100% found this document useful (1 vote)
438 views

Modern Physics Module

This document discusses Einstein's theory of special relativity. It begins by outlining the key objectives to be covered, which include discussing Einstein's two postulates of relativity and how measurements can differ between observers. It then presents an example experiment involving observers on a moving train measuring the motion of a tossed ball. The document goes on to explain Einstein's postulates in more detail, including his principle of relativity stating that physical laws are the same in all inertial frames, and his postulate that the speed of light is constant in all frames. It also discusses how this challenges traditional notions of relative velocity and sets the stage for developing the mathematical transformations relating measurements between frames.

Uploaded by

Michael Esmalla
Copyright
© © All Rights Reserved
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
438 views

Modern Physics Module

This document discusses Einstein's theory of special relativity. It begins by outlining the key objectives to be covered, which include discussing Einstein's two postulates of relativity and how measurements can differ between observers. It then presents an example experiment involving observers on a moving train measuring the motion of a tossed ball. The document goes on to explain Einstein's postulates in more detail, including his principle of relativity stating that physical laws are the same in all inertial frames, and his postulate that the speed of light is constant in all frames. It also discusses how this challenges traditional notions of relative velocity and sets the stage for developing the mathematical transformations relating measurements between frames.

Uploaded by

Michael Esmalla
Copyright
© © All Rights Reserved
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 111

General Physiology

Modern Physics
Module 1: Relativity

The behavior of space and time is


fundamentally explained by special
relativity. It was created by Albert
Einstein and is based on the idea that all
observers who are not moving faster
than light are subject to the same
physical laws. It is crucial for
comprehending both the structure of the
universe and the behavior of fast-moving
objects, and it has numerous practical
applications.
By the end of the end of this
module you shall be able to:

 Discuss the two postulates of


Einstein’s special theory of relativity,
and what motivates these postulates.
 Investigate why different observers
can disagree about whether two
events are simultaneous.
 Provide experimental evidence that
confirms relativity predictions about
clocks running slow.
 Determine how the length of an object
changes due to the object’s motion.
 Demonstrate how the velocity of an
object depends on the frame of
reference from which it is observed.
 Discuss how the theory of relativity
modifies the relationship between
velocity and momentum.
 Solve problems involving work and
kinetic energy for particles moving at
relativistic speeds.
General Physiology

Lesson 1
Theory of Special Relativity

Objectives: At the end of the lesson the students shall be able to:
 Discuss the two postulates of Einstein’s special theory of relativity,
and what motivates these postulates.
 Investigate why different observers can disagree about whether two
events are simultaneous.
 Provide experimental evidence that confirms relativity predictions about
clocks running slow.
 Determine how the length of an object changes due to the object’s
motion.
 Demonstrate how the velocity of an object depends on the frame of
reference from which it is observed.

Activity:
Ben and Dan carried out an experiment. They toss a ball
inside a moving train and then watch the ball move upward and
downward. Ben and Dan decided to go out the train and do the
same thing on the ground with the same force they used on the
ball.
 Will there be any differences between the ball inside the train
and the ball outside the ground?
 What do you think Ben and Dan will notice about the ball
inside and outside the train?
General Physiology

Ben and Dan then decided to repeat the experiment, but this time
Ben is inside the moving train and Dan is on the ground.
 When observing the ball at the same time, is there any difference
between Ben's and Dan's observations?
 How do Ben and Dan perceive the ball differently?

Analysis: From the activity, how does the movement of the train impact
the trajectory of the ball, and how does the observation point of
the experiment affect the perception of the ball's movement?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________
How does the concept of relative motion apply to Ben and Dan's
experiment, and what implications does it have for the results
they observe inside and outside the train?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________

