneurons,
which in humans make up 99% of all neurons in the body, lie entirely within the central nervous
system (CNS). Afferent neurons are connected to receptors. Receptors function to convert
environmental stimuli into nerve impulses, which are carried by the afferent neurons into the
CNS. Impulses also move to efferent neurons, which carry them via the peripheral system to
effectors, such as muscles or glands.

Glial cells (glia, neuroglia) are non-conducting nerve cells, performing supportive and protective
roles for neurons. They include astrocytes, oligodendrocytes, Schwann cells, microglia and
ependyma cells. Astrocytes attach neurons to blood vessels and serve as nutrient and ion
reservoirs for neurons, as well as a scaffold during brain development.

Astrocytes, and smaller microglial cells, are essential for the regenerative process that follows
brain injury. Unfortunately, astrocytes also participate in several diseases of the nervous system,
including Parkinsonism and multiple sclerosis. Oligodendrocytes form myelin sheaths of axons
of central nervous system and Schwann cells form myelin sheaths outside CNS. Ependyma cells
line ventricles of the brain and cerebrospinal canal.

3.1.2. Conduction of nervous impulse

Axons achieve high conduction velocities by a cooperative relationship between axons and the
myelin sheath laid down by the Schwann cells. Insulating myelin sheaths are interrupted at
intervals by nodes (called nodes of Ranvier) where the surface of the axon is exposed to fluid
surrounding the nerve. In these myelinated fibers the action potential depolarizes the axon
membrane only at the nodes because the myelin sheath prevents depolarization elsewhere.
18

The ion pumps (Na+-K+ pump) and channels that move ions across the membrane are
concentrated in each node. Once an action potential starts down an axon, depolarization of the
first node initiates an electrical current that stretches out to the neighboring node, causing it to
depolarize and trigger an action potential. Thus the action potential leaps from node to node, a
kind of conduction called saltatory. The gain in efficiency as compared with non-myelinated
fibers is impressive. For example, a frog myelinated axon only 12 μm in diameter conducts nerve
impulses at the same speed as a squid axon 350 μm in diameter.

Impulse conduction in unmyelinated and myelinated fibers

In unmyelinated fibers (A), the action potential


spreads continuously along the entire length of the
axon. In myelinated fibers (B), the action potential
leaps from node to node, bypassing the insulated
portions of the fiber. This is saltatory conduction,
which is much faster than continuous conduction.

3.1.3. Synapses: junctions between nerves

Synapses are specialized junctions that set separation between nerves or from effector organs.
When an action potential passes down an axon to its terminal, it must cross a small gap, the
synapse, separating it from another neuron or an effector organ. Two distinct kinds of synapses
are known: electrical and chemical.

More complex than electrical synapses are chemical synapses, which contain packets of
specialized chemicals called neurotransmitters. Neurons bringing impulses toward chemical
synapses are called presynaptic neurons; those carrying impulses away are postsynaptic
neurons. At a synapse, membranes are separated by a narrow gap, the synaptic cleft, having a
width of approximately 20 nm.
19

Axon terminals and synapse

3.2. Divisions of the nervous system

The nervous system of vertebrates has two main divisions: central and peripheral. The central
nervous system is composed of the brain and spinal cord and is the site of information
processing. The peripheral nervous system is composed of all the nerves of the body outside the
brain and spinal cord.

Central and peripheral NS


20

3.2.1. Central nervous system

3.2.1.1. Brain

Anatomically, the vertebrate brain develops at the anterior end of the spinal cord. During
embryonic development, the brain undergoes regional expansion as a hollow tube of nervous
tissue forms and develops into the hindbrain, midbrain, and forebrain.

Table illustrating divisions of the vertebrate brain


Embryonic vesicle Main components
Function
Early embryo Late embryo in adults
Motor area controls voluntary muscle movements; sensory
cortex is center of conscious perception of touch, pressure,
vibration, pain, temperature, and taste; association areas
Telencephalon Cerebrum integrate and process sensory data
Forebrain Part of limbic system; integrates sensory information arriving
(Prosencephalon) Thalamus at thalamus, projects to cerebral frontal lobes
Diencephalon Controls autonomic functions; sets appetitive drives (thirst,
hunger, sexual desire) and reproductive behavior; participates
in emotional responses; secretes ADH, oxytocin; secretes
Hypothalamus releasing hormones for anterior pituitary regulation
Optic lobes Integrates visual information with other sensory inputs; relays
Midbrain (tectum) auditory information
Mesencephalon
(Mesencephalon) Involuntary control of muscle tone; processing of incoming
Midbrain nuclei sensations and outgoing motor commands
Involuntary coordination and control of outgoing movements
Cerebellum for equilibrium, muscle tone, posture
Metencephalon Links cerebellum with other brain centers and with medulla
Hindbrain and spinal cord; modifies output of respiratory centers in
(Rhombencephalon) Pons medulla
Regulates heart rate and force of contraction; vasomotor
Medulla control; sets rate of respiration; relays information to the
Myelencephalon oblongata cerebellum

a) Hindbrain

The medulla, a conical continuation of the spinal cord, together with the more anterior midbrain,
constitutes the brainstem an area that controls numerous vital and largely subconscious activities
such as heartbeat, respiration, vascular tone, gastric secretions, and swallowing. The pons, also a
part of the hindbrain, contains a thick bundle of fibers that carry impulses from one side of the
cerebellum to the other, and also connects both medulla and cerebellum to other brain regions.
21

The cerebellum lying dorsal to the medulla controls equilibrium, posture, and movement. Its
development is directly correlated with the animal’s mode of locomotion, agility of limb
movement, and balance. Unlike in other animals, the cerebellum is greatly expanded and folded
in birds and mammals. Primates and especially humans have the most complex cerebellum due
to a most complex manual dexterity.

b) Midbrain

The midbrain consists mainly of the tectum (including the optic lobes). The roof of the midbrain,
the optic tectum, is a thickened region of gray matter that integrates visual and auditory signals.
The midbrain in vertebrates has taken on added functions relating to tactile and auditory input. In
mammals, the midbrain is mainly a relay center for information on its way to higher brain
centers.

c) Forebrain

Just anterior to the midbrain lie the thalamus and hypothalamus, the most posterior elements of
the forebrain. The thalamus is a major relay station that analyzes and passes sensory information
to higher brain centers. In the hypothalamus are several centers that regulate body temperature,
water balance, appetite, hunger, and thirst all functions concerned with maintenance of internal
constancy (homeostasis).

Neurosecretory cells located in the hypothalamus produce several neurohormones. The


hypothalamus also contains centers for regulating reproductive function and sexual behavior, and
it participates in emotional behaviors. In the posterior part of the fore brain are also located the
pineal gland and pituitary gland. The pineal gland, or epiphysis, is an endocrine gland that
controls some body rhythms. The pituitary, or hypophysis, is a major endocrine gland.

The anterior portion of the forebrain is the cerebrum. Its deep section includes the limbic system
that mediates several species-specific behaviors that relate to fulfilling needs such as feeding and
sex. One region of the limbic system, the hippocampus, has been extensively studied as a site
involved with spatial learning and memory. The roof (outermost section) called cerebral cortex,
contains discrete motor and sensory areas and large association areas, concerned with memory,
judgment, reasoning, and other integrative functions. This part has progressively increased in
22

size and complexity in mammals, so that it has folded back on itself to a remarkable extent. It
contains primary sensory and motor areas. Other areas of the cortex are involved in the
perception of visual or auditory signals from the environment. In humans, this includes the
ability to use language, both written and spoken.

The cerebrum is also divided into right and left cerebral hemispheres. The right and left
hemispheres of the cerebral cortex are bridged through the corpus callosum, a neural connection
through which the two hemispheres are able to transfer information and coordinate mental
activities. In humans, the two hemispheres of the brain are specialized for entirely different
functions: the left hemisphere for language development, mathematical and learning capabilities,
and sequential thought processes; and the right hemisphere for spatial, musical, artistic, intuitive,
and perceptual activities. Each hemisphere also controls the opposite side of the body.

In mammals, and especially in humans, separate parts of the brain mediate conscious and
unconscious functions. The unconscious mind, all of the brain except the cerebral cortex,
governs numerous vital functions that are removed from conscious control: respiration, blood
pressure, heart rate, hunger, thirst, temperature balance, salt balance, sexual drive, and basic
(sometimes irrational) emotions. It is also a complex endocrine gland that regulates and receives
feedback from the body’s subservient endocrine system. The conscious mind, the cerebral
cortex, is the site of higher mental activities (e.g. planning and reasoning), memory, and
integration of sensory information.
23

3.2.1.2. Spinal cord

The spinal cord is the part of the CNS that extends from the brain to near or into the tail. The
spinal cord serves two important functions in an animal: it is the connecting link between the
brain and most of the body, and it is involved in spinal reflex actions. A reflex is an automatic
involuntary response or reaction of the nervous system to a stimulus. An example is when
somebody moves quickly the hand when accidently touching a hot object, or feels an insect bite.

A cross section in spinal cord shows a neural canal, which contains a cerebrospinal fluid,
surrounded by white and gray matter. The gray matter consists of cell bodies and dendrites, and
is concerned mainly with reflex connections at various levels of the spinal cord. The ventral and
dorsal roots of the spinal nerves contain the main motor and sensory fibers (axons and/or
dendrites), respectively. Notice that the relative position of gray and white matter is contrasted
between brain and spinal cord. In brain the gray matter is peripheral and white matter central.

Simple reflex arc

The white matter of the spinal cord gets its name from the whitish myelin that covers the axons.
Continuous with similar ones that cover the brain, 3 layers of protective membranes (meninges)
surround the spinal cord. The outer layer, the dura mater, is a tough, fibrous membrane. The
middle layer, the arachnoid, is delicate and connects to the innermost layer, the pia mater that
contains small blood vessels that nourish the spinal cord.
24

3.2.2. Peripheral nervous system

The nerves that make up the PNS are commonly divided into two groups: sensory or afferent
nerves, which transmit sensory information to the CNS; and motor or efferent nerves, which
convey motor commands from CNS to muscles and glands. The motor nerves divide into the
somatic nervous system (voluntary), which innervates and relays commands to skeletal muscles,
and the visceral or autonomic nervous system (involuntary) , which stimulates other muscles
(smooth and cardiac) and glands of the body. The nerves of the autonomic nervous system divide
into sympathetic and parasympathetic systems.

CNS and nerves

Nerves are also classified considering the centers where they take origin. Cranial nerves are
nerves that emerge directly from the brain and brainstem. This is in contrast to spinal nerves
which emerge from various segments of the spinal cord. The 12 pairs of cranial nerves in
reptiles, birds, and mammals are: Olfactory, Optic, Oculomotor, Trochlear, Trigeminal
(Ophthalmic division, Maxillary division, Mandibular division), Abducens, Facial,
Vestibulocochlear (Vestibular branch, Cochlear branch), Glossopharyngeal, Vagus, Accessory
(Cranial branch, Spinal branch), and Hypoglossal. Fishes and amphibians have only the first 10
pairs. Some of the nerves (e.g., optic nerve) contain only sensory axons, which carry signals to
the brain. Others contain sensory and motor axons, and are termed mixed nerves.
25

Somatic nervous system controls voluntary actions that affect consciousness such as locomotion
and lifting objects. The autonomic nervous system governs involuntary, internal functions of the
body that do not ordinarily affect consciousness, such as movements of the alimentary canal and
heart, contraction of the smooth muscle of blood vessels, urinary bladder, iris of the eye, and
others, plus secretions of various glands. Both types of nerves originate in the brain or spinal
cord. Most organs in the body are innervated by both sympathetic and parasympathetic fibers,
whose actions are antagonistic (if one fiber stimulates an activity, the other inhibits it).

Parasympathetic input stimulates salivary gland secretions and intestinal movements, contracts
pupillary muscles in the eyes, and relaxes sphincter muscles. Sympathetic input controls
antagonistic actions, respectively the inhibition of salivary gland secretions and intestinal
movements, the relaxation of pupillary muscles, and the contraction of sphincters. However,
neither kind of nerve is exclusively excitatory or inhibitory. For example, parasympathetic fibers
inhibit heartbeat but excite peristaltic movements of the intestine; sympathetic fibers increase
heartbeat but inhibit intestinal peristaltic movement.

The sensory branch (afferent) of peripheral nervous system comprises visceral afferent nerves
and sensory afferent nerves. These nerves are responsible for conduction of nervous impulse
that is perceived by sensory organs. Animals require a constant inflow of information from the
environment to regulate their lives. Sense organs are specialized receptors designed for detecting
environmental status and change. An animal’s sense organs are its first level of environmental
perception; they are channels for bringing information to the brain.

3.3. Sensory reception

3.3.1. Sensory response to stimulus

A sense organ transforms energy from a stimulus into nerve impulses, the common language of
the nervous system. A stimulus (pl., stimuli) is any form of energy (electrical, mechanical,
chemical, or radiant) an animal can detect with its receptors. Sense organs are, as a rule, specific
for one kind of stimulus. Eyes respond only to light, ears to sound, pressure receptors to pressure,
and chemoreceptors to chemical molecules.
26

There are various ways animals can obtain information from external factors or stimuli, and five
of them are most common in mammals including humans, and discrete organs are involved in the
process. These are vision, hearing, taste, smell, and touch. The vision involves the eye, hearing is
made possible by the ear, taste is perceived by taste buds on the tongue, smell is felt through
smell nervous sensors in the nasal tract, and touch signals are conveyed by sensory nerve endings
in the skin.

3.3.2. Classification of receptors

Sensory receptors consist of cells that can convert environmental information (stimuli) into nerve
impulses. Receptors are traditionally classified by their location. Those near the external surface,
called exteroceptors, keep the animal informed about the external environment. Internal parts of
the body are provided with interoceptors, which receive stimuli from internal organs.

Sometimes receptors are classified by the form of energy to which the receptors respond, such as
chemical, mechanical, light, or thermal. Baroreceptors sense changes in pressure;
chemoreceptors respond to chemicals; georeceptors respond to the force of gravity;
hygroreceptors detect the water content of air; phonoreceptors respond to sound; photoreceptors
are sensitive to light; proprioceptors, commonly called “stretch receptors”, are internal sense
organs that respond to mechanically induced changes caused by stretching, compression,
bending, or tension; tactile receptors are generally derived from modifications of epithelial cells
associated with sensory neurons; thermoreceptors respond to temperature changes;
mechanoreceptors are excited by mechanical pressures or distortions (e.g. sound, touch, and
muscular contractions).
27

Chapter 4. ENDOCRINE GLANDS AND FUNCTIONS

4.1. Body metabolism regulation

The two basic systems for communication and regulation throughout the body are nervous and
endocrine systems. Chemical signaling by hormones is the function of the endocrine system.
Hormones secreted by endocrine cells regulate reproduction, development, energy metabolism,
growth, and behavior. They also maintain homeostasis and mediate responses to environmental
stimuli; for example, hormones coordinate the body's responses to stress, dehydration, and low
blood glucose. They also control the appearance of characteristics of a juvenile animal from an
adult. The endocrine system communicates by hormones. These are chemicals released into the
blood in small amounts and transported to distant target cells where they initiate physiological
responses.

Many hormones are secreted by endocrine glands, small ductless glands composed of groups of
epithelial cells modified for secretion. Since they have no ducts, their only connection with the
rest of the body is by the bloodstream. Exocrine glands, in contrast, are provided with ducts for
discharging their secretions onto a free surface, e.g. sweat glands and sebaceous glands of skin,
salivary glands, and various enzyme-secreting glands in stomach and intestine. New information
appears however that some neurosecretions may never enter the general circulation and that
some hormones may be secreted by non-endocrine parts (e.g. insulin by nerve cells, cytokines by
immune system). Such hormones may function as neurotransmitters in the brain or as local
tissue factors (parahormones), which stimulate cell growth or some biochemical process.

The endocrine system is slow acting compared with the nervous system because of the time
required for a hormone to reach the appropriate tissue, cross the capillary endothelium, and
diffuse through tissue fluid to cells. Hormonal responses in general are long lasting (minutes to
days; minimum response time is seconds) whereas those under nervous control are short term
(milliseconds to minutes). Despite such differences, the nervous and endocrine systems function
without sharp separation as a single, interdependent system. Endocrine glands often receive
directions from the brain. Conversely, many hormones act on the nervous system and
significantly affect a wide array of animal behaviors.
28

4.2. Mode of chemical regulatory signals

Hormones and other signaling molecules trigger responses by binding to specific receptor
proteins in or on target cells. Only cells that have receptors for a particular secreted molecule are
target cells; other cells are unresponsive to that molecule.

Hormones

Hormones secreted into extracellular fluids by endocrine cells reach target cells via the
bloodstream (or hemolymph). Some endocrine system cells are found in organs that are part of
other organ systems. For example, within the digestive and excretory systems, the stomach and
kidney both contain endocrine cells. Other endocrine cells are grouped in ductless organs called
endocrine glands. Endocrine glands secrete hormones directly into the surrounding fluid.

Local Regulators

Many types of cells produce local regulators, secreted molecules that act over short distances
and reach their target cells solely by diffusion. Examples are cytokines helping communications
between immune cells and growth factors stimulating cell growth and differentiation. Local
regulators play roles in many other processes, including blood pressure regulation, nervous
function, and reproduction. In paracrine signaling, target cells lie near the secreting cell. In
autocrine signaling, the secreted molecules act on the secreting cell itself. Some secreted
molecules have both activities.

Neurotransmitters and neurohormones

Neurons communicate with target cells, such as other neurons and muscle cells, at specialized
junctions, synapses, where molecules called neurotransmitters are secreted and diffuse a very
short distance to bind receptors on the target cells. Neurotransmitters are central to sensation,
memory, cognition, and movement. In neuroendocrine signaling, neurosecretory cells
(specialized neurons typically found in the brain) secrete molecules that diffuse from nerve cell
endings into the bloodstream. These molecules, which travel through the bloodstream to reach
target cells, are a class of hormones called neurohormones. One example is ADH (vasopressin),
a hormone critical to kidney function and water balance.
29

Pheromones

Not all secreted signaling molecules act within the body. Members of the same animal species
sometimes communicate with pheromones, chemicals that are released into the external
environment. Pheromones serve many functions, including marking trails leading to food,
defining territories, warning of predators, and attracting potential mates.

Intercellular communication by secreted molecules

(a) In endocrine signaling, secreted molecules diffuse into the bloodstream and trigger responses in target
cells anywhere in the body

(b) In paracrine signaling, secreted molecules diffuse locally and trigger a response in neighboring cells.

(c) In autocrine signaling, secreted molecules diffuse locally and trigger a response in the cells that
secrete them
30

(d) In synaptic signaling, neurotransmitters diffuse across synapses and trigger responses in cells of target
tissues (neurons, muscles, or glands)

(e) In neuroendocrine signaling, neurohormones diffuse into the bloodstream and trigger responses in
target cells anywhere in the body

Prostaglandins and nitric oxide

A group of local regulators called prostaglandins are modified fatty acids. In semen that reaches
the female reproductive tract, prostaglandins stimulate the smooth muscles of the female's
uterine wall to contract, helping sperm reach an egg. At the onset of childbirth, prostaglandin-
secreting cells of the placenta cause the nearby muscles of the uterus to become more excitable,
helping to induce labor. In the immune system, prostaglandins promote fever and inflammation
and also intensify the sensation of pain. The anti-inflammatory and pain-relieving effects of
aspirin and ibuprofen are due to the inhibition of prostaglandin synthesis by these drugs.
Prostaglandins also help regulate the aggregation of platelets, one step in blood clotting.

The gas nitric oxide (NO), serves in the body as both a neurotransmitter and a local regulator.
When the level of oxygen (O2) in the blood falls, endothelial cells in blood vessel walls
synthesize and release NO. Nitric oxide activates an enzyme that relaxes the neighboring smooth
muscle cells, resulting in vasodilatation, which improves blood flow to tissues. In human males,
31

the ability of NO to promote vasodilatation enables sexual function by increasing blood flow into
the penis, producing an erection. The drug Viagra (sildenafil citrate), a treatment for male
erectile dysfunction, sustains an erection by interfering with this breakdown of NO.

4.3. Hormone chemical forms and solubility

Hormones chemical forms and solubility


Based on their structure and pathway for
synthesis, hormones are often divided into three
groups: polypeptides (proteins and peptides),
amines, and steroids.

The polypeptide hormone insulin is made up of


two polypeptide chains. Like most hormones in
this group, insulin is formed by cleavage of a
longer protein chain. Epinephrine and thyroxine
are amine hormones, which are synthesized from
a single amino acid, either tyrosine or
tryptophan. Steroid hormones (all derived from
the steroid cholesterol), such as cortisol, are
lipids that contain four fused carbon rings.

Hormones vary in their solubility in aqueous and


lipid-rich environments. Polypeptides and many
amine hormones are water-soluble. Being
insoluble in lipids, these hormones cannot pass
through the plasma membranes of cells. In
contrast, steroid hormones, as well as other
largely nonpolar hormones, such as thyroxine,
are lipid-soluble and can pass through cell
membranes readily.
32

4.4. Major endocrine glands and hormones

Major human endocrine glands


33

Major endocrine glands, hormones, and representative action

Gland Hormone Target cells and principal actions


Thyroid-stimulating
hormone (TSH) Stimulates release of TSH by anterior pituitary
Adrenocorticotropin-
releasing hormone (CRH) Stimulates release of ACTH by anterior pituitary
Hypothalamus Gonadotropin-releasing
hormone (GnRH) Stimulates gonadotropin release by anterior pituitary
Prolactin-inhibiting factor
(PIF) Inhibits prolactin release by anterior pituitary
Somatostatin Inhibits prolactin release by anterior pituitary
Stimulates growth of bone and muscle; promotes protein
Growth hormone (GH) or synthesis; affects lipid and carbohydrate metabolism;
Somatotropin (STH) increases cell division
Initiates milk production and secretion by mammary
glands; acts on crop sacs of some birds; stimulates
Prolactin (PRL) maternal behavior in birds
In ovary: Stimulates growth of follicles; functions with LH
Follicle-stimulating to cause estrogen; In testis: Acts on seminiferous tubules to
Anterior hormone (FSH) promote spermatogenesis
pituitary Luteinizing hormone (LH) In ovary: Forms corpora lutea; secretes progesterone;
gland or Interstitial cell probably acts in conjunction with FSH; In testis:
stimulating hormone Stimulates the interstitial cells, thus promoting the
(ICSH) secretion of testosterone
Stimulates thyroid gland to synthesize and release thyroid
Thyroid-stimulating hormones concerned with growth, development, metabolic
hormone (TSH) rate
Adrenocorticotropic Stimulates secretion of adrenocortical steroids; is involved
hormone (ACTH) in stress response
Endorphins Decrease pain
Intermediate Promotes melanin synthesis; darkens the skin; responds to
lobe of Melanocyte-stimulating external stimuli; expands amphibian melanophores
pituitary hormone (MSH) (pigment-containing cells)
Affects postpartum mammary gland, causing ejection of
milk; promotes contraction of uterus; has possible action in
Posterior
parturition and in sperm transport in female reproductive
pituitary
Oxytocin tract
gland
Antidiuretic hormone Elevates blood pressure by acting on arterioles; promotes
(ADH) or Vasopressin reabsorption of water by kidney tubules
34

Triiodothyronine (T3) and Affect growth, amphibian metamorphosis, molting,


thyroxine (T4) metabolic rate in birds and mammals, development
Thyroid gland
Lowers blood calcium level by inhibiting calcium
Calcitonin reabsorption from bone

Parathyroid Parathyroid hormone


glands (PTH) or Parathormone Regulates phosphate and calcium concentration in blood
Promotes glycogen synthesis and glucose utilization and
Insulin (beta cells) uptake from blood
Pancreas
Raises blood glucose concentration by stimulating the
Giucagon (alpha cells) breakdown of glycogen into glucose
Promote synthesis of carbohydrates and breakdown of
Glucocorticoids (e.g. proteins; initiate antiinflammatory and antiallergic actions;
Adrenal
cortisol) mediate response to stress
cortex
Mineralocorticoids (e.g. Regulate sodium retention and potassium loss through
aldosterone) kidneys, and water balance

Mobilizes glucose; increases blood flow through skeletal


Adrenal
Epinephrine (Adrenaline) muscle; increases oxygen consumption; increases heart rate
medulla
Norepinephrine Elevates blood pressure; raises the level of blood glucose;
(Noradrenaline) constricts arterioles and venules
Support sperm formation (spermatogenesis); promote
Androgens (e.g. development and maintenance of male secondary sex
Testes testosterone) characteristics
Stimulate uterine lining growth; promote oogenesis;
regulate the menstrual and estrus cycles; promote
Ovaries
development and maintenance of female secondary sex
Estrogens (e.g. estradiol) characteristics
Ovaries &
testes Inhibin Inhibits the secretion of FSH
Progestins (e.g. Promote uterine lining growth (for fertilized egg);
progesterone) stimulates development of mammary glands
Corpus
Relaxes the pelvic ligaments of female mammals during
luteum
pregnancy; softens the opening of the uterus (cervix) at the
Relaxin time of delivery
Involved in biological rhythms (physiological cycles);
Pineal gland Melatonin control variations in skin color
Thymopoietin (TP), and Development and differentiation of T-cells of the immune
Thymus Thymosin system
35

In addition to the major endocrine glands, other glands and organs carry on hormonal activity.

Gland/ Organ
Hormone Function Target area
Estrogens, progesterone, human
Ovaries,
chorionic gonadotropin (hCG),
Placenta Maintain pregnancy mammary
human chorionic
glands, uterus
somatomammotropin (hCS)
Digestive Stimulates release of pancreatic juice to Cells of
tract Secretin neutralize stomach acid pancreas
Stimulates digestive enzymes and HCl Stomach
Gastrin in stomach mucosa
Stimulates release of pancreatic Pancreas,
Cholecystokinin (CCK) enzymes and bile from gallbladder gallbladder
Lowers blood pressure, maintains fluid Blood vessels,
Heart Atriopeptin balance kidneys
Kidneys Erythropoietin Stimulates red blood cell production Bone marrow
Urotensin Stimulates constriction of arteries Major arteries
Aids in the absorption of dietary
Calcitrol calcium and phosphorus Small intestine
Adipose
tissue Leptin Suppresses appetite Brain

4.5. Endocrine and nervous coordination by hypothalamus

The hypothalamus plays a central role in integrating the endocrine and nervous systems. The
hypothalamus receives information from nerves throughout the body and from other parts of the
brain. In response, it initiates endocrine signaling appropriate to environmental conditions.
Signals from the hypothalamus travel to the pituitary gland, a gland located at its base. Modified
hypothalamic nerve cells (neurosecretary cells) project their axons down a stalk of nerve cells
and blood vessels, called the infundibulum, into the pituitary gland, directly linking the nervous
and endocrine systems. The hypothalamic releasing and inhibiting hormones are secreted near
capillaries at the base of the hypothalamus.

The pituitary has discrete posterior and anterior parts (lobes), which are actually two glands, the
posterior pituitary and the anterior pituitary. The pituitary of many vertebrates (but not in
humans, birds, and cetaceans) also has a functional intermediate lobe (pars intermedia) of mostly
glandular tissue. Its secretions (e.g. melanophore-stimulating hormone) in response to external
stimuli induce changes in the coloration of the body surface of many animals.
36

The posterior pituitary, or neurohypophysis, an extension of the hypothalamus, stores and


secretes two hormones made by the hypothalamus. Hormones released by the hypothalamus
regulate secretion of hormones by the anterior pituitary, or adenohypophysis. Under the control
of the hypothalamus, the anterior pituitary and posterior pituitary produce a set of hormones
central to endocrine signaling throughout the body.

Posterior pituitary hormones

The posterior pituitary releases two neurohormones, oxytocin and antidiuretic hormone (ADH).
Synthesized in the hypothalamus, these hormones travel along the long axons of neurosecretory
cells to the posterior pituitary. There they are stored and released when needed.

Hypothalamus and anterior pituitary

The anterior pituitary synthesizes and secretes many different hormones and is itself regulated by
hormones secreted by the hypothalamus. Each hypothalamic hormone is either a releasing
hormone or an inhibiting hormone, reflecting its role in promoting or inhibiting release of one or
more specific hormones by the anterior pituitary. Thyrotropin-releasing hormone (TRH), for
example, is a product of the hypothalamus that stimulates the anterior pituitary to secrete
thyrotropin, also known as thyroid-stimulating hormone (TSH). Every anterior pituitary hormone
is controlled by at least one releasing hormone. Some have both a releasing hormone and an
inhibiting hormone.

Sets of hormones from the hypothalamus, the anterior pituitary, and a target endocrine gland are
often organized into a hormone cascade pathway. Signals to the brain stimulate the
hypothalamus to secrete a hormone that in turn either stimulates or inhibits release of a particular
anterior pituitary hormone. The anterior pituitary hormone acts on a target endocrine tissue,
stimulating secretion of yet another hormone that exerts systemic metabolic or developmental
effects. When a young child's body temperature drops, the hypothalamus secretes TRH. TRH
targets the anterior pituitary, which responds by secreting TSH. TSH acts on the thyroid gland to
stimulate release of thyroid hormone. As it accumulates, thyroid hormone increases metabolic
rate, releasing thermal energy that raises body temperature.
37

Tropic hormones

TSH is an example of a tropic hormone, a hormone that regulates the function of endocrine cells
or glands. Three other anterior pituitary hormones act primarily or exclusively as tropic
hormones: follicle-stimulating hormone (FSH), luteinizing hormone (LH) and adrenocortico-
tropic hormone (ACTH). FSH and LH stimulate the activities of the male and female gonads,
the testes and ovaries, respectively. For this reason, FSH and LH are also known as
gonadotropins. Their role will be addressed again in the reproductive function.

Nontropic hormones

Two major hormones of the anterior pituitary target nonendocrine tissues and are thus nontropic.
They are prolactin (PRL) and melanocyte-stimulating hormone (MSH). PRL stimulates
mammary gland growth and milk synthesis in mammals. PRL is remarkable for the diversity of
other effects among vertebrate species: it regulates fat metabolism and reproduction in birds,
delays metamorphosis in amphibians, and regulates salt and water balance in freshwater fishes.

Growth Hormone

Growth hormone (GH), which is secreted by the anterior pituitary, stimulates growth through
tropic and nontropic effects. A major target, the liver; responds to GH by releasing insulin-like
growth factors (IGFs), which circulate in the blood and directly stimulate bone and cartilage
growth. In the absence of GH, the skeleton of an immature animal stops growing. GH also exerts
diverse metabolic effects that tend to raise blood glucose levels, thus opposing the effects of
insulin.

Abnormal production of GH in humans can result in several disorders, depending on when the
problem occurs and whether it involves hypersecretion (too much) or hyposecretion (too little).
Hypersecretion of GH during childhood can lead to gigantism, in which the person grows
unusually tall. In adulthood GH hypersecretion stimulates bony growth in the few tissues that are
still responsive to the hormone, resulting is an overgrowth of the extremities, an abnormality
called acromegaly. Hyposecretion of GH in childhood retards long-bone growth and can lead to
pituitary dwarfism. If diagnosed before puberty, pituitary dwarfism can be treated successfully
with human GH.
38

4.6. Hormone function mechanisms

4.6.1. Simple hormone pathways

In response to an internal or environmental stimulus, a specific hormone travels in the


bloodstream to target cells, where it interacts with its specific receptors. Signal transduction
within target cells brings about a physiological response.

Let us examine the function of epinephrine or adrenaline. When you find yourself in a stressful
situation, perhaps running to catch a bus, your adrenal glands secrete epinephrine. When
epinephrine reaches liver cells, it binds to a G protein-coupled receptor in the plasma membrane.
The binding of hormone to receptor triggers a cascade of events involving synthesis of cyclic
AMP (cAMP, cyclic adenosine monophosphate that activates enzymes in many hormone-
induced biochemical reactions) as a short-lived second messenger. Activation of protein-kinase
A by cAMP leads to activation of an enzyme required for glycogen breakdown and inactivation
of an enzyme necessary for glycogen synthesis. The net result is that the liver releases glucose
into the bloodstream, providing the fuel you need to chase the departing bus.

Physiological responses to hormone activity alter rates of many different biochemical processes
such as enzymatic activity, membrane permeability, synthesis of cellular proteins, or release of
hormones. Therefore they must be regulated by the appropriate hormones, not merely activated.
However, the concentration of a hormone in the plasma depends on its rate of secretion and that
of inactivation and removal from the circulation. If secretion is to be correctly controlled, an
endocrine gland requires information about the level of its own hormone(s) in the plasma.

Negative hormone feedback mechanism

Many hormones are controlled by negative feedback systems that operate between glands
secreting the hormones and target cells. A feedback pattern is one in which output is constantly
compared with a set point. For example, CRH (corticotropin-releasing hormone), secreted by the
hypothalamus, stimulates the pituitary (the target cells) to release ACTH. ACTH stimulates the
adrenal gland (the target cells) to secrete cortisol.
39

As the level of ACTH rises in the plasma, it acts on, or “feeds back” on, the hypothalamus to
inhibit release of CRH. Similarly, as cortisol level rises in the plasma, it “feeds back” on the
hypothalamus and pituitary to inhibit release of both CRH and ACTH respectively. Thus any
deviation from the set point (a specific plasma level of each hormone) leads to corrective action
in the opposite direction. Such a negative feedback system is highly effective in preventing
extreme oscillations in hormonal output.

Positive feedback mechanism

One function of oxytocin in mammals is to regulate milk release during nursing. A stimulus
received by a sensory neuron stimulates a neurosecretory cell to secrete a neurohormone, which
diffuses into the bloodstream and travels to target cells. The initial stimulus is the infant is
suckling. Stimulation of sensory nerve cells in the nipples generates signals in the nervous
system that reach the hypothalamus. A nerve impulse from the hypothalamus then triggers the
release of oxytocin from the posterior pituitary gland. In response to circulating oxytocin, the
mammary glands secrete milk.