Abstraction:
The Principle of Relativity
Einstein’s First Postulate
Einstein’s first postulate, called the principle of relativity,
states: The laws of physics are the same in every inertial frame of
reference. If the laws differed, that difference
could distinguish one inertial frame from the
others or make one frame somehow more
“correct” than another. Here are two examples.
Suppose you watch two children playing catch
with a ball while the three of you are aboard a
train moving with constant velocity. Your
observations of the motion of the ball, no
matter how carefully done, can’t tell you how
fast (or whether) the train is moving. This is
because Newton’s laws of motion are the same
in every inertial frame. Another example is the
electromotive force (emf) induced in a coil of
wire by a nearby moving permanent magnet. In
the frame of reference in which the coil is
General Physiology

stationary, the moving magnet causes a change of magnetic flux through the
coil, and this induces an emf. In a different frame of reference in which the
magnet is stationary, the motion of the coil through a magnetic field induces the
emf. According to the principle of relativity, both frames of reference are equally
valid. Hence the same emf must be induced in both situations. This is indeed
the case, so Faraday’s law is consistent with the principle of relativity.

Indeed, all the laws of electromagnetism are the same in every inertial
frame of reference. Equally significant is the prediction of the speed of
electromagnetic radiation, derived from Maxwell’s equations. According to this
analysis, light and all other electromagnetic waves travel in vacuum with a
constant speed, now defined to equal exactly 299,792,458 m/s (We often use
the approximate value of c = 3.00x108 m/s which is within one part in 1000 of
the exact value.)

As we will see, the speed of light in vacuum plays a central role in the
theory of relativity.

Einstein’s Second Postulate

During the 19th century, most physicists believed that light traveled
through a hypothetical medium called the ether, just as sound waves travel
through air. If so, the speed of light measured by observers would depend on
their motion relative to the ether and would therefore be different in different
directions. The Michelson-Morley experiment was an effort to detect motion of
the earth relative to the ether. Einstein’s conceptual leap was to recognize that if
Maxwell’s equations are valid in all inertial frames, then the speed of light in
vacuum should also be the same in all frames and in all directions. In fact,
Michelson and Morley detected no ether motion across the earth, and the ether
concept has been discarded. Although Einstein may not have known about this
negative result, it supported his bold hypothesis of the constancy of the speed of
light in vacuum.
Einstein’s second postulate states: The speed of light in vacuum is the
same in all inertial frames of reference and is independent of the motion of
the source.
General Physiology

Let’s think about what this means. Suppose two observers measure
the speed of light in vacuum. One is at rest with respect to the light source,
and the other is moving away from it. Both are in inertial frames of
reference. According to the principle of relativity, the two observers must
obtain the same result, despite the fact that one is moving with respect to
the other.

If this seems too easy, consider the following situation. A spacecraft


moving past the earth at 1000 m/s fires a missile straight ahead with a
speed of 2000 m/s (relative to the spacecraft). What is the missile’s speed
relative to the earth? Simple, you say; this is an elementary problem in
relative velocity. The correct answer, according to Newtonian mechanics, is
3000 m/s.

But now suppose the spacecraft turns on a searchlight, pointing in


the same direction in which the missile was fired. An observer on the
spacecraft measures the speed of light emitted by the searchlight and
obtains the value According to Einstein’s second postulate, the motion of
the light after it has left the source cannot depend on the motion of the
source. So, the observer on earth who measures the speed of this same
light must also obtain the value not This result contradicts our elementary
notion of relative velocities, and it may not appear to agree with common
sense. But “common sense” is intuition based on everyday experience, and
this does not usually include measurements of the speed of light.

Galilean Coordinate Transformation

In order to show the underlying equivalence of measurements made in


different reference frames and hence the equivalence of different frames for
doing physics, we need a mathematical formula that systematically relates
measurements made in one reference frame to those in another. Such a
relation is called a transformation, and the one satisfying Newtonian
relativity is the so called Galilean transformation, which owes its origin to
Galileo. It can be derived as follows.
General Physiology

Consider two
inertial systems or
frames S and S’. The
frame S’ moves with a constant velocity v along the xx’ axes, where v is
measured relative to the frame S. Clocks in S and S’ are synchronized, and
the origins of S and S’ coincide at t=t’=0. We assume that a point event, a
physical phenomenon such as a lightbulb flash, occurs at the point P. An
observer in the system S would describe the event with space–time
coordinates (x, y, z,t), whereas an observer in S’ would use (x’, y’, z’, t’) to
describe the same event. As we can see from, these coordinates are related
by the equations:

These equations constitute what is known as a Galilean


transformation of coordinates. Note that the fourth coordinate, time, is
assumed to be the same in both inertial frames. That is, in classical
mechanics, all clocks run at the same rate regardless of their velocity, so
that the time at which an event occurs for an observer in S is the same as
the time for the same event in S’. Consequently, the time interval between
two successive events should be the same for both observers. Although this
assumption may seem obvious, it turns out to be incorrect when treating
General Physiology

situations in which v is comparable to the speed of light. In fact, this point


represents one of the most profound differences between Newtonian
concepts and the ideas contained in Einstein’s theory of relativity.