The oxytocin pathway regulating the mammary gland provides an example of a positive-
feedback mechanism. Unlike negative feedback, which dampens a stimulus, positive feedback
reinforces a stimulus, leading to an even greater response. Thus, oxytocin stimulates milk
release, which leads to more suckling and therefore more stimulation. Activation of the pathway
is sustained until the baby stops suckling.
40

4.6.2. Insulin and glucagon

Two antagonistic hormones insulin and glucagon regulate the concentration of glucose in the
blood. Each hormone operates in a simple endocrine pathway regulated by negative feedback.
When blood glucose rises above the set point, release of insulin triggers uptake of glucose from
the blood, decreasing the blood glucose concentration. When blood glucose drops below the set
point, the release of glucagon promotes the release of glucose into the blood, increasing the
blood glucose concentration. Because insulin and glucagon have opposing effects, the combined
activity of these two hormones tightly controls the concentration of glucose in the blood.

Glucagon and insulin are produced in the pancreas. Scattered throughout the pancreas are dusters
of endocrine cells known as the islets of Langerhans. Each islet has alpha cells, which make
glucagon, and beta cells, which make insulin. Like all hormones, insulin and glucagon are
secreted into the interstitial fluid and enter the circulatory system. Overall, hormone-secreting
cells make up only 1-2% of the mass of the pancreas. Other cells in the pancreas produce and
secrete bicarbonate ions and digestive enzymes. These secretions are released into small ducts
that empty into the pancreatic duct, which leads to the small intestine. Thus, the pancreas is both
an endocrine gland and an exocrine gland with functions in the endocrine and digestive systems.

Insulin lowers blood glucose levels by stimulating nearly all body cells outside the brain to take
up glucose from the blood. Brain cells can take up glucose without insulin, so the brain almost
always has access to circulating fuel. Insulin also decreases blood glucose by slowing glycogen
breakdown in the liver and inhibiting the conversion of amino acids and glycerol (from fats) to
glucose.

Glucagon influences blood glucose levels through its effects on target cells in the liver. The liver,
skeletal muscles, and adipose tissues store large amounts of fuel. The liver and muscles store
sugar as glycogen, whereas cells in adipose tissue convert sugars to fats. Of these tissues, only
those in the liver are sensitive to glucagon. When the blood glucose level decreases to or below
the set point (approximately 90 mg/100 mL), glucagon signals the liver cells to increase
glycogen hydrolysis, convert amino acids and glycerol to glucose, and release glucose into the
bloodstream.
41

Diabetes mellitus is a physiological disorder caused by a deficiency of insulin or a decreased


response to insulin in target tissues. Blood glucose levels rise, but cells are unable to take up
enough glucose to meet metabolic needs. Instead, fat becomes the main substrate for cellular
respiration. In severe cases, acidic metabolites formed during fat breakdown accumulate in the
blood, threatening life by lowering blood pH and depleting sodium and potassium ions from the
body. In people with diabetes mellitus, the high level of glucose in blood exceeds the capacity of
the kidneys to reabsorb this nutrient. Glucose that remains in the filtrate is excreted. For this
reason, the presence of sugar in urine is one test for this disorder.

There are two main types of diabetes mellitus. Each is marked by high blood glucose, but with
very different causes. Type 1 diabetes, or insulin-dependent diabetes, is an autoimmune disorder
in which the immune system destroys the beta cells of the pancreas. Type 1 diabetes, which
usually appears during childhood, destroys the person's ability to produce insulin. Type 2
diabetes or non-insulin-dependent diabetes, is characterized by a failure of target cells to respond
normally to insulin. Insulin is produced, but target cells fail to take up glucose from the blood,
and blood glucose levels remain elevated.
42

Chapter 5. REPRODUCTIVE PROCESS AND REGULATION

5.1. Sexual reproduction

Sexual reproduction is the production of individuals with gametes (eggs and sperm). The
distinction between male and female is fundamentally based on the size and mobility of the
gametes they produce. The ovum (egg) is produced by the female. Ova are large (because of
stored yolk to sustain early development), non-motile, and produced in relatively small numbers.
The spermatozoon (sperm) is produced by the male. Sperm are small, motile, and produced in
enormous numbers. Each contains a highly condensed genetic material designed for the single
purpose of reaching and fertilizing an egg.

Meiosis is a distinctive type of gamete-producing nuclear division. Meiosis differs from ordinary
cell division (mitosis) in being a double division. The chromosomes split once, but the cell
divides twice, producing four cells, each with half the original number of chromosomes (the
haploid number). Meiosis is followed by fertilization in which two haploid gametes are
combined to restore the normal (diploid) chromosomal number of the species. The new cell
(zygote), which now begins to divide by mitosis, has equal numbers of chromosomes from each
parent and accordingly is different from each. It is a unique individual bearing a recombination
of parental characteristics (genetic recombination).

The male-female distinction is more clearly evident in most animals. Organs that produce germ
cells are called gonads. The gonad that produces sperm is a testis and that which forms eggs is
an ovary. Gonads represent the primary sex organs, the only sex organs found in certain groups
of animals. Most metazoa, however, have various accessory sex organs (such as penis, vagina,
uterine tubes, and uterus) that transfer and receive germ cells.

5.2. The origin and maturation of germ cells

The vertebrate body is composed of non-reproductive somatic cells, which are differentiated for
specialized functions and germ cells, which form the gametes (eggs and sperm). Germ cells
provide continuity of life between generations and ensure the species’ survival. Somatic cells are
43

equally necessary for the support, protection, and nourishment of the germ cells during their
development (gametogenesis).

The series of transformations that results in the formation of mature gametes is called
gametogenesis. Although the same essential processes are involved in the maturation of both
sperm and eggs, there are some important differences. Gametogenesis in testes is called
spermatogenesis, and in ovaries, oogenesis.

5.2.1. Spermatogenesis

The walls of the seminiferous tubules contain differentiating germ cells arranged in a stratified
Human
layer five to eight cells deep. Germ cells develop in close contact with large sustentacular
sperm
(Sertoli) cells, which extend from the periphery of the seminiferous tubules to the lumen and
provide nourishment during germ cell development and differentiation.

The outermost layers contain


spermatogonia, diploid cells
that have increased in number
by ordinary mitosis. Each
spermatogonium increases in
size and becomes a primary spermatocyte. Each primary spermatocyte then undergoes the first
44

meiotic division, to become two secondary spermatocytes. Each secondary spermatocyte enters
the second meiotic division without the intervention of a resting period. In the two steps of
meiosis each primary spermatocyte gives rise to four spermatids, each containing the haploid
number (23 in humans) of chromosomes.

Without further divisions the spermatids are transformed into mature spermatozoa or (sperm).
Modifications include great reduction of cytoplasm, condensation of the nucleus into a head,
formation of a middle piece containing mitochondria, and a whiplike, flagellar tail for
locomotion. The head consists of a nucleus containing the chromosomes for heredity and an
acrosome, a distinctive feature of nearly all the metazoa. In many animal species, the acrosome
contains lysins that serve to clear an entrance through the layers that surround the egg. The
fusion of the egg and sperm plasma membranes is the initial event of fertilization.

5.2.2. Oogenesis

Early germ cells in the ovary, called oogonia, increase in number by ordinary mitosis. Each
oogonium contains the diploid number of chromosomes. After the oogonia cease to increase in
number, they grow in size and become primary oocytes. Before the first meiotic division, the
chromosomes in each primary oocyte meet in pairs, paternal and maternal homologues, just as in
spermatogenesis.

When the first maturation (reduction) division occurs, the cytoplasm is divided unequally. One
of the two daughter cells, the secondary oocyte, is large and receives most of the cytoplasm; the
other is very small and is called the first polar body. Each of these daughter cells, however, has
received half of the chromosomes.

In the second meiotic division, the secondary oocyte divides into a large ootid and a small polar
body. If the first polar body also divides in this division, which sometimes happens, there are
three polar bodies and one ootid. The ootid develops into a functional ovum. The polar bodies
are nonfunctional, and they disintegrate.

Formation of nonfunctional polar bodies is necessary to enable the egg to dispose of excess
chromosomes, and the unequal cytoplasmic division makes possible a large cell with the
cytoplasm containing sufficient yolk for the development of the young. Thus a mature ovum has
45

the N (haploid) number of chromosomes, the same as a sperm. However, each primary oocyte
gives rise to only one functional gamete instead of four as in spermatogenesis.

Oogenesis

In most vertebrates and many invertebrates the egg does not actually complete all the meiotic
divisions before fertilization occurs. The general rule is that development is arrested during
prophase I of the first meiotic division. Meiosis resumes and is completed either at the time of
ovulation (birds and most mammals) or shortly after fertilization (many invertebrates, teleost
fishes, amphibians, and reptiles). In humans, the ova begin the first meiotic division at about the
thirteenth week of fetal development, then their development arrests in prophase I as the primary
oocyte until puberty, at which time one of these primary oocytes typically develops each
menstrual month into a functional egg. Meiosis II is completed only when the ovum is penetrated
by a spermatozoon.

The most obvious feature of egg maturation is the deposition of yolk. Yolk, usually stored as
granules or more organized platelets, is not a definite chemical substance but may be lipid or
protein or both. In insects and vertebrates, all having more or less yolky eggs, yolk may be
synthesized within an egg from raw materials supplied by surrounding follicle cells, or
46

preformed lipid or protein yolk may be transferred by pinocytosis from follicle cells to the
oocyte.

The result of the enormous accumulation of yolk granules and other nutrients (glycogen and lipid
droplets) is that an egg grows well beyond the normal limits that force ordinary body (somatic)
cells to divide. A young frog oocyte 50 μm in diameter, for example, grows to 1500 μm in
diameter when mature after 3 years of growth in the ovary, and its volume has increased by a
factor of 27,000. Bird eggs attain even greater absolute size; a hen egg will increase 200 times in
volume in only the last 6 to 14 days of rapid growth preceding ovulation.

5.3. Male reproductive system

The male reproductive system of vertebrates such as that of human males includes testes, vasa
efferentia, vas deferens, accessory glands, and (in some birds and reptiles, and all mammals) a
penis. The paired testes are the sites of sperm production. Each testis is composed of numerous
seminiferous tubules, in which the sperm develop. The sperm are surrounded by sustentacular
cells (or Sertoli cells), which nourish the developing sperm. Between the tubules are interstitial
cells, which produce the male sex hormone (testosterone).

Human male reproductive system in sagittal view


47

The sperm travel from the seminiferous tubules to the vasa efferentia, small tubes passing to a
coiled epididymis (one for each testis), where final sperm maturation takes place, and then to a
vas deferens, the ejaculatory duct. In mammals the vas deferens joins the urethra, a duct that
serves to carry both sperm and urinary products through the penis, or external intromittent organ.
The term semen is used to mean the thick white fluid containing sperm that a male ejaculates, but
sometimes sperm is used for semen to mean the ejaculated fluid.

In most mammals three sets of accessory glands open into the reproductive channels: a pair of
seminal vesicles, a single prostate gland, and the pair of bulbourethral (Cowper’s) glands. Fluid
secreted by these glands furnishes food to the sperm, lubricates the passageways for sperm, and
counteracts the acidity of the urine so that the sperm are not harmed.

5.4. Female reproductive system

The ovaries of female vertebrates produce both ova and female sex hormones (estrogens and
progesterone). In most vertebrates, mature ova from each ovary enter the funnel-like opening of
a uterine tube or oviduct, which typically has a fringed margin that envelops the ovary at the
48

time of ovulation. In cartilaginous fishes, reptiles, and birds that produce a large shelled egg,
special regions have developed for production of albumin and shell. In amniotes (reptiles, birds,
and mammals) the terminal portion of the uterine tube is expanded into a muscular uterus in
which shelled eggs are held before laying or in which embryos complete their development. In
placental mammals, the walls of the uterus establish a close vascular association with the
embryonic membranes through a placenta.

The paired ovaries of the human female, slightly smaller than the male testes, contain many
thousands of oocytes. Each oocyte develops within a follicle that enlarges and finally ruptures to
release a secondary oocyte. During a woman’s fertile years, except following fertilization,
approximately 13 oocytes mature each year, and usually the ovaries alternate in releasing
oocytes. Because a woman is fertile for only about 30 years, of the approximately 400,000
primary oocytes in her ovaries at birth, only 300 to 400 have a chance to reach maturity; the
others degenerate and are absorbed.

Human female reproductive system in sagittal section


49

The uterine tubes, or oviducts, are lined with cilia for propelling the egg in its course. The two
ducts open into the upper corners of the uterus, or womb, which is specialized for housing the
embryo during the 9 months of its intrauterine existence. It is provided with thick muscular
walls, many blood vessels, and a specialized lining: the endometrium. The vagina is a muscular
tube adapted for receiving the male’s penis and for serving as birth canal during expulsion of a
fetus from the uterus. Where vagina and uterus meet, the uterus projects down into the vagina to
form a cervix.

The external genitalia of human females, or vulva, include folds of skin, the labia majora and
labia minora, and a small erectile organ, the clitoris (the female homolog of the glans penis of
males). The opening into the vagina is often reduced in size in the virgin state by a membrane,
the hymen, although in sexually active females, this membrane may be much reduced in extent.
50

5.5. Endocrine regulation in reproduction

Reproduction in vertebrates is usually a seasonal or cyclic activity. Timing is crucial, because the
young should appear when food is available and other environmental conditions are optimal for
survival. Neurosecretory centers of the brain (hypothalamus) regulate the release of anterior
pituitary gland hormones. This delicately balanced hormonal system controls development of the
gonads, accessory sex structures, and secondary characteristics, as well as timing of
reproduction.

The cyclic reproductive patterns of mammals are of two types: estrous cycle, characteristic of
most mammals, and menstrual cycle, characteristic only of the anthropoid primates (monkeys,
apes, and humans). In estrous cycles, females are receptive to males only during brief periods of
estrus, or “heat,” whereas in the menstrual cycle receptivity may occur throughout the cycle. A
menstrual cycle, but not an estrous cycle, ends with collapse and discharge of the inner portion of
the endometrium (uterine lining). In an estrous animal, each cycle ends with the uterine lining
simply reverting to its original state, without the discharge characteristic of the menstrual cycle.

5.5.1. Gonadal steroids and their control

The ovaries produce two kinds of steroid sex hormones (growth-promoting hormones, GPH) in
females: estrogens and progesterone. There are 3 kinds of estrogens: estradiol, estrone and
estriol, of which estradiol is secreted in the highest amounts during reproductive cycles.
Estrogens are responsible for development of female accessory sex structures (oviducts, uterus,
and vagina) and for stimulating female reproductive activity.

Secondary sex characters, those characteristics that are not primarily involved in formation and
delivery of ova (or sperm in the male), but that are essential for behavioral and functional
success of reproduction, are also controlled or maintained by estrogens. These include
characteristics such as distinctive skin or feather coloration, bone development, body size and, in
mammals, initial development of the mammary glands.

In mammals, both estrogen and progesterone are responsible for preparing the uterus to receive a
developing embryo. These hormones are controlled by pituitary gonadotropins: FSH and LH.
The two gonadotropins are in turn governed by GnRH produced by neurosecretory cells in the
51

hypothalamus. Through this control system, environmental factors such as light, nutrition, and
stress may influence reproductive cycles.

The male sex steroid, testosterone, is manufactured by the interstitial cells of the testes.
Testosterone, and its metabolite, dihydrotestosterone (DHT), are necessary for the growth and
development of the male accessory sex structures (penis, sperm ducts, and glands), development
of secondary male sex characters (such as bone and muscle growth, male plumage or pelage
coloration, antlers in deer, and, in humans, voice quality), and male sexual behavior.
Development of the testes and secretion of testosterone is controlled by FSH and LH, the same
pituitary hormones that regulate the female reproductive cycle, and ultimately by GnRH from the
hypothalamus.

Testosterone and DHT makes feedback to the hypothalamus and anterior pituitary to keep the
secretion of GnRH, FSH, and LH in check. The testes also secrete a second hormone, the peptide
inhibin, which is secreted by the sustentacular cells (or Sertoli cells). This hormone is an
additional regulator of the secretion of FSH from the anterior pituitary in a negative feedback
manner.

5.5.2. The menstrual cycle

The human menstrual cycle consists of two distinct phases within the ovary: follicular phase and
luteal phase, and three distinct phases within the uterus: menstrual phase, proliferative phase and
secretory phase. Menstruation (the “period”) signals the menstrual phase, when part of the
lining of the uterus (endometrium) degenerates and sloughs off, producing the menstrual
discharge. Meanwhile, the follicular phase within the ovary is occurring, and by day 3 of the
cycle blood levels of FSH and LH begin to rise slowly, prompting some of the ovarian follicles
to begin growing and to secrete estrogen.

As estrogen levels in the blood increase, the uterine endometrium heals and begins to thicken,
and uterine glands within the endometrium enlarge (proliferative phase). By day 10 most of the
ovarian follicles that began to develop at day 3 now degenerate (become atretic), leaving only
one (sometimes two or three) to continue ripening until it appears like a blister on the surface of
the ovary. This is a mature follicle or graafian follicle. During the latter part of the follicular
52

phase, the graafian follicle secretes more estrogen, and also inhibin. Inhibin acts as a negative
feedback regulator of FSH (as in males), and as the levels of inhibin rise, the levels of FSH fall.

Human ovarian and menstrual cycles


53

At day 13 or 14 in the cycle, the now high levels of estrogen from the graafian follicle stimulate
a surge of GnRH from the hypothalamus, which induces a surge of LH (and to a lesser extent,
FSH) from the anterior pituitary. The LH surge causes the largest follicle to rupture (ovulation),
releasing the oocyte from the ovary. Now follows a critical period, for unless a mature oocyte is
fertilized within a few hours, it will die. During the ovarian luteal phase, a corpus luteum forms
from the remains of the ruptured follicle that released the oocyte at ovulation.

The corpus luteum, responding to continued stimulation of LH, becomes a transitory endocrine
gland that secretes progesterone (and estrogen in primates). Progesterone, as its name implies,
stimulates the uterus to undergo final maturational changes that prepare it for gestation
(secretory phase). The uterus is now fully ready to house and nourish an embryo. If fertilization
has not occurred, the corpus luteum degenerates, and its hormones are no longer secreted. Since
the uterine lining (endometrium) depends on progesterone and estrogen for its maintenance, their
declining levels cause the uterine lining to deteriorate, leading to menstrual discharge of the next
cycle.

GnRH from the hypothalamus, and LH and FSH from the anterior pituitary, are controlled by
negative feedback of ovarian steroids (and inhibin). This negative feedback occurs throughout
the menstrual cycle, except for a few days before ovulation. As mentioned above, ovulation is
due to the high levels of estrogen causing a surge of GnRH, LH (and FSH). Such positive
feedback mechanisms are rare in the body, since they move events away from stable set points.
This event is terminated by ovulation when estrogen levels fall as an oocyte is released from the
follicle.

The following figure illustrates the human menstrual cycle, showing changes in blood hormone
levels and uterine endometrium during the 28-day cycle. FSH promotes maturation of ovarian
egg follicles, which secrete estrogen. Estrogen prepares the uterine endometrium and causes a
surge in LH, which in turn stimulates the corpus luteum to secrete progesterone and estrogen.
Progesterone and estrogen production will persist only if the egg is fertilized; without pregnancy
progesterone and estrogen levels decline and menstruation follows.
54

5.5.3. Regulation of human pregnancy and birth

If fertilization occurs, it normally does so in the first third of the uterine tube (ampulla), and the
zygote travels from here to the uterus, dividing by mitosis to form a blastocyst by the time it
reaches the uterus. The developing blastocyst will contact the uterine surface after about 6 days
and bury itself in the endometrium. This process is called implantation.

Growth of the embryo continues, producing a spherically shaped trophoblast. This embryonic
stage contains three distinct tissue layers, the amnion, chorion, and embryo proper, the inner cell
mass. The chorion becomes the source of human chorionic gonadotropin (hCG), which appears
in the bloodstream soon after implantation. hCG stimulates the corpus luteum to synthesize and
release both estrogen and progesterone.

The point of attachment between trophoblast and uterus becomes the placenta. Besides serving
as a medium for the transfer of materials between maternal and fetal bloodstreams, the placenta
also serves as an endocrine gland. The placenta continues to secrete hCG and also produces
estrogen (mainly estriol) and progesterone. After about the third month of pregnancy, the corpus
luteum degenerates, but by then the placenta itself is the main source of both progesterone and
estrogen.

Preparation of the mammary glands for secretion of milk requires two additional hormones,
prolactin (PRL) and human placental lactogen (hPL) (or human chorionic somatomammotropin).
PRL is produced by the anterior pituitary, but in nonpregnant women its secretion is inhibited.
During pregnancy, elevated levels of progesterone and estrogen depress the inhibitory signal, and
PRL begins to appear in the blood.

PRL, in combination with hPL, prepare the mammary glands for secretion. hPL, together with
maternal growth hormone, also stimulates an increase in available nutrients in the mother, so that
more are provided to the developing embryo. Later the placenta begins to synthesize a peptide
hormone called relaxin; this hormone allows some expansion of the pelvis by increasing the
flexibility of the pubic symphysis, and also dilates the cervix in preparation for delivery.
55

Birth, or parturition, begins with a series of strong, rhythmic contractions of the uterine
musculature, called labor. The exact signal that triggers birth is not fully understood in humans,
but several important factors have been identified in other mammals. Just before birth, secretion
of estrogen, which stimulates uterine contractions, rises sharply, while the level of progesterone,
which inhibits uterine contractions, declines. This removes the “progesterone block” that keeps
the uterus quiescent throughout pregnancy.

Prostaglandins, a large group of hormones (long-chain fatty acid derivatives), also increase at
this time, making the uterus more “irritable”. Finally, the stretching of the uterus sets in motion
neural reflexes that stimulate secretion of oxytocin from the posterior pituitary. Oxytocin also
stimulates uterine smooth muscle, leading to stronger and more frequent labor contractions.

After birth, secretion of milk is triggered when the infant sucks on its mother’s nipple. This leads
to a reflex release of oxytocin from the pituitary; when oxytocin reaches the mammary glands it
causes contraction of smooth muscles lining ducts and sinuses of the mammary glands and
ejection of milk. Suckling also stimulates release of prolactin from the anterior pituitary gland,
which stimulates continued production of milk by the mammary glands.
56

Chapter 6. SKELETON, MUSCLE, AND MOTION

6.1. Animal skeletons and their roles

6.1.1. Exoskeleton and endoskeleton

There are two principal types of rigid skeletons in animals: exoskeleton, typical of mollusks,
arthropods and many other invertebrates and endoskeleton, characteristic of echinoderms and
vertebrates. Unlike an endoskeleton, which grows with the animal, an exoskeleton is often a
limiting coat of armor that must be periodically molted to make way for an enlarged
replacement. The vertebrate endoskeleton is formed inside the body and is composed of bone
and cartilage, which are forms of dense connective tissue.

The skeleton is defined as the rigid framework of interconnected bones and cartilage that protects
and supports the internal organs and provides attachment for muscles in humans and other
vertebrate animals. Skeletons are supportive systems that provide rigidity to the body, surfaces
for muscle attachment, and protection for vulnerable body organs. All ribs protect the heart and
lungs. Skull bones protect the fragile nervous system, especially the brain. The articulation
between the thoracic vertebrae provides the flexibility. Bone not only supports and protects but is
also the major body reservoir for calcium and phosphorus. In amniote vertebrates red blood cells
and certain white blood cells are formed in the bone marrow.

Bone is a specialized component of the human skeleton that provides a point of attachment for
muscles and transmits the force of muscular contraction from one part of the body to another
during movement. In addition, bones of the skeleton support the internal organs of many
animals, store reserve calcium and phosphate, and manufacture red blood cells and some white
blood cells. Bone tissue contains inorganic salts mainly calcium phosphate and calcium
carbonate. When an animal needs the calcium or phosphate stored within bones, metabolic
reactions (under endocrine control) release the required amounts.

This chapter will emphasize the role and function of the skeleton rather than identification and
description of the skeleton structures. The following figure shows the general view of the human
skeleton.
57

Human skeleton (ventral and dorsal view)


58

6.1.2. Bone structure and cells

The typical cells (osteocytes) of a compact bone are in minute chambers called lacunae, which
are arranged in concentric rings around osteonic canals or osteons (formerly called Haversian
systems). These cells communicate with nearby cells by means of cellular processes passing
through small channels called canaliculi. Bone consists of bundles of osteons cemented together
and interconnected with blood vessels and nerves. This tiny network of channels also helps
osteocytes communicate with blood vessels that penetrate into bone. The matrix around them is
heavily impregnated with calcium phosphate, making this kind of tissue hard and ideally suited
for its functions of support and protection.

Structure of a bone
59

6.1.3. Bone formation and growth

Most bone develops from cartilage and is called endochondral (within cartilage) or replacement
bone. Intramembranous bone develops directly from sheets of embryonic cells. Whatever the
embryonic origin, once fully formed, endochondral and intramembranous bone look the same.
Fully formed bone, however, may vary in density. Cancellous (or spongy) bone consists of an
open, interlacing framework of bony tissue. All bone develops first as cancellous bone, but some
bones, through further deposition of bone salts, become compact.

Bone growth is a complex restructuring process, involving both its destruction internally by bone
resorbing cells (osteoclasts, large multinucleate cells that function in bone dissolution) and its
deposition externally by bone building cells (osteoblasts, bone-forming cells).
Bone growth responds to several hormones, in particular parathyroid hormone from the
parathyroid gland, which stimulates bone resorption, and calcitonin from the thyroid gland,
which inhibits bone resorption. These two hormones, together with a derivative of vitamin D, are
responsible for maintaining a constant level of calcium in the blood.

6.1.4. Bone marrow function

Structure of long bone


60

Bone marrow, the soft pulpy tissue that fills bone cavities, contains a network of blood vessels
and fibers surrounded by fat and blood-producing cells. In children, the cells that give rise to
blood cells can be found throughout the marrow. In adults, these cells are found mostly in the red
marrow of the bones of the chest, hips, back, skull, and of the upper arms and legs.

The marrow in the long shafts of bones gradually loses its ability to manufacture blood. This
marrow, which is dominated by fat cells and takes on a yellowish color, is called yellow marrow.
This cross section of a long bone shows yellow bone marrow in the shaft of a long bone.

Hematopoiesis is the process of blood cell production (formation and development). All blood
cells initially begin their lives in the bone marrow of long bones within a vertebrate’s body.
Red marrow produces all of the body’s blood cells: red blood cells, white blood cells, and
platelets. Red blood cells in the circulatory system transport oxygen to body tissues and carbon
dioxide away from tissues. White blood cells are critical for fighting bacteria and other foreign
invaders of the body. Platelets are essential for the formation of blood clots to heal wounds.
61

Within red bone marrow, all blood cells originate from a single type of cell, called a
hematopoietic stem cell. Stimulated by hormones and growth factors, these stem cells divide to
produce immature or progenitor blood cells. Most of these progenitor cells remain in the stroma
and rapidly undergo series of cell divisions, producing either red blood cells or white blood cells.

At any one time, the stroma consists largely of progenitor cells in various stages of development.
At the appropriate developmental stage, new cells squeeze through the walls of the capillaries.
From there, the cells leave the bone and enter the body’s circulatory system. In mammals and
birds, red cells form continuously from large nucleated erythroblasts in red bone marrow (in
other vertebrates kidneys and spleen are the principal sites of red blood cell production).

6.2. Skeletal muscle structure and function

6.2.1. Skeletal muscle structure

The cell unit of the muscular tissue is called muscle fiber, specialized for contraction. When
viewed with a light microscope, striated muscle appears transversely striped (striated), with
alternating dark and light bands. The skeletal muscle is attached to skeletal elements and is
responsible for movements of the trunk, appendages, respiratory organs, eyes, mouthparts, and
other structure.

Skeletal muscle fibers are extremely long, cylindrical, multinucleate cells that may reach from
one end of the muscle to the other. They are packed into bundles called fascicles, which are
enclosed by tough connective tissue. The fascicles are in turn grouped into a discrete muscle
surrounded by a thick connective tissue layer. Skeletal muscle is sometimes called voluntary
muscle because it is stimulated by motor fibers and is under conscious cerebral control.

A skeletal muscle fiber contains many myofibrils, each consisting of functional units called
sarcomeres. The characteristic striations of a sarcomere are due to the arrangement of actin and
myosin filaments. This striation of whole fibers arises from the alternating dark and light bands
of the many smaller, threadlike myofibrils in each muscle fiber. Electron microscopy and
biochemical analysis show that these bands are due to the placement of the actin and myosin
muscle proteins within the myofibrils. Myosin occurs as thick filaments and actin as thin
62

filaments. The lightest region of a myofibril (I band) contains only actin, whereas the darkest
region (A band) contains both actin and myosin.

Structure of skeletal muscle


63

6.2.2. Muscle contraction mechanism and control

The functional (contractile) unit of a myofibril is the sarcomere, each of which extends from one
Z line to another Z line. During sarcomere contraction, the actin filaments slide past the myosin
filaments as they approach one another. This movement of actin in relation to myosin is called
the sliding-filament model of muscle contraction.

Each thin filament consists of two strands of actin, tropomyosin and the troponin complex. The
thin myofilament consists of a double strand of actin surrounded by two tropomyosin strands. A
globular protein complex, troponin, occurs in pairs at every seventh actin unit. The myosin
molecule is composed of two polypeptides coiled together and expanded at their ends into a
globular head. The thick myofilament is composed of a bundle of myosin molecules with the
globular heads extended outward.

Myosin contains globular


projections that attach to actin at
specific active binding sites,
forming attachments called cross-
bridges. Once cross-bridges form,
they exert a force on the thin actin
filament and cause it to move (by
ratchet mechanism).

Release of bond energy from ATP activates the myosin head, which swings 45 o, at the same time
releasing a molecule of ADP. This is the power stroke that pulls the actin filament a distance of
64

about 10 nm, and it comes to an end when another ATP molecule binds to the myosin head,
inactivating the site. Thus each cycle requires expenditure of energy in the form of ATP.

When a motor nerve conducts nerve impulses to skeletal muscle fibers, the fibers are stimulated
to contract via a motor unit. A motor unit consists of one motor nerve fiber and all the muscle
fibers with which it communicates. A space separates the specialized end of the motor nerve
fiber from the membrane (sarcolemma) of the muscle fiber.

The sarcolemma is normally polarized; the outside is positive, and the inside is negative. The
endoplasmic reticulum of muscle cells, called sarcoplasmic reticulum, releases calcium ions
(Ca2+), which diffuse into the cytoplasm. The calcium then binds with a regulatory protein called
troponin that is on another protein called tropomyosin. This binding exposes the myosin binding
sites on the actin molecule that tropomyosin had blocked. Once the binding sites are open, the
myosin filament can form cross-bridges with actin, and power strokes of cross-bridges result in
filament sliding and muscular contraction.

In a muscle fiber at rest, tropomyosin covers the myosin binding sites along the thin filament,
preventing actin and myosin from interacting. Calcium ions (Ca 2+) and proteins bound to actin
play a critical role in muscle cell contraction and relaxation. When Ca 2+ accumulates in the
cytosol, it binds to the troponin complex, causing the proteins bound along the actin strands to
shift position and expose the myosin-binding sites on the thin filament.
65

When the Ca2+ concentration rises in the


cytosol, the thin and thick filaments slide past
each other and the muscle fiber contracts.
When the Ca2+ concentration falls, the binding
sites are covered, and contraction stops.

Relaxation follows contraction. During relaxation, an active transport system pumps calcium
back into the sarcoplasmic reticulum for storage. By controlling the nerve impulses that reach the
sarcoplasmic reticulum, the nervous system controls Ca2+ levels in skeletal muscle tissue, thereby
exerting control over contraction.

Contracted and relaxed state of muscle

A is the relaxed state, and B the contracted state of muscle within sarcomere.
66

6.3. Interaction of muscles and skeleton for movement

6.3.1. Muscle and bone joints

Attachments between skeletal structures or to muscles are achieved by tendons and ligaments.
Tendon is an inextensible cord or band of tough white fibrous connective tissue that attaches a
muscle to a bone or other muscles. The main component of tendons and ligaments consists of the
collagen fibers. In tendon the collagenous fibers are extremely long and tightly packed together.
Achilles tendon connects the heel bone to the calf muscles.

Ligament is a sheet or band of tough fibrous tissue that usually connects bones to bones. Some
ligaments connect cartilage or support an organ, muscle, or other body part. Yellow elastic
ligaments consist primarily of elastic fibres and form relatively extensible ligaments joining
vertebrae, and true vocal cords; collagenous ligaments by contrast consist largely of parallel
bundles of collagen and resist extension.

6.3.2. Movement by muscle and skeleton

Skeletons function in support and protection as well as movement. Most land animals would
perish from their own weight if they had no skeleton to support them. Since the skeleton bears
the weight of the body, muscle mass is often concentrated in the upper appendages and girdles.
Many running mammals have little muscle in their lower leg that would slow leg movement.
Instead, tendons run from muscles high in the leg to cause movement at the lower joints. Muscles
can only contract; to be lengthened they must be extended by the pull of an antagonistic set of
muscles. Rigid skeletons provide the anchor points required by opposing sets of muscles, such as
flexors and extensors.

Although we call such muscles an antagonistic pair, their function is actually cooperative,
coordinated by the nervous system. For example, when you extend your arm, motor neurons
trigger your triceps muscle to contract while the absence of neuronal input allows your biceps to
relax. This arrangement works with either an internal skeleton, as in mammals, or an external
skeleton, as in insects. To move an animal in part or in whole, muscles must work in concert
with the skeleton.
67

The interaction of muscles and bones in upper limb

Because muscles
exert force only
during contraction,
moving a body part
back and forth
typically requires two
muscles attached to
the same section of
the skeleton. Back-
and-forth movement
of our body parts,
especially limbs, is
generally
accomplished by
antagonistic muscles.