An immediate and important consequence of the invariance of the


distance between two points under the Galilean transformation is the
invariance of force. For example, if gives the electric force
between two charges q,Q located at x1 and x2 on the x-axis in frame S, F’, the
force measured in S’, is given by.
In fact any force would be invariant under the Galilean transformation as long
as it involved only the relative positions of interacting particles.

Now suppose two events are separated by a distance dx and a time


interval dt as measured by an observer in S. It follows the equation that the
corresponding displacement dx’ measured by an observer in S’ is given by

dx’=dx-v dt, where dx is the displacement measured by an observer in S.


Because dt = dt’, we find that

or

where ux and ux’ are the instantaneous velocities of the object relative to S
and S’, respectively. This result, which is called the Galilean addition law for
velocities (or Galilean velocity transformation), is used in everyday
observations and is consistent with our intuitive notions of time and space.
To obtain the relation between the accelerations measured by observers in S
and S’, we take a derivative of with respect to time and
use the results that dt = dt’ and v is constant:

Thus, observers in different inertial frames measure the same acceleration for

an accelerating object. The mathematical terminology is to say that lengths


(Δx), time intervals, and accelerations are invariant under a Galilean
transformation. It points up the distinction between invariant and covariant
and shows that transformation equations, in addition to converting
General Physiology

measurements made in one inertial frame to those in another, may be used to


show the covariance of physical laws.

Relativity of Simultaneity

Measuring times and time intervals involves the concept of simultaneity. In


a given frame of reference, an event is an occurrence that has a definite position
and time. When you say that you awoke at seven o’clock, you mean that two
events (you’re awakening and your clock showing 7:00) occurred simultaneously.
The fundamental problem in measuring time intervals is this: In general, two
events that are simultaneous in one frame of reference are not simultaneous in a
second frame that is moving relative to the first, even if both are inertial frames.

This may seem to be contrary to common sense. To illustrate the point,


here is a version of one of Einstein’s thought experiments—mental experiments
that follow concepts to their logical conclusions.

Imagine a train moving with a speed comparable to with uniform velocity.


Two lightning bolts strike a passenger car, one near each end. Each bolt leaves a
mark on the car and one on the ground at the instant the bolt hits. The points on
the ground are labeled A and B in the figure, and the corresponding points on the
car are A' and B’. Stanley is stationary on the ground at O, midway between A
and B. Mavis is moving with the train at O’ in the middle of the passenger car,
midway between A’ and B’. Both Stanley and Mavis see both light flashes emitted
from the points where the lightning strikes.

Suppose the two wave fronts from the lightning strikes reach Stanley at O
simultaneously. He knows that he is the same distance from B and A so Stanley
concludes that the two bolts struck and simultaneously. Mavis agrees that the
two wave fronts reached Stanley at the same time, but she disagrees that the

flashes were emitted simultaneously.

Stanley and Mavis agree that the two wave fronts do not reach Mavis at the
same time. Mavis at O’ is moving to the right with the train, so she runs into the
wave front from B’ before the wave front from A’ catches up to her. However,
because she is in the middle of the passenger car equidistant from A’ and B’ her
observation is that both wave fronts took the same time to reach her because
both moved the same distance at the same speed c. Thus she concludes that the
lightning bolt B’ at struck before the one at A’. Stanley at O measures the two
General Physiology

events to be simultaneous, but Mavis at O’ does not! Whether or not two events at
different x-axis locations are simultaneous depends on the state of motion of the
observer.