Chapter 7. DIGESTIVE PROCESS AND NUTRITION

Almost all animals are heterotrophic organisms that depend on already synthesized organic
compounds of plants and other animals to obtain the materials they will use for growth,
maintenance, and reproduction of their kind. Since the food of animals, normally the complex
tissues of other organisms, is usually too massive to be absorbed directly by cells, it must be
broken down, or digested, into soluble molecules that are small enough to be used.

The ingestion of foods and their simplification by digestion are only initial steps in nutrition.
Foods reduced by digestion to soluble, molecular form are absorbed into the circulatory system
and transported to the body’s tissues. There they are assimilated into the structure of cells.
Oxygen is also transported by blood to the tissues, where food products are oxidized, or burned
to yield energy and heat. Food not immediately used is stored for future use. Wastes produced by
oxidation must be excreted. Food products unsuitable for digestion are egested in the form of
feces. This part will be focused particularly on the function of the alimentary canal in nutrition.
68

7.1. Food capture and mastication

Few animals can absorb nutrients directly from their external environments. Some blood
parasites, certain intestinal protozoan parasite, and tapeworms nourish themselves on primary
organic molecules absorbed directly across their body surfaces. Most animals however must
work for their meals. They are active feeders that have evolved numerous specializations for
obtaining food.

Fish, amphibians, and reptiles use their teeth principally to grip the prey and prevent its escape
until they can swallow it whole. True mastication, the chewing of food as opposed to tearing or
crushing, is found only among mammals. Mammals usually have four different types of teeth,
each adapted for specific functions.

Incisors are designed for biting, cutting, and stripping; canines are for seizing, piercing and
tearing; premolars and molars, at the back of the jaw, are for grinding and crushing. This basic
pattern is often greatly modified in animals having specialized food habits. For example,
herbivores have suppressed canines but well-developed molars with enamel ridges for grinding.

7.2. Mechanisms of digestion

Digestion is the process by which organic foods are mechanically and chemically broken into
small units for absorption. Although food solids consist principally of carbohydrates, proteins,
and fats, the very components that make up the body of the consumer, these components must
first be reduced to their simplest molecular units and dissolved before they can be assimilated.

7.2.1. Action of digestive enzymes

Mechanical processes of cutting and grinding by teeth and muscular mixing by the intestinal tract
are important in digestion. However, reduction of foods to small, absorbable units relies
principally on chemical breakdown by enzymes. Digestive enzymes are hydrolytic enzymes, or
hydrolases, so called because food molecules are split by the process of hydrolysis, breaking of a
chemical bond by adding the components of water across it:
69

There are specific enzymes for each class of organic compounds. These enzymes are located in
specific regions of the alimentary canal.

Mammalian digestive secretions, principal components, and pH

7.2.2. Motility in the alimentary canal

Food is moved through the digestive tract by cilia or by specialized musculature, and often by
both. In animals with well-developed digestive tube, the gut is usually lined with two opposing
layers of smooth muscle: a longitudinal layer (outer), in which the smooth muscle fibers run
parallel with the length of the gut, and a circular layer (inner), in which the muscle fibers
embrace the circumference of the gut.

The most characteristic gut movement in mammals is a kind of muscular action called peristalsis
that pushes the food down the gut with waves of involuntary contraction of circular muscle. This
peristalsis consists of alternating waves of contraction and relaxation in the smooth muscles
lining the canal. At some of the junctions between specialized compartments, the muscular layer
forms ringlike valves called sphincters which regulate the passage of material between
compartments.
70

7.3. Organization of the alimentary canal

In animals with tissue level of organization, the alimentary canal can be divided into five major
regions: (1) reception, (2) conduction and storage, (3) grinding and early digestion, (4) terminal
digestion and absorption, and (5) water absorption and concentration of solids.

7.3.1. Receiving region

The first region of the alimentary canal consists of devices for feeding and swallowing. These
include mouthparts (as in insects), buccal (oral) cavity and muscular pharynx. Most metazoans
have salivary glands (buccal glands) that produce lubricating secretions containing mucus to
assist swallowing. Salivary glands often have other specialized functions such as secretion of
toxic enzymes for quieting struggling prey and secretion of salivary enzymes to begin digestion.

In humans, ingestion and the initial steps of digestion occur in the mouth, or oral cavity.
Mechanical digestion begins when teeth of various shapes cut, smash, and grind food, making it
easier to swallow and increasing its surface area. Meanwhile, the presence of food stimulates a
nervous reflex that causes the salivary glands to deliver saliva through ducts to the oral cavity.
Saliva may also be released before food enters the mouth, triggered by a learned association
between eating and the time of day, a cooking odor, or another stimulus.

Saliva initiates chemical digestion while also protecting the oral cavity. Salivary amylase, an
enzyme in saliva, is found only in certain herbivorous mollusks, some insects, and in primate
mammals. This enzyme hydrolyzes starch (a glucose polymer from plants) and glycogen (a
glucose polymer from animals) into smaller polysaccharides and the disaccharide maltose. It
does not completely hydrolyze starch; when the food mass (bolus) is swallowed, salivary
amylase continues to act for some time, digesting perhaps half of the starch before the enzyme is
inactivated by the acidic environment of the stomach. Further starch digestion resumes beyond
the stomach in the intestine.

Mucin, a slippery glycoprotein (carbohydrateprotein complex) in saliva protects the lining of the
mouth from abrasion. Mucin also lubricates food for easier swallowing. Additional components
of saliva include buffers, which help prevent tooth decay by neutralizing acid, and antibacterial
agents (such as lysozyme), which protect against microorganisms that enter the mouth with food.
71

Bolus passage in oral cavity and throat

The tongue is a vertebrate innovation, usually attached to the floor of the mouth, that assists in
food manipulation and swallowing. It may be used for other purposes, however, such as food
capture (for example, chameleons, woodpeckers, anteaters) or as an olfactory sensor (many
lizards and snakes).

In humans, swallowing begins with the tongue pushing moistened food toward the pharynx. The
nasal cavity closes reflexively by raising the soft palate. As the food slides into the pharynx, the
epiglottis tips down over the trachea, nearly closing it. Some particles of food may enter the
opening of the trachea but contraction of laryngeal muscles prevents it from going farther. Once
food is in the esophagus, peristaltic contraction of esophageal muscles forces it smoothly toward
the stomach.

7.3.2. Conduction and storage region

The esophagus of vertebrates and many invertebrates serves to transfer food to the digestive
region. In many invertebrates (annelids, insects, octopods) the esophagus is expanded into a
crop, used for food storage before digestion. Among vertebrates, only birds have a crop. This
72

crop serves to store and soften food (grain, for example) before it passes to the stomach, or to
allow mild fermentation of food before it is regurgitated to feed nestlings.

7.3.3. Region of grinding and early digestion

In most vertebrates, and in some invertebrates, the stomach provides initial digestion as well as
storage and mixing of food with digestive juices. Mechanical breakdown of food, especially
plant food with its tough cellulose cell walls, often continues in herbivorous animals by grinding
and crushing devices in the stomach.

The woody cellulose that encloses plant cells can be broken down only by an enzyme, cellulase.
Animals cannot produce intestinal cellulase for the direct digestion of cellulose; however many
herbivores harbor microorganisms (bacteria and protozoa) in their gut (e.g. in ruminants) that do
produce cellulase. These microorganisms ferment cellulose under the anaerobic conditions of the
gut, producing fatty acids and sugars that the herbivore can use.

When food arrives at the stomach, the cardiac sphincter opens reflexively to allow the food to
enter, then closes to prevent regurgitation back into the esophagus. In humans, gentle peristaltic
waves pass over the filled stomach at the rate of approximately three each minute. Churning is
most vigorous at the intestinal end where food is steadily released into the duodenum, first
region of the small intestine. A pyloric sphincter regulates the flow of food into the intestine and
prevents regurgitation in the stomach.

Deep tubular glands in the stomach wall secrete gastric juice, in humans approximately 2 liters
each day. Two types of cells line these glands: chief cells, which secrete pepsin, and parietal
cells, which secrete hydrochloric acid.

Pepsin is a protease (protein-splitting enzyme) that acts only in an acid medium (pH 1.6 to 2.4).
This highly specific enzyme splits large proteins by preferentially breaking down certain peptide
bonds scattered along the peptide chain of the protein molecule. Although pepsin, because of its
specificity, cannot completely degrade proteins, it effectively hydrolyzes them into smaller
polypeptides. Other proteases that together can split all peptide bonds complete digestion of
protein in the intestine. Pepsin is present in the stomachs of nearly all vertebrates.
73

Rennin (not to be confused with renin, an enzyme produced by the kidney) is a milk-curdling
enzyme found in the stomach of ruminant mammals. It probably occurs in many other mammals.
By clotting and precipitating milk proteins (to favor bacterial action), it slows the movement of
milk through the stomach. Rennin extracted from stomachs of calves is used in making cheese.

7.3.4. Region of terminal digestion and absorption: the Intestine

a) Digestion in the vertebrate small intestine

Food is released into the small intestine through the pyloric sphincter, which relaxes at intervals
to allow entry of acidic stomach contents into the initial segment of the small intestine, the
duodenum. Two secretions pour into this region: pancreatic juice and bile. Both of these
secretions have high bicarbonate content, especially pancreatic juice, which effectively
neutralizes gastric acid, raising the pH of the liquefied food mass, now called chyme, from 1.5 to
7 as it enters the duodenum. This change in pH is essential because all intestinal enzymes are
effective only in a neutral or slightly alkaline medium.

Pancreatic enzymes

The pancreatic secretion of vertebrates contains several enzymes of major importance in


digestion. Two powerful proteases, namely trypsin and chymotrypsin, continue enzymatic
digestion of proteins begun by pepsin which is now inactivated by the alkalinity of the intestine.
Trypsin and chymotrypsin, like pepsin, are highly specific proteases that split apart peptide
bonds deep inside the protein molecule. The hydrolysis of a peptide linkage may be shown as:

Pancreatic juice also contains carboxypeptidase, which removes amino acids from the carboxyl
ends of polypeptides; pancreatic lipase, which hydrolyzes fats into fatty acids and glycerol;
pancreatic amylase, a starch-splitting enzyme identical to salivary amylase in its action; and
nucleases, which degrade RNA and DNA to nucleotides.
74

Membrane enzymes

The cells lining the intestine have digestive enzymes embedded in their surface membrane that
continue digestion of carbohydrates, proteins, and phosphate compounds. These enzymes of the
microvillus membrane include aminopeptidase that splits terminal amino acids from the amino
end of short peptides, and several disaccharidases, enzymes that split 12-carbon sugar molecules
into 6-carbon units. The disaccharidases include maltase, which splits maltose into two
molecules of glucose; sucrase, which splits sucrose to fructose and glucose; and lactase, which
breaks lactose (milk sugar) into glucose and galactose. Also present is alkaline phosphatase, an
enzyme that attacks a variety of phosphate compounds.

Table summarizing the major digestive events

Secretory Main
Region of region or components & Substrate/ Product of
action and pH organs Secretion enzymes Function digestion
Smaller
Starch or polysaccahrides
Amylase
Salivary Glycogen Maltose
Mouth
Mouth glands or (disaccharide)
(pH=6.5)
saliva Mucin Protecting mouth lining; lubrication of
(glucoprotein) bolus
Bicarbonate Neutralizes acidity  
Stomach Smaller
(tubular Pespin Protein polypeptides
Stomach glands: chief Hydrochloric acid
Gastric juice
(pH=1.6-2.4) cells & (HCl) Providing acid medium to stomach
parietal Rennin (in young Milk-curdling (to favor bacterial action);
cells) ruminants) slow movement of milk
Emulsification of fats (breakdown into
Bile salts tiny droplets)
Small Liver &
Bile Bile pigments Components of bile
intestine gallbladder
Cholesterol Formation of bile acid
(pH=7-8)
Bicarbonate Neutralizes gastric acidity (in duodenum)
Pancreas Pancreatic Trypsin Completes digestion of proteins started
75

by pepsin
Completes digestion of proteins started
Chymotrypsin by pepsin
Removes a.a. from carboxyl ends of
Carboxypeptidase polypeptides
Fatty acids &
juice Fats glycerol
Lipase
Completes digestion of fats started by
the bile
Continues and completes digestion of
Amylase starch and glucogen started in mouth
Nucleases DNA & RNA Nucleotides
Bicarbonate Neutralizes gastric acidity
Removes terminal a.a. from amino ends
Aminopeptidase of short peptides
2 glucose
Small Maltase Maltose molecules
Membrane
intestine Glucose &
enzymes
walls Lactase Lactose galactose
Sucrase Sucrose Fructose & Glucose
Alkaline Phosphate
phosphatase compounds Phoshate
Bile and fat digestion

The liver secretes bile into the bile duct, which drains into the upper intestine (duodenum).
Between meals bile collects in the gallbladder, an expansible storage sac that releases bile when
stimulated by the presence of fatty food in the duodenum. Bile contains water, bile salts, and
pigments, but no enzymes. Bile salts are essential for digestion of fats, which are especially
resistant to enzymatic digestion because of their tendency to remain in water-insoluble globules.

Bile salts reduce surface tension of fat globules, allowing the churning action of the intestine to
break fats into tiny droplets (emulsification). With total surface exposure of fat particles greatly
increased, fat-splitting lipases are able to reach and hydrolyze the triglyceride molecules. The
yellow-green color of bile is produced by bile pigments, breakdown products of hemoglobin
from worn-out red blood cells. Bile pigments also give the feces its characteristic color.

Bile production is only one of the liver’s many functions. This highly versatile organ is a
storehouse for glycogen, production center for plasma proteins, site of protein synthesis and
detoxification of protein wastes, site for destruction of worn-out red blood cells, and center for
metabolism of fat, amino acids, and carbohydrates.
76

b) Intestine structure

Devices for increasing the internal surface area of the intestine are highly developed in
vertebrates, but are generally absent among invertebrates. Perhaps the most direct way to
increase the absorptive surface of the gut is to increase its length. Coiling of the intestine is
common among all vertebrate groups and reaches its highest development in mammals, in which
the length of the intestine may exceed eight times the length of the body.

Nearly all vertebrates have developed elaborate folds (amphibians and reptiles) and minute
fingerlike projections called villi (birds and mammals), which give the inner surface of fresh
intestinal tissue the appearance of velvet. The electron microscope reveals that each cell lining
the intestinal cavity additionally is bordered by hundreds of short, delicate processes called
microvilli. These processes, together with larger villi and intestinal folds, may increase the
internal surface area of the intestine more than a million times as compared to a smooth cylinder
of the same diameter. This elaborate surface greatly facilitates the absorption of food molecules.

Structure of the Microvilli borne on


intestine lining intestinal epithelium

c) Absorption
77

Little food is absorbed in the stomach because digestion is still incomplete and because of
limited absorptive surface area. However, some materials, such as drugs and alcohol, are
absorbed mostly there, which contributes to their rapid action. Most digested food is absorbed
from the small intestine where the numerous finger-shaped villi provide an enormous surface
area through which materials can pass from the intestinal lumen into the circulation.

Carbohydrates are absorbed almost exclusively as simple sugars (monosaccharides, for example,
glucose, fructose, and galactose) because the intestine is virtually impermeable to
polysaccharides. Proteins are absorbed principally as their amino acid subunits, although a
limited amount of small proteins or peptide fragments sometimes may be absorbed. Both active
and passive processes transfer simple sugars and amino acids across the intestinal epithelium.

Immediately after a meal these materials are in such high concentration in the gut that they
readily diffuse into the blood, where their concentration is initially lower. However, if absorption
were passive only, we would expect transfer to cease as soon as concentrations of a substance
became equal on both sides of the intestinal epithelium.

Passive transfer alone would permit valuable nutrients to be lost in the feces. In fact, very little is
lost because passive transfer is supplemented by an active transport mechanism located in the
epithelial cells that transfers food molecules into the blood. Materials thus are moved against
their concentration gradient, a process requiring expenditure of energy. Although not all food
products are actively transported, those that are, such as glucose, galactose, and most amino
acids, are handled by transport mechanisms that are specific for each kind of molecule.

As mentioned previously, fat droplets are emulsified by bile salts and then digested by pancreatic
lipase. Triglycerides are broken into fatty acids and monoglycerides, which complex with bile
salts to form minute droplets called micelles. When micelles contact the microvilli of the
intestinal epithelium, the fatty acids and monoglycerides are absorbed by simple diffusion.

They then enter the endoplasmic reticulum of the absorptive cells, where they are resynthesized
into triglycerides before passing into lacteals. From the lacteals, fat droplets enter the lymph
system and eventually pass into the blood circulation through the thoracic duct. After a fatty
78

meal, even a peanut butter sandwich, the presence of numerous fat droplets in the blood imparts
a milky appearance to the blood plasma.

7.3.5. Region of water absorption and concentration of solids

The large intestine consolidates the indigestible remnants of digestion by reabsorption of water
to form solid or semisolid feces for removal from the body by defecation. Specialized rectal
glands in insects absorb water and ions as needed, leaving behind fecal pellets that are almost
completely dry. In reptiles and birds, which also produce nearly dry feces, most of the water is
reabsorbed in the cloaca. White pastelike feces are formed containing both indigestible food
wastes and uric acid.

The colon of humans contains enormous numbers of bacteria, which first enter the sterile colon
of the newborn infant with its food. Egested feces contain numerous harmless bacteria as well as
bacteria that can cause serious illness that otherwise should infest the abdomen or bloodstream.
Normally the body’s defenses prevent invasion of such bacteria. Bacteria degrade organic wastes
in the feces and provide some nutritional benefit by synthesizing certain vitamins (vitamin K and
small quantities of some of the B vitamins), which are absorbed by the body.

7.4. Regulation of food intake and digestion

A hunger center located in the hypothalamus of the brain regulates the intake of food. A drop in
the blood glucose level stimulates a craving for food. Fat stores are supervised by the
hypothalamus, which may be set at a point higher or lower than the norm. A high setting can be
lowered somewhat by exercise, but as dieters are painfully aware, the body defends its fat stores
with remarkable tenacity.

The hormone, called leptin, appears to operate through a feedback system that tells the
hypothalamus how much fat the body carries. If levels are high, release of leptin by fat cells
leads to diminished appetite and increased thermogenesis. The discovery of leptin has initiated a
flurry of research on obesity and a resurgence of commercial interest in producing a weight-loss
drug based on leptin.
79

The digestive process is coordinated by a family of hormones produced by the body’s most
diffuse endocrine tissue, the gastrointestinal tract. These hormones are examples of the many
substances produced by the vertebrate body that have hormonal function, yet are not necessarily
produced by discrete endocrine glands.

Among the principal Gastro-intestinal (GI) hormones are gastrin, cholecystokinin (CCK), and
secretin. Gastrin is a small polypeptide hormone produced by endocrine cells in the pyloric
portion of the stomach. Gastrin is secreted in response to stimulation by the parasympathetic
nerve endings, and when protein food enters the stomach. Its main actions are to stimulate
hydrochloric acid secretion and to increase gastric motility. Gastrin is an unusual hormone in that
it exerts its action on the same organ from which it is secreted.

CCK, also a polypeptide hormone with a striking structural resemblance to gastrin, is secreted by
endocrine cells in the walls of the upper small intestine in response to the presence of fatty acids
and amino acids in the duodenum. It has at least three distinct functions. It stimulates gallbladder
contraction and thus increases the flow of bile salts into the intestine; it stimulates an enzyme-
rich secretion from the pancreas; and it acts on the brain to contribute a feeling of satiety after a
meal, particularly one rich in fats.

Secretin is produced by endocrine cells in the duodenal wall. It is secreted in response to food
and strong acid in the stomach and small intestine, and its principal action is to stimulate the
release of an alkaline pancreatic fluid that neutralizes stomach acid as it enters the intestine. It
also aids fat digestion by inhibiting gastric motility and increasing production of an alkaline bile
secretion from the liver.

7.5. Nutritional requirements

7.5.1. Categories of nutrients

The food of animals must include carbohydrates, proteins, fats, water, mineral salts, and
vitamins. Carbohydrates and fats are required as fuels for energy and for the synthesis of various
substances and structures. Proteins (actually the amino acids of which they are composed) are
needed for the synthesis of specific proteins and other nitrogen-containing compounds. Water is
required as the solvent for body chemistry and as a major component of all fluids of the body.
80

Inorganic salts are required as the anions and cations of body fluids and tissues and form
important structural and physiological components throughout the body. Vitamins are accessory
factors from food that are often built into the structure of many enzymes.

A vitamin is a relatively simple organic compound that is not a carbohydrate, fat, protein, or
mineral and that is required in very small amounts in the diet for some specific cellular function.
For humans, 13 essential vitamins have been identified. Depending on the vitamin, the required
amount ranges from about 0.01 to 100 mg per day. Vitamins are not sources of energy but are
often associated with the activity of important enzymes that serve vital metabolic roles.

Vitamin B2 is converted in the body to FAD, a coenzyme used in many metabolic processes,
including cellular respiration. Vitamins are classified as water-soluble or fat-soluble. The water-
soluble vitamins include the B complex, which are compounds that generally function as
coenzymes, and vitamin C which is required to produce connective tissue.

Among the fat-soluble vitamins are vitamin A, which is incorporated into visual pigments of the
eye, and vitamin K, which functions in blood clotting. Another is vitamin D, which aids in
calcium absorption and bone formation. Our dietary requirement for vitamin D is variable
because we synthesize this vitamin from other molecules when the skin is exposed to sunlight.

Almost all animals, vertebrate and invertebrate, require B vitamins. The dietary need for vitamin
C and the fat-soluble vitamins A, D, E, and K is mostly restricted to vertebrates. Requirements
for those vitamins are relative even within groups of close relationship: a rabbit does not require
vitamin C, but guinea pigs and humans do; some songbirds require vitamin A, but others do not.
Generally vitamins represent synthetic gaps in the metabolic machinery of animals.

Dietary minerals are inorganic nutrients, such as zinc and potassium, that are usually required in
small amounts-from less than 1 mg to about 2,500 mg per day. Mineral requirements vary among
animal species. For example, humans and other vertebrates require relatively large quantities of
calcium and phosphorus for building and maintaining bone. In addition, calcium is necessary for
the functioning of nerves and muscles, and phosphorus is an ingredient of ATP and nucleic
acids.
81

Iron is a component of the cytochromes that function in cellular respiration and of hemoglobin,
the oxygen-binding protein of red blood cells. Many minerals are cofactors built into the
structure of enzymes; magnesium, for example, is present in enzymes that split ATP. Vertebrates
need iodine to make thyroid hormones, which regulate metabolic rate. Sodium, potassium, and
chloride ions are important in the functioning of nerves and in maintaining osmotic balance
between cells and the surrounding body fluid.

7.5.2. Essential nutrients

The recognition years ago that many human diseases and those of domesticated animals were
caused by or associated with dietary deficiencies led biologists to search for specific nutrients
that would prevent such diseases. These studies eventually yielded a list of essential nutrients
for human beings and other animal species studied.

Essential nutrients are those needed for normal growth and maintenance and that must be
supplied in the diet. In other words, it is “essential” that these nutrients be in the diet because the
animal cannot synthesize them from other dietary constituents. Nearly 30 organic compounds
(amino acids and vitamins) and 21 elements are essential for humans.

Human nutrient requirements

Water-soluble Minerals Amino acids Polyunsaturated


vitamins Major Trace Phenylalanine fatty acids
Thiamine (B1) Calcium Iron Lysine Arachidonic
Riboflavin (B2) Phosphorus Fluorine Isoleucine Linoleic
Niacin (nicotinic Sulfur Zinc Leucine Linolenic
acid) Potassium Copper Valine
Pyridoxine (B6) Chlorine Silicon Methionine
Pantothenic acid Sodium Vanadium Tryptophan
Folacin (folic acid) Magnesium Tin Threonine
Vitamin B12 Nickel Arginine*
(cobalamin) Selenium Histidine*
Biotin Manganese
Ascorbic acid (C) Iodine
Molybdenum
Fat-soluble Chromium
vitamins Cobalt
82

A, D, E, and K

Lipids are needed principally to provide energy. However, at least three fatty acids are essential
for humans because we cannot synthesize them. The arachidonic acid found in most animal fats
is a precursor of prostaglandins. The three fatty acids are obtained from plants (seeds, fruits,
plant oils, legumes…). Much interest and research have been devoted to lipids in our diets
because of the association between fatty diets and the disease atherosclerosis. The matter is
complex, but evidence suggests that atherosclerosis may occur when the diet is high in saturated
lipids but low in polyunsaturated lipids.

Proteins are not themselves the essential nutrients but rather contain essential amino acids. Of the
20 amino acids commonly found in proteins, eight and possibly 10 are essential to humans. We
can synthesize the rest from other amino acids. All eight of the essential amino acids must be
present simultaneously in the diet for protein synthesis. Each plant complements the other by
having adequate amounts of those amino acids that are deficient in the other. Because animal
proteins are rich in essential amino acids, they are in great demand in all countries. Their
uncontrolled consumption leads to risks of the so-called ‘diseases of affluence’: heart disease,
stroke, and certain kinds of cancer.
83

Chapter 8. BLOOD CIRCULATION AND FUNCTION

Blood is a red body fluid that is pumped by the heart and circulates throughout the body via
specialized vessels (arteries, veins, capillaries) in which a fluid matrix called plasma suspends
specialized red blood cells, white blood cells, and platelets. Blood transports various substances
(oxygen, carbon dioxide, nutrients, wastes, hormones, minerals, vitamins, and other substances)
throughout the bodies of animals.

In a closed circulatory system of vertebrates, blood never leaves the blood vessels. Blood moves
from the heart, through arteries, arterioles, capillaries, venules, veins, and back to the heart.
Exchange between the blood and extracellular fluid only occurs at the capillary level. In general,
the role of vertebrate blood is to transport oxygen, carbon dioxide, and nutrients; defend the body
against harmful microorganisms, cells, and viruses; prevent blood loss through coagulation
(clotting); and contribute in regulating body temperature and pH.

8.1. Blood composition

Plasma, straw-colored, is the liquid part of blood. In mammals, plasma is about 90% water and
provides the solvent for dissolving and transporting nutrients. A group of proteins (albumin,
fibrinogen, and globulins) comprises another 7% of the plasma. Albumin represents about 60%
of the total plasma proteins. Fibrinogen is necessary for blood coagulation (clotting), and the
globulins include the immunoglobulins and various metal-binding proteins. Serum is plasma
from which the proteins involved in blood clotting have been removed. The gamma globulin
portion functions in the immune response because it consists mostly of antibodies. The
remaining 3% of plasma is composed of electrolytes, amino acids, glucose and other nutrients,
various enzymes, hormones, metabolic wastes, and traces of many inorganic and organic
molecules.

The formed-element fraction (cellular component) of vertebrate blood consists of erythrocytes


(red blood cells; RBCs), leukocytes (white blood cells; WBCs), and platelets (thrombocytes).

White blood cells (leukocytes) are present in lower number than are red blood cells, generally
being 1 to 2% of the blood by volume. White blood cells are divided into agranulocytes (without
granules in the cytoplasm) and granulocytes (have granules in the cytoplasm). The two types of
84

agranulocytes are lymphocytes and monocytes. The three types of granulocytes are eosinophils,
basophils, and neutrophils. Fragmented cells are called platelets (thrombocytes). WBCs are
scavengers that destroy microorganisms at infection sites, remove foreign chemicals, and remove
debris that results from dead or injured cells. All WBCs are derived from immature cells (called
stem cells) in bone marrow by a process called hematopoiesis.

Formed elements of
human blood

Among the granulocytes, eosinophils are phagocytic, and ingest foreign proteins and immune
complexes rather than bacteria. In mammals, eosinophils also release chemicals that counteract
the effects of certain inflammatory chemicals released during allergic reactions. When basophils,
the least numerous WBCs, react with a foreign substance, their granules release histamine
(causes blood vessels to dilate and leak fluid at a site of inflammation) and heparin (prevents
blood clotting). Neutrophils, the most numerous of the white blood cells, are chemically attracted
to sites of inflammation and are active phagocytes.

Two distinct types of lymphocytes are B cells and T cells, both of which are central to the
immune response. B cells originate in the bone marrow and colonize the lymphoid tissue, where
they mature. In contrast, T cells are associated with and influenced by the thymus gland before
they colonize lymphoid tissue and play their role in the immune response, especially in
combating viral infections and cancers. B cells create antibodies in response to a specific antigen.
When B cells are activated, they divide and differentiate to produce plasma cells (lymphocytes
that produce antibodies).
85

Platelets or thrombocytes are disk-shaped cell fragments that initiate blood clotting. When a
blood vessel is injured, platelets immediately move to the site and clump, attaching themselves to
the damaged area, and thereby beginning the process of blood coagulation.

Red blood cells (or erythrocytes) vary dramatically in size, shape, and number in the different
vertebrates. RBCs of most vertebrates are nucleated, but mammalian RBCs are enucleated (i.e.
lack a nucleus). Some fishes and amphibians also have enucleated RBCs. Most mammalian
RBCs are biconcave disks. The shape of a biconcave disk provides a larger surface area for gas
diffusion than a flat disk or sphere.

Almost the entire mass of a RBC consists of hemoglobin, an iron-containing protein. The major
function of an erythrocyte is to pick up oxygen from the environment, bind it to hemoglobin to
form oxyhemoglobin, and transport it to body tissues. Blood rich in oxyhemoglobin is bright red.
Hemoglobin also carries waste carbon dioxide (in the form of carbaminohemoglobin) from the
tissues to the lungs (or gills) for removal from the body.

8.2. Open and closed circulations

Many invertebrates have an open circulation in which there are no small blood vessels or
capillaries connecting arteries with veins. In insects and other arthropods, in most molluscs, and
in many smaller invertebrate groups blood sinuses, collectively called a hemocoel, replace
capillary beds found in animals with closed systems. Blood in those animal groups called
hemolymph freely circulates and is so-called because there is no separation of the extracellular
fluid into blood plasma and lymph, thus the blood volume is large and may constitute 20% to
40% of body volume while by contrast, blood volume in animals with closed circulations
(vertebrates, for example) is only about 5% to 10% of body volume.

All closed systems have certain features in common. A heart pumps blood into arteries that
branch and narrow into arterioles and then into a vast system of capillaries. Blood leaving
capillaries enters venules and then veins that return the blood to the heart. Capillary walls are
thin, permitting rapid rates of transfer of materials between blood and tissues.

Closed systems are more suitable for large and active animals because blood can be moved
rapidly to tissues needing it. In addition, flow to various organs can be readjusted to meet
86

changing needs by varying the diameters of blood vessels. Because blood pressures are much
higher in closed than in open systems, fluid is constantly filtered across capillary walls into the
surrounding tissue spaces. Most of this fluid is drawn back into capillaries by osmosis. The
remainder is recovered by the lymphatic system which has evolved in parallel with the high-
pressure system of vertebrates.

8.3. Mammalian heart

8.3.1. Heart structure and function

Vertebrates have developed a high-pressure double circulation: a systemic circuit that provides
oxygenated blood to the capillary beds of the body organs; and a pulmonary circuit that serves
the lungs. The four-chambered mammalian heart is a muscular organ located in the thorax and
covered by a tough, fibrous sac, the pericardium.

Blood returning from the lungs collects in the left atrium, passes into the left ventricle, and is
pumped into the body (systemic) circulation. Blood returning from the body flows into the right
atrium, and passes into the right ventricle, which pumps it into the lungs. Backflow of blood is
prevented by two sets of valves that open and close passively in response to pressure differences
between the heart chambers. The bicuspid (between left atrium and ventricle) and tricuspid
(between right atrium and ventricle) valves separate the cavities of the atrium and ventricle in
each half of the heart. Where the great arteries, the pulmonary from the right ventricle and the
aorta from the left ventricle, leave the heart, semilunar valves prevent backflow into the
ventricles.

Contraction is called systole, and relaxation, diastole. When the atria contract (atrial systole), the
ventricles relax (ventricular diastole), and ventricular systole is accompanied by atrial diastole.

Rate of the heartbeat depends on age, sex, and especially exercise. Exercise may increase
cardiac output (volume of blood forced from either ventricle each minute) more than fivefold.
Both heart rate and stroke volume increase. Heart rates among vertebrates vary with general
level of metabolism and body size. Ectothermic codfish have a heart rate of approximately 30
beats per minute; endothermic rabbits of about the same weight have a rate of 200 beats per
minute. Small animals have higher heart rates than do large animals.
87

Human heart

In human heart, deoxygenated blood enters right side of heart and is pumped to the lungs.
Oxygenated blood returning from the lungs enters left side of the heart and is pumped to the
body. The left ventricular wall is thicker than that of the right ventricle, which needs less
muscular force to pump blood into the nearby lungs.

8.3.2. Coronary circulation

As a very active organ, the heart may need a generous blood supply of its own. In birds and
mammals, however, the thickness of the heart muscle and its high rate of metabolism require that
the heart have its own vascular supply, the coronary circulation. Coronary arteries divide to
88

form an extensive capillary network surrounding the muscle fibers and provide them with
oxygen and nutrients.

Heart muscle has an extremely high oxygen demand. Even at rest the heart removes 70% of
oxygen from the blood, in contrast to most other body tissues, which remove only about 25%.
Therefore, an increase in the work of the heart must be met by a massive increase in coronary
blood flow. Any reduction in coronary circulation due to partial or complete blockage (coronary
artery disease) may lead to a heart attack (myocardial infarction) in which heart cells die from
lack of oxygen.