You may want to argue that in this example the lightning bolts really are
simultaneous and that if Mavis at O’ could communicate with the distant points
without the time delay caused by the finite speed of light, she would realize this.
But that would be erroneous; the finite speed of information transmission is not
the real issue. If O’ is midway between A’ and B’ then in her frame of reference the
time for a signal to travel from A’ to O’ is the same as that from B’ to O’. Two
signals arrive simultaneously at O’ only if they were emitted simultaneously at A’
and B’. In this example they do not arrive simultaneously at O’ and so Mavis
must conclude that the events at A’ and B’ were not simultaneous.
General Physiology

Furthermore, there is no basis for saying that Stanley is right, and Mavis is
wrong, or vice versa. According to the principle of relativity, no inertial frame of
reference is more correct than any other in the formulation of physical laws. Each
observer is correct in his or her own frame of reference. In other words,
simultaneity is not an absolute concept. Whether two events are simultaneous
depends on the frame of reference. As we mentioned at the beginning of this
section, simultaneity plays an essential role in measuring time intervals. It follows
that the time interval between two events may be different in different frames of
reference. So, our next task is to learn how to compare time intervals in different
frames of reference.

Time Dilation
The fact that observers in different inertial frames always measure
different time intervals between a pair of events can be illustrated in another
way by considering a vehicle moving to the right with a speed v. A mirror is fixed
to the ceiling of the vehicle, and observer O’, at rest in this system, holds a laser
a distance d below the mirror. At some instant the laser emits a pulse of light
directed toward the mirror, and at some later time, after reflecting from the
mirror, the pulse arrives back at the laser (event 2). Observer O’ carries a clock,
C’, which she uses to measure the time interval Δt’ between these two events.
Because the light pulse has the speed c, the time it takes to travel from O’ to the
mirror and back can be found from the definition of speed:
General Physiology

Now consider the same set of events as viewed by observer O in a second frame.
According to this observer, the mirror and laser are moving to the right with a
speed v, and as a result, the sequence of events appears different to this
observer. By the time the light from the laser reaches the mirror, the mirror has
moved to the right a distance vΔt/2, where Δt is the time interval required for
the light pulse to travel from O’ to the mirror and back as measured by O. In
other words, O concludes that, because of the motion of the vehicle, if the light
is to hit the mirror, it must leave the laser at an angle with respect to the
vertical direction. Comparing the two scenarios, we see that the light must
travel farther in (b) than in (a). (Note that neither observer “knows” that he or
she is moving. Each is at rest in his or her own inertial frame.) According to the
second postulate of special relativity, both observers must measure c for the
speed of light. Because the light travels farther according to O, it follows that the
time interval Δt measured by O is longer than the time interval Δt’ measured by
O’. To obtain a relationship between Δt and Δt’, it is convenient to use the right
triangle equation. The Pythagorean theorem gives:

The Twin Paradox

Consider identical twin astronauts named Eartha and Astrid. Eartha


General Physiology

remains on earth while her twin Astrid takes off on a high-speed trip through the
galaxy. Because of time dilation, Eartha observes Astrid’s heartbeat and all other
life processes proceeding more slowly than her own. Thus to Eartha, Astrid ages
more slowly; when Astrid returns to earth she is younger (has aged less) than
Eartha. Now here is the paradox: All inertial frames are equivalent. Can’t Astrid
make the same arguments to conclude that Eartha is in fact the younger? Then
each twin measures the other to be younger when they’re back together, and
that’s a paradox. To resolve the paradox, we recognize that the twins are not
identical in all respects. While Eartha always remains in an approximately inertial
frame, Astrid must accelerate with respect to that inertial frame during parts of
her trip in order to leave, turn around, and return to earth. Eartha’s reference
frame is always approximately inertial; Astrid’s is often far from inertial. Thus,
there is a real physical difference between the circumstances of the two twins.
Careful analysis shows that Eartha is correct; when Astrid returns, she is
younger than Eartha.