8.3.3. Control of the heart

The vertebrate heart is a muscular pump composed of cardiac muscle. Cardiac muscle resembles
skeletal muscle (both are types of striated muscle) but the cells are branched and joined end-to-
end by junctional complexes to form a complex branching network. Unlike skeletal muscle,
vertebrate cardiac muscle does not depend on nerve activity to initiate a contraction. Instead,
regular contractions are established by specialized cardiac muscle cells, called pacemaker cells.

The pacemaker is a specialized region in heart-muscle tissue located in the wall of the upper
right chamber of the heart that sends out rhythmic electrical impulses to regulate the heartbeat.
Electrical activity initiated in the pacemaker spreads over the muscle of the two atria and then,
after a slight delay, to the muscle of the ventricles.

The cardiac control center in the brain is located in the medulla and connects to two sets of
nerves. Impulses sent along one set, the vagus nerves, apply a braking action to the heart rate,
and impulses sent along the other set, the accelerator nerves, speed it up. Both sets of nerves
terminate in the sinus node, thus guiding the activity of the pacemaker.

8.4. Prevention of blood loss

It is essential that animals have ways of preventing rapid loss of body fluids after an injury.
When a vessel is damaged, smooth muscle in the wall of the vessel contracts, which causes the
vessel lumen to narrow, sometimes so strongly that blood flow is completely stopped.
89

In vertebrates blood coagulation is the dominant hemostatic defense. Blood clots form as a
tangled network of fibers from one of the plasma proteins, fibrinogen. The transformation of
fibrinogen into a fibrin meshwork that entangles blood cells to form a gel-like clot is catalyzed
by the enzyme thrombin. Thrombin is present in blood in an inactive form called prothrombin,
which must be activated for coagulation to occur. In this process, blood platelets play a vital role.

Platelets form in red bone marrow from certain large cells that regularly pinch off bits of their
cytoplasm; thus they are fragments of cells. When the normally smooth inner surface of a blood
vessel is disrupted, either by a break or by deposits of a cholesterol-lipid material, platelets
rapidly adhere to the surface and release thromboplastin and other clotting factors (including
Ca2+ and Vitamin K). These factors, along with factors released from damaged tissue and with
calcium ions, initiate conversion of prothrombin to active thrombin.

Left: Fibrin function in blood clotting

Right: Formation of fibrin

Clotting is initiated after tissue damage by disintegration of platelets in blood, resulting in a


complex series of intravascular reactions that end with conversion of a plasma protein,
fibrinogen, into long, tough, insoluble polymers of fibrin. Fibrin and entangled erythrocytes form
the blood clot, which arrests bleeding.

Several kinds of clotting abnormalities in humans are known. One of these, hemophilia is a
condition characterized by failure of blood to clot, so that even insignificant wounds can cause
continuous severe bleeding. It is caused by a rare mutation (the condition occurs in about 1 in
10,000 males) on the X sex chromosome, resulting in an inherited lack of one of the platelet
factors in males and in homozygous females.
90

8.5. Blood groups and transfer

Blood cells differ chemically from person to person, and when two different (incompatible)
blood types are mixed, agglutination (clumping together) of erythrocytes results. The basis of
these chemical differences is naturally occurring antigens on the membranes of red blood cells.
An antigen is a substance, usually a protein, on the surface of a cell or bacterium that stimulates
the production of an antibody. An antibody is a protein produced by B cells in the body in
response to the presence of an antigen. Antibodies are a primary form of immune response in
resistance to disease and act by attaching themselves to a foreign antigen and weakening or
destroying it.

The best known of these inherited immune systems is the ABO blood group. Antigens A and B
are inherited as codominant alleles of a single gene. Homozygotes for a recessive allele at the
same gene have type O blood, which lacks A and B antigens. Type A persons acquire anti-B
antibodies soon after birth, even without exposure to type B cells. Similarly, type B individuals
come to carry anti-A antibodies at a very early age. Type AB blood has neither anti-A nor anti-B
antibodies (since if it did, it would destroy its own blood cells), and type O blood has both anti-A
and anti-B antibodies.

Blood groups, properties, and transfer possibilities

Blood group Antigens Antibodies in serum Can give Can receive


on RBCs blood to blood from
O None Anti-A and Anti-B All O
A A Anti-B A, AB O, A
B B Anti-A B, AB O, B
AB AB None AB All

We see then that the blood-group names identify their antigen content. Persons with type O
blood are called universal donors because, lacking antigens, their blood can be infused into a
person with any blood type. Even though it contains anti-A and anti-B antibodies, these are so
diluted during transfusion that they do not react with A or B antigens in a recipient. Persons with
AB blood are universal recipients because they lack antibodies to A and B antigens.
91

8.6. Lymphatic system

The lymphatic system of vertebrates is an extensive network of thin-walled vessels that arise as
blind-ended lymph capillaries in most tissues of the body. These unite to form a treelike structure
of increasingly larger lymph vessels, which finally drain into veins in the lower neck.

Lymph is the extracellular fluid that accumulates in the lymph vessels. Lymphatic capillaries are
in direct contact with the extracellular fluid surrounding tissues. The small lymphatic capillaries
merge to form larger lymphatic vessels called lymphatics. Lymphatics are thin-walled vessels
with valves that ensure the one-way flow of lymph, i.e. to the heart. These vessels pass through
the lymph nodes on their way back to the heart. Lymph nodes concentrate in several areas of the
body and play an important role in the body’s defense against disease as they produce white
blood cells, esp. lymphocytes that are responsible for immunity.

The lymphatic system has four major functions: (1) to collect and drain most of the fluid that
seeps from the bloodstream and accumulates in the extracellular fluid; (2) to return small
amounts of proteins that have left the cells; (3) to transport lipids that have been absorbed from
the small intestine; and (4) to transport foreign particles and cellular debris to disposal centers
(lymph nodes).

A principal function of the lymphatic system is to return to the blood the excess fluid (lymph)
filtered across capillary walls into interstitial spaces. Lymph is similar to plasma but has a much
lower concentration of protein. Large molecules, especially fats absorbed from the gut, also
reach the circulatory system by way of the lymphatic system.

The lymphatic system also plays a central role in the body’s defenses. Located at intervals along
the lymph vessels are lymph nodes that have several defense-related functions. Cells in the
lymph glands such as macrophages remove foreign particles, especial bacteria, which might
otherwise enter the general circulation. They are also centers (together with bone marrow and
thymus gland) for production, maintenance, and distribution of lymphocytes that produce
antibodies, essential components of the body’s defense mechanisms
92

Chapter 9. RESPIRATORY SYSTEM AND BREATHING

9.1. Oxygen transport in animal respiration

Energy bound up in food is released by oxidative processes, usually with molecular oxygen as
the terminal electron acceptor. Oxygen for this purpose is taken into the body across some
respiratory surface. Physiologists find it is convenient to distinguish two separate but interrelated
respiratory processes: cellular respiration, the oxidative processes that occur within cells, and
external respiration, the exchange of oxygen and carbon dioxide between the organism and its
environment.

In single-celled organisms, oxygen is acquired and carbon dioxide liberated by direct diffusion
across surface membranes. Gas exchange by diffusion alone is possible only for very small
organisms less than 1 mm in diameter, where diffusion paths are short and the surface area of the
organism is large relative to volume. As animals became larger and evolved a waterproof
covering, specialized devices such as lungs and gills evolved to increase the effective surface for
gas exchange. But, because gases diffuse so slowly through living tissue, a circulatory system
was necessary to distribute gases to and from the deep tissues of the body.

9.2. Respiratory structures or organs

9.2.1. Gas exchange by direct diffusion

Protozoa, sponges, cnidarians, and many worms respire by direct diffusion of gases between
organism and environment. It is a kind of cutaneous respiration. Cutaneous respiration frequently
supplements gill or lung breathing in larger animals such as amphibians and fishes.

9.2.2. Gas exchange through tubes: tracheal systems

Insects and certain other terrestrial arthropods (centipedes, millipedes, and some spiders) have a
highly specialized type of respiratory system, consisting of a branching system of tubes
(tracheae) that extends to all parts of the body. Air enters the tracheal system through valvelike
openings (spiracles). Carbon dioxide diffuses out through spiracles.
93

9.2.3. Efficient exchange in water: gills

Gills of various types are effective respiratory devices for life in water. Fish gills are thin
filamentous structures, richly supplied with blood vessels arranged so that blood flow is opposite
to the flow of water across the gills (countercurrent flow). For extracting oxygen, water flows
over the gills in a steady stream, pulled and pushed by an efficient, two-valved, branchial pump.
Gill ventilation is often assisted by the fish’s forward movement through the water.

9.2.4. Lungs

Most air-breathing vertebrates possess lungs. Lungs are also found in certain invertebrates
(pulmonate snails, scorpions, some spiders, some small crustaceans), but these structures cannot
be very efficiently ventilated. The total surface available for gas exchange is much increased in
lungs of reptiles which are subdivided into numerous interconnecting air sacs. Most elaborate of
all are mammalian lungs complexes of millions of small sacs, called alveoli, each covered by a
rich vascular network. A large surface area is essential for the high oxygen uptake required to
support the elevated metabolic rate of endothermic mammals.

9.3. Mammalian respiratory system

9.3.1. Respiratory system structure

Air enters a mammalian respiratory system through nostrils (external nares), passes through a
nasal chamber, lined with mucus-secreting epithelium, and then through internal nares, nasal
openings connected to the pharynx. Here, where pathways of digestion and respiration cross,
inhaled air leaves the pharynx by passing into a narrow opening, the glottis, while food enters
the esophagus to pass to the stomach. The glottis opens into the larynx, or voice box, and then
into the trachea, or windpipe.

The trachea branches into two bronchi, one to each lung. Within the lungs each bronchus divides
and subdivides into small tubes (bronchioles) that lead via alveolar ducts to the air sacs
(alveoli). The single-layered endothelial walls of the alveoli are thin and moist to facilitate
exchange of gases between air sacs and adjacent blood capillaries.
94

The lungs consist of a great deal of elastic connective tissue and some muscle. They are covered
by a thin layer of tough epithelium known as the visceral pleura. A similar layer, the parietal
pleura, lines the inner surface of the walls of the chest. The two layers of the pleura are in
contact and slide over one another as the lungs expand and contract.

The “space” between the pleura, called the pleural cavity, maintains a partial vacuum, which
helps keep the lungs expanded to fill the pleural cavity. Therefore no real pleural space exists;
the two pleura rub together, lubricated by tissue fluid (lymph). The chest cavity is bounded by
the spine, ribs, and breastbone, and floored by the diaphragm, a dome-shaped, muscular partition
between the chest cavity and abdomen. A muscular diaphragm is found only in mammals.

Structure of human lungs and alveoli

A: Lungs of human with right lung shown in section. B: Terminal portion of bronchiole showing
air sacs with their blood supply. Arrows show direction of blood flow.

Air passageways are lined with both mucus-secreting and ciliated epithelial cells, which play an
important role in conditioning the air before it reaches the alveoli. Partial cartilage rings in the
walls of the tracheae, bronchi, and even some of the bronchioles prevent those structures from
collapsing. In its passage to the air sacs, air undergoes three important changes: (1) it is filtered
free from most dust and other foreign substances, (2) it is warmed to body temperature, and (3) it
is saturated with moisture.
95

9.3.2. Ventilating the lungs

The chest cavity is an air-tight chamber. Inspiration pulls the ribs upward, flattens the
diaphragm, and enlarges the chest cavity. The resultant increase in volume of the chest cavity
causes air pressure in the lungs to fall below atmospheric pressure: air rushes in through
passageways to equalize the pressure. Normal expiration is a less active process than inspiration.
When the muscles relax, the ribs and diaphragm return to their original position, and the chest
cavity decreases in size, the elastic lungs deflate, and air exits.

Unlike other respiratory organs, gas in lungs is exchanged between blood and air only in the
alveoli, located at the ends of a branching tree of air tubes (trachea, bronchi, and bronchioles); air
must enter and exit a lung through the same channel. After exhalation, the air tubes are filled
with “used” air from the alveoli which, during the following inhalation, is pulled back into the
lungs. In fact, lung ventilation in humans is so inefficient that in normal breathing only
approximately one-sixth of the air in the lungs is replenished with each inspiration. Even after
forced expiration, 20% to 35% of the air remains in the lungs.

9.3.3. Breathing coordination

Breathing is normally involuntary and automatic but can come under voluntary control. Neurons
in the medulla of the brain regulate normal, quiet breathing. They spontaneously produce
rhythmical bursts that stimulate contraction of the diaphragm and external intercostal muscles.
However, respiration must adjust itself to changing requirements of the body for oxygen. Carbon
dioxide rather than oxygen has the greatest effect on respiratory rate. Even a small rise in carbon
dioxide level in the blood, has a powerful effect on respiratory activity. Actually, the stimulatory
effects of carbon dioxide are due in part to an increase in hydrogen ion concentration in
cerebrospinal fluid.

CO2 +H2O H2CO3 H+ + HCO3-

This reaction shows that carbon dioxide combines with water to form carbonic acid. Carbonic
acid then dissociates to release hydrogen ions, making the cerebrospinal fluid more acidic, and
stimulating respiratory receptors in the medulla of the brain. Both rate and depth of respiration
increase.
96

9.3.4. Gaseous exchange in lungs and body tissues

Air (the atmosphere) is a mixture of gases: about 71% nitrogen, 20.9% oxygen, in addition to
fractional percentages of other gases, such as carbon dioxide (0.03%). This leads to evaluate the
pressure of each component of gas mixture as partial pressure which is the pressure that one gas
in a mixture of gases would exert if it were the only gas present.

The composition of air changes as soon as it enters the respiratory tract. Inspired air becomes
saturated with water vapor as it travels through the air-filled passageways toward the alveoli.
When inspired air reaches the alveoli, it mixes with residual air remaining from the previous
respiratory cycle. Partial pressure of oxygen drops and that of carbon dioxide rises. Upon
expiration, air from the alveoli mixes with air in the dead space (volume of inhaled air that does
not take part in the gas exchange) to produce still a different mixture.

Because the partial pressure of oxygen in lung alveoli is greater (100 mm Hg) than it is in venous
blood of lung capillaries (40 mm Hg), oxygen diffuses into the lung capillaries. In a similar
manner carbon dioxide in blood of the lung capillaries has a higher concentration (46 mm Hg)
than has this same gas in lung alveoli (40 mm Hg), so carbon dioxide diffuses from the blood
into the alveoli.

In tissues respiratory gases also move along their concentration gradients. Partial pressure of
oxygen in the blood (100 mm Hg) is greater than in the tissues (0 to 30 mm Hg), and partial
pressure of carbon dioxide in tissues (45 to 68 mm Hg) is greater than that in blood (40 mm Hg).
In each case gases diffuse from a location of higher concentration to one of lower concentration.

9.4. Transport of respiratory gases

In some invertebrates respiratory gases are simply carried, dissolved in body fluids. However,
solubility of oxygen is so low in water that it is adequate only for animals with low rates of
metabolism. For example, only approximately 1% of a human’s oxygen requirement can be
transported in this way. Consequently in many invertebrates and in virtually all vertebrates,
nearly all oxygen and a significant amount of carbon dioxide are transported by special colored
proteins, or respiratory pigments, in the blood.
97

The most widespread respiratory pigment in the animal kingdom is hemoglobin, a red, iron-
containing protein. Each molecule of hemoglobin is 5% heme, an iron-containing compound
giving the red color to blood, and 95% globin, a colorless protein. The heme portion of
hemoglobin has a great affinity for oxygen; each gram of hemoglobin can carry a maximum of
approximately 1.3 ml of oxygen. Because there are approximately 15 g of hemoglobin in each
100 ml of blood, fully oxygenated blood contains approximately 20 ml of oxygen per 100 ml.
When the oxygen concentration is high, as it is in the capillaries of the lung alveoli, hemoglobin
loads up with oxygen; in tissues where the prevailing oxygen partial pressure is low, hemoglobin
releases its stored oxygen reserves.

The same blood that transports oxygen to the tissues from the lungs must carry carbon dioxide
back to the lungs on its return trip. However, unlike oxygen that is transported almost
exclusively in combination with hemoglobin, carbon dioxide is transported in three different
forms. A small fraction of the blood-borne carbon dioxide, only about 7%, is carried as gas
physically dissolved in the plasma. The remainder diffuses into red blood cells.

In red blood cells, most carbon dioxide, approximately 70%, becomes carbonic acid through
action of the enzyme carbonic anhydrase. Carbonic acid immediately dissociates into hydrogen
ion and bicarbonate ion. We can summarize the entire reaction as follows:
CO2 +H2O H2CO3 H+ + HCO3-

Another fraction of the carbon dioxide, approximately 23%, combines reversibly with
hemoglobin. Carbon dioxide does not combine with the heme group but with amino groups of
several amino acids to form a compound called carbaminohemoglobin.
98

Chapter 10. EXCRETION AND HOMEOSTASIS

10.1. Water and osmotic regulation

10.1.1. Salt and water balance in marine invertebrates

Most marine invertebrates are in osmotic equilibrium with their seawater environment. They
have body surfaces that are permeable to salts and water so that their body fluid concentration
rises or falls in conformity with changes in concentrations of seawater.