Application:
Read the case below and answer the following questions.
“The Case of the Choking Curse”
A young woman, Sarah, comes to the hospital with
symptoms of choking and difficulty breathing. She is convinced
that she has been cursed by an evil spirit and that her
symptoms are supernatural in nature.
The medical team performs a thorough evaluation and
discovers that Sarah has a condition called globus pharyngeus,
which is a sensation of a lump in the throat that is not caused
by any physical obstruction. This condition is often triggered by
stress and anxiety and is not related to any supernatural
forces.
The medical team explains to Sarah that her symptoms
are caused by physiological changes in her body and that she
does not have a curse or evil spirit. They give her a treatment
plan that includes reducing stress and anxiety, practicing
relaxation techniques, and seeking counseling if needed.
1. Why do you think Sarah believed she was cursed by an evil
spirit?
2. How can medical professionals help patients who believe they
have a supernatural medical condition?
3. Why is it important to separate superstitions from medical
conditions?
General Physiology

Assessment:

Read the following questions and choose the choices that


corresponds to the best answer.
1. What is Physiology and how is it different from anatomy?
a. Physiology is the study of the structures of the body while anatomy focuses
on the functioning of the body.
b. Physiology is the study of the functioning of the body while anatomy focuses
on the structures of the body.
c. Physiology is the study of the evolution of species while anatomy focuses on
the functioning of the body.
d. Physiology and anatomy are the same.

2. How does the study of Physiology contribute to our understanding of the


human body?
a. By observing the structures of the body.
b. By observing the interactions between the various parts of the body.
c. By integrating knowledge from different scientific fields.
d. All of the above

3. How does physics contribute to the study of Physiology?


a. By understanding electrical conduction and fluid dynamics.
b. By understanding the structure of large biological molecules.
c. By understanding the ecology of organisms.
General Physiology

d. By understanding the chemical reactions in living processes.

4. What is the impact of genome studies on human understanding of evolution?


a. It reveals the complex and often illogical nature of evolution at its most
fundamental level.
b. It helps to deepen our understanding of evolution and the role of genetics in
shaping species.
c. It has no impact on evolution.
d. It helps to simplify our understanding of evolution and the role of genetics in
shaping species.

5. What is the hypothetico-deductive method and why is it commonly used in


scientific investigation?
A. A method used in social sciences to understand human behavior.
B. A process used in engineering to design products.
C. A logical approach that involves creating and testing hypotheses to explain
natural phenomena.
D. A mathematical formula used to calculate the probability of an event.

6. How does the scientific method help us understand the world?


a. By using intuition and personal experience.
b. By using religion and faith-based beliefs.
c. By using a systematic and logical approach to uncover consistent patterns
and processes in nature.
d. By using random trial and error.

7. How does the integrative nature of physiology present challenges for


researchers?
a. Physiology is not challenging and does not require a lot of effort.
b. Physiology is limited to only molecular research.
c. Physiology requires a comprehensive understanding of processes from the
functions of specific molecules to the behaviors of animals.
d. Physiology only involves the study of plants.

8. When can a hypothesis be elevated to a scientific theory?


a. When it is supported by one experiment.
b. When it is supported by intuition and personal experience.
c. When it is consistently supported by multiple tests and all alternative
hypotheses have been falsified.
d. When it is supported by a majority vote.

9. How does the discovery phase of science involve asking questions about
nature?
General Physiology

a. Scientists gather all available information, existing hypotheses, and explore


the unknown through observation and experimentation.
b. Scientists create hypotheses to explain the phenomenon without exploring or
gathering information.
c. Scientists only observe and do not gather information or ask questions.
d. Scientists do not ask questions about nature.

10. How are experiments or observations conducted in the hypothetico-


deductive method?
a. by relying on intuition and personal experience.
b. by using random trial and error.
c. by perturbing an organism or natural process and measuring the outcome
with controls.
d. by using methods that are not testable or falsifiable.

Lesson 2
The Functions and Requirement of Human Life
Objectives:
At the end of the lesson the students shall be able to:
 Distinguish between metabolism, anabolism, and catabolism;
 Compare and contrast growth, differentiation, and reproduction;
 Describe the structure of the human body in terms of six levels
of organization; and
 List the eleven organ systems of the human body and identify at
least one organ and one major function of each.

Activity:

Picture Analysis
General Physiology

From the activity, How does the concept of levels of organization


help to understand the relationships between different biological
systems and structures, from the molecular to the organismal
level?
_______________________________________________________________
Analysis:
_______________________________________________________________
_______________________________________________________________
The cow’s digestive system
In what and
ways the an
does worm’s digestive system
understanding both
of levels breakdown
of organization
food, but they doinform
not look the same. Compare and contrast the two
our understanding of the relationships between systems.
What do you think is the and
organisms similarities and the differences
their environments, from thebetween the
cellular to two
the
digestive
ecosystem level? system?

_______________________________________________________________
_______________________________________________________________
_______________________________________________________________

Abstraction:
Levels of Organization
Before you begin to study the different functions of the human
body, it is helpful to consider its basic architecture; that is, how
its smallest parts are assembled into larger structures. It is
convenient to consider the structures of the body in terms of
fundamental levels of organization that increase in complexity: subatomic
particles, atoms, molecules, organelles, cells, tissues, organs, organ systems,
organisms, and biosphere.
General Physiology

The Chemical Level


All matter in the universe is composed of one
or more unique pure substances called
elements, familiar examples of which are
hydrogen, oxygen, carbon, nitrogen, calcium,
and iron. To study the chemical level of
organization, scientists consider the simplest
building blocks of matter: subatomic
particles, atoms, and molecules. The smallest
unit of any of these pure substances
(elements) is an atom. Atoms are made up of
subatomic particles such as the proton,
electron, and neutron. Two or more atoms
combine to form a molecule, such as the
water molecules, proteins, and sugars found Figure 1.3 Levels of Structural Organization of the Human
in living things. Molecules are the chemical Body The organization of the body often is discussed in
terms of six distinct levels of increasing complexity, from
building blocks of all body structures. the smallest chemical building blocks to a unique human
organism.

The Cell Level

The Cell Theory of Life: A Scientific Breakthrough in the 19th Century

The study of life sciences in the 19th century led to the development of
the cell theory of life, which states that the cell is the fundamental unit of life.
Despite the challenge of defining "life," cells are essential for all living organisms
and play a crucial role in carrying out life processes on Earth. A cell consists of
a membrane that acts as an outer barrier and a fluid-filled interior called the
cytosol. The cytosol contains water, salts, small organic molecules, and
macromolecules, including DNA and RNA, which carry genetic information, and
proteins, which are complex shaped molecules that carry out cellular functions
and reactions.

Cells can be classified into two categories: prokaryotic cells, found in the
groups Bacteria and Archaea, which lack complex internal structures such as
nuclei, and eukaryotic cells, which have these structures and are found in the
kingdoms Protista, Fungi, Plantae, and Animalia. However, due to the
fundamental differences between the two prokaryotic groups, biologists now
categorize all life into three domains: archaea, bacteria, and eukarya.

The simplest forms of life are either unicellular or colonial, typically


prokaryotes or certain eukaryotic protists. Meanwhile, more complex organisms
such as animals are multicellular, made up of hundreds to trillions of
cooperating eukaryotic cells with specific structures and functions.
Multicellularity was necessary for large body size due to physical limitations,
such as diffusion, on the size of a single cell.
General Physiology

All cells, whether they are single or part of a multicellular organism,


perform fundamental functions crucial for survival, including cell metabolism,
communication, and reproduction. These basic functions are essential for the
survival of the cell and ultimately the organism.

The Tissue Level

The cells must be specifically organized to


perform the life-sustaining processes of the
organism, such as digestion, respiration, and
circulation. There are four main types of tissues
in animals, each composed of similar cells with
similar structures and functions: epithelial,
connective, muscular, and nervous. The cells
in each tissue are specialized and accompanied
by varying amounts of extracellular material.

1. EPITHELIAL TISSUE is made up of specialized cells that facilitate the


exchange of materials between the organism and its environment. There are two
main structures of epithelial tissue: sheets and secretory glands.

 The epithelial cells are tightly joined to form sheets that cover and line
various organs, such as the skin and digestive tract. These sheets serve
as boundaries, separating the animal from its surroundings and the
contents of its internal cavities. Only certain materials are allowed to
cross these barriers, depending on the location and function of the tissue.
For example, the epithelial cells in the digestive tract are specialized for
absorbing water and nutrients, while the skin only allows limited
exchange with the environment