10.1.2. Hyperosmotic regulation in brackish-water crab

Hyperosmotic refers to a solution whose osmotic pressure is greater than that of a solution to
which it is compared; it contains a greater concentration of dissolved particles and gains water
through a selectively permeable membrane from a solution containing fewer particles. The
animal is a hyperosmotic regulator, meaning that it maintains its body fluids more concentrated
than the surrounding water. By regulating against excessive dilution, thus protecting the cells
from extreme changes, these crabs can live successfully in the physically unstable but
biologically rich coastal environment.

For the shore crab and other coastal invertebrates to achieve hyperosmotic regulation by
preventing body fluids to become diluted and unbalanced, the problem is solved by the kidneys,
which can excrete the excess water as dilute urine. The second problem is salt loss. Because the
animal is saltier than its environment, it cannot avoid loss of ions by outward diffusion across the
gills. Salt is also lost in urine. This problem is solved by special salt-secreting cells in the gills
that actively remove ions from dilute seawater and move them into the blood by active
transport, thus maintaining the internal osmotic concentration.

10.2. Salt and water balance in terrestrial animals

Terrestrial animals lose water by evaporation from respiratory and body surfaces, excretion in
urine, and elimination in the feces. They replace such losses by water in the food, drinking water
when available, and retaining metabolic water formed in cells by oxidation of foods, especially
carbohydrates. Certain insects are able to absorb water vapor directly from atmospheric air. In
some desert rodents, metabolic water gain may constitute most of the animals’ water intake.
99

Water balance in a human and a desert rodent


Particularly revealing is a comparison of
Human Kangaroo
water balance in human beings, non-
(%) Rat (%)
Gains Drinking 48 0 desert mammals that drink water, with

Free water in food 40 10 that of kangaroo rats, desert rodents that


may drink no water at all. Kangaroo rats
Metabolic water 12 90
acquire all their water from their food:
Losses Urine 60 25
90% is metabolic water derived from
Evaporation 34 70
(lungs and skin) oxidation of foods and 10% as free
Feces 6 5 moisture in food.

Terrestrial insects, reptiles, and birds have no convenient way to rid themselves of toxic
ammonia; instead, they convert it into uric acid, a nontoxic, almost insoluble compound. This
conversion enables them to excrete a semisolid urine with little water loss. Marine birds and
turtles have evolved an effective solution for excreting large loads of salt eaten with their food.
Located above each eye is a special salt gland capable of excreting a highly concentrated
solution of sodium chloride. In birds the salt solution runs out the nares. Marine lizards and
turtles shed their salt gland secretion as salty tears.

10.3. Vertebrate kidney structure and function

10.3.1. Vertebrate kidney structure

The vertebrate kidney plays a prominent role in maintaining homeostasis because it is the
principal organ that regulates the volume and composition of the internal fluid environment.
While described as an organ of excretion, the removal of metabolic wastes is incidental to its
regulatory function.

The basic functional unit is the nephron, and urine is formed by three well-defined physiological
processes: filtration, reabsorption, and secretion. The following discussion focuses mainly on
the mammalian kidney, which is the most completely understood regulatory organ.
100

The two human kidneys are small organs comprising less than 1% of the body weight. Yet they
receive a remarkable 20% to 25% of the total cardiac output, some 2000 liters of blood each day.
This vast blood flow is channeled to approximately 2 million nephrons, which make up the bulk
of the two kidneys.

Each nephron begins with an expanded chamber, the Bowman’s capsule (renal corpuscle),
containing a tuft of capillaries called the glomerulus. Blood pressure in the capillaries forces a
protein-free filtrate into a renal tubule, consisting of several segments that perform different
functions in the process of urine formation. The filtrate passes first into a proximal convoluted
tubule, then into a long, thin-walled loop of Henle, which drops deep into the inner portion of
the kidney (the medulla) before returning to the outer portion (the cortex) where it joins a distal
convoluted tubule. From this, the fluid empties into a collecting duct which drains into the renal
pelvis. Here the urine is collected before being carried by the ureter to the urinary bladder.

Urinary system of humans


Left: Kidney section showing regions
Right: Details of a single nephron

The arrows nephron diagram show the direction of filtrate and blood flow
101

The urine that leaves the collecting duct is very different from the filtrate produced in the renal
corpuscle. During its travels through the renal tubule and collecting duct, both the composition
and concentration of the original filtrate change. Some solutes such as glucose and sodium have
been reabsorbed while other materials, such as hydrogen ions and urea, have been concentrated
in the urine.

The nephron, with its pressure filter and tubule, is intimately associated with blood circulation.
Blood from the aorta enters each kidney through a large renal artery, which divides into a
branching system of smaller arteries. The arterial blood reaches the renal corpuscle through an
afferent arteriole and leaves by way of an efferent arteriole.

From the efferent arteriole the blood travels to an extensive capillary network that surrounds and
supplies the proximal and distal convoluted tubules and the loop of Henle. This capillary
network provides a means for the pickup and delivery of materials that are reabsorbed or
secreted by the kidney tubules. From these capillaries blood is collected by veins that unite to
form the renal vein. This vein returns the blood to the vena cava.

10.3.2. Glomerular filtration

The glomerulus is the place where the process of urine formation begins. It acts as a specialized
mechanical filter in which a protein-free filtrate of the plasma is driven by the blood pressure
across the capillary walls and into the fluid-filled space of the renal corpuscle.

Solute molecules small enough to pass through the pores of the capillary wall are carried through
with the water. RBCs and plasma proteins, however, are withheld because they are too large to
pass through these pores. The filtrate continues through the renal tubular system. Human kidneys
form each day a volume much exceeding the total blood volume. If such volume of water and the
valuable nutrients and salts it contains were lost, the body would soon be depleted of these
compounds. In fact, nearly all of the filtrate is reabsorbed.

The final urine volume in humans averages 1.2 liters per day. Conversion of filtrate into urine
involves two processes: (1) modification of the composition of the filtrate through tubular
reabsorption and secretion, and (2) changes in the total osmotic concentration of the urine
through the regulation of water excretion.
102

10.3.3. Tubular reabsorption

Approximately 60% of the filtrate volume and virtually all of the glucose, amino acids, vitamins
and other valuable nutrients are reabsorbed in the proximal convoluted tubule. Much of this
reabsorption is by active transport. Electrolytes such as sodium, potassium, calcium,
bicarbonate, and phosphate are reabsorbed by ion pumps, which are carrier proteins driven by the
hydrolysis of ATP. These are differently reabsorbed with specific pumps depending on the
body’s need to conserve each mineral.

Some materials are passively reabsorbed. Negatively charged chloride ions passively accompany
active reabsorption of positively charged sodium ions in the proximal convoluted tubule. Water,
too, is withdrawn passively from the tubule, as it follows osmotically the active reabsorption of
solutes. In descending limb of the loop of Henle numerous water channels formed by aquaporin
proteins make the transport epithelium freely permeable to water. In contrast, there is a near
absence of channels for salt and other small solutes, resulting in a very low permeability for
these substances. Unlike the descending limb, the ascending limb has a transport epithelium that
contains ion channels, but not water channels. Indeed, this membrane is impermeable to water.
When the kidneys are conserving water, aquaporin channels in the collecting duct also allow
water molecules to cross the epithelium.

For most substances there is an upper limit to the amount of substance that can be reabsorbed.
This upper limit is termed the transport maximum (renal threshold) for that substance. For
example, glucose normally is reabsorbed completely by the kidney because the transport
maximum for glucose is poised well above the amount of glucose usually present in the plasma
filtrate. Should the plasma glucose concentration exceed this threshold level, as in the disease
diabetes mellitus, glucose appears in the urine.

Unlike glucose, most electrolytes are excreted in the urine in variable amounts. Sodium is the
dominant cation in the plasma. The human kidney filters approximately 600 g of sodium every
day. Nearly all of this sodium is reabsorbed, but the exact amount is matched precisely to sodium
intake. With a very high salt intake, much above 20 g per day, the kidney cannot excrete sodium
as fast as it enters; the unexcreted sodium chloride holds additional water in the body fluids, and
the person begins to gain weight.
103

In the distal convoluted tubule, sodium reabsorption is controlled by aldosterone, a steroid


hormone from the adrenal gland. Aldosterone increases active reabsorption of sodium by the
distal tubules and thus decreases loss of sodium in the urine. The secretion of aldosterone is
regulated mainly by the enzyme renin, produced by the juxtaglomerular apparatus, a complex
of cells located in the afferent arteriole at its junction with the glomerulus.

Renin is released in response to a low blood sodium level or to low blood pressure. Renin then
initiates a series of enzymatic events culminating in the production of angiotensin, a blood
protein that has several related effects. First, it stimulates the release of aldosterone, which acts
in turn to increase sodium reabsorption by the distal tubule. Second, it increases the secretion of
antidiuretic hormone (vasopressin), which promotes water conservation by the kidney. Third, it
increases blood pressure. Finally, it stimulates thirst.

10.3.4. Tubular secretion

In addition to reabsorbing materials from plasma filtrate, the nephron can secrete materials
across the tubular epithelium and into the filtrate. In this process, the reverse of tubular
reabsorption, carrier proteins in the tubular epithelial cells selectively transport substances from
blood in capillaries outside the tubule to the filtrate inside the tubule. Tubular secretion enables
the kidney to build up the urine concentrations of materials to be excreted, such as hydrogen and
potassium ions, drugs, and various foreign organic materials. The distal convoluted tubule is the
site of most tubular secretion. The distal tubule plays a key role in regulating the K + and NaCl
concentration of body fluids. This regulation involves variation in the amount of the K + that is
secreted into the filtrate, as well as the amount of NaCl reabsorbed from the filtrate.

In the kidneys of bony marine fishes, reptiles, and birds, tubular secretion is a much more highly
developed process than it is in mammalian kidneys. Instead of urea, reptiles and birds excrete
uric acid which is actively secreted by the tubular epithelium. Since uric acid is nearly insoluble,
it forms crystals in the urine and requires little water for excretion. Thus excretion of uric acid is
an important adaptation for water conservation.
104

Movement of materials in the nephron and collecting duct

10.4. Physiological adaptations to


temperature

Biochemical reactions are extremely sensitive to temperature. All enzymes have an optimum
temperature; at temperatures above or below this optimum, enzyme function is impaired. When
body temperature drops too low, metabolic processes slow, reducing the amount of energy the
animal can muster for activity and reproduction.
105

If body temperature rises too high, metabolic reactions become unbalanced and enzymatic
activity is hampered or even destroyed. Thus animals can succeed only in a restricted range of
temperature, usually between 0° to 40° C. Animals must either find a habitat where they do not
have to contend with temperature extremes, or they must develop means of stabilizing their
metabolism independent of temperature extremes.

10.4.1. Ectothermy and endothermy

The term poikilothermic (variable body temperature) and homeothermic (constant body
temperature) are frequently used by zoologists as alternatives to “cold-blooded” and “warm-
blooded,” respectively. These terms, which refer to variability of body temperature, are more
precise and more informative, but still offer difficulties. Physiologists prefer yet another way to
describe body temperatures, one that reflects the fact that an animal’s body temperature is a
balance between heat gain and heat loss.

All animals produce heat from cellular metabolism, but in most the heat is conducted away as
fast as it is produced. In these animals, the ectotherms body temperature is determined solely by
the environment. Many ectotherms exploit their environment behaviorally to select areas of more
favorable temperature (such as basking in the sun). The source of energy used to increase body
temperature comes from the environment, not from within the body.

Some other groups of animals are able to generate and retain enough heat to elevate their own
body temperature to a high but stable level. Because the source of their body heat is internal,
they are called endotherms. This category includes birds and mammals, as well as a few reptiles
and fast-swimming fishes, and certain insects that are at least partially endothermic.

Endothermy allows birds and mammals to stabilize their internal temperature so that biochemical
processes and nervous system functions can proceed at steady high levels of activity.
Endotherms can thus remain active in winter and exploit habitats denied to ectotherms.
106

10.4.2. Temperature independence in ectotherms

Behavioral adjustments

Ectotherms often have the option of seeking areas in their environment where the temperature is
favorable to their activities. Behavioral patterns in desert lizards help to maintain a relatively
steady body temperature of 36° to 39° C while the air temperature varies between 29° and 44° C.

Metabolic adjustments

Most ectotherms can adjust their metabolic rates to the prevailing temperature such that the
intensity of metabolism remains mostly unchanged. This is called temperature compensation,
analogous to homeostasis, and involves complex biochemical and cellular adjustments that can
help such animals benefit from almost the same level of activity in both warm and cold
environments.

10.4.3. Temperature regulation in endotherms

Most mammals have body temperatures between 36° and 38° C, somewhat lower than those of
birds, which range between 40° and 42° C. Constant temperature is maintained by a delicate
balance between heat production and heat loss. Heat is produced by the animal’s metabolism;
this includes oxidation of foods, basal cellular metabolism, and muscular contraction.
Accordingly, endotherms need eat more food than ectotherms of the same size.

If such animals become too cool, they can generate heat by increasing muscular activity
(exercise or shivering) and by decreasing heat loss by increasing its insulation. If it becomes too
warm, it decreases heat production and increases heat loss.

Adaptations for hot environments

Despite the harsh conditions of deserts many kinds of animals live there successfully. The
smaller desert mammals are mostly fossorial (living mainly in the ground) or nocturnal (active
at night). The lower temperature and higher humidity of burrows help to reduce water loss by
evaporation.
107

Large desert ungulates such as camels and desert antelopes possess a number of adaptations for
coping with heat and dehydration. The glossy, pallid color of fur reflects direct sunlight, and fur
itself is an excellent insulation that resists heat. Fat tissue, an essential food reserve, is
concentrated in a single hump on the back, instead of being uniformly distributed under the skin
where it would impair loss of heat by radiation.

Elands avoid evaporative water loss by permitting their body temperature to drop during the cool
night and then to rise slowly during the day as the body stores heat. Only when the body
temperature reaches 41° C must elands prevent further rise through evaporative cooling by
sweating and panting. They conserve water by producing a concentrated urine and dry feces.
Camels have all of these adaptations developed to a similar or even greater degree; they are
perhaps the most perfectly adapted of all large desert mammals.

Adaptations for cold environments

In cold environments mammals and birds use two major mechanisms to maintain homeothermy:
(1) decreased conductance, reduction of heat loss by increasing the effectiveness of the
insulation, and (2) increased heat production.

In all mammals living in the cold regions of the earth, fur thickness increases in winter,
sometimes by as much as 50%. However, unlike the well-insulated trunk of the body, the body
extremities (legs, tail,ears, nose) of arctic mammals are thinly insulated and exposed to rapid
cooling. To prevent these parts from becoming major avenues of heat loss, they are allowed to
cool to very low temperatures. The heat in the warm arterial blood is not lost from the body
because of a countercurrent heat exchange between the outgoing warm blood and the returning
cold blood prevents heat loss.

A consequence of peripheral heat exchange is that legs and feet of mammals and birds living in
cold environments must function at low temperatures. Temperatures of the feet of arctic foxes
and barren-ground caribou are just above the freezing point. To keep feet supple and flexible at
such low temperatures, fats in the extremities have very low melting points. In severely cold
conditions all mammals can produce more heat by augmented muscular activity through
108

exercise or shivering. Another source of heat is increased oxidation of foods, especially from
stores of brown fat. This mechanism is called nonshivering thermogenesis.

10.4.4. Adaptive hypothermia

Endothermy is energetically expensive. Whereas an ectotherm can survive for weeks in a cold
environment without eating, an endotherm must always have energy resources to supply its high
metabolic rate. Consequently a few small birds and mammals have evolved ways to abandon
homeothermy for periods ranging from a few hours to several months, allowing their body
temperature to fall until it approaches or equals the temperature of surrounding air.

Some very small mammals, such as bats, maintain high body temperatures when active but allow
their body temperature to drop profoundly when inactive and asleep. This kind of adaptive
hypothermia is called daily torpor. Hummingbirds also may drop their body temperature at night
when food supplies are low.

Many small and medium-sized mammals in northern temperate regions solve the problem of
winter scarcity of food and low temperature by entering a prolonged and controlled state of
dormancy: hibernation. True hibernators, such as ground squirrels, jumping mice, marmots, and
woodchucks, prepare for hibernation by storing body fat. Subsequently their body metabolism
respiratory and heart rates decrease in a characteristic way. During arousal a hibernator both
shivers violently and employs nonshivering thermogenesis to produce heat.

Some mammals, such as bears, badgers, raccoons, and opossums, enter a state of prolonged sleep
in winter with little or no decrease in body temperature. Prolonged sleep is not true hibernation.
Bears of the northern forest sleep for several months. A bear’s heart rate may decrease from 40
to 10 beats per minute, but body temperature remains normal and the bear is awakened if
sufficiently disturbed.

You might also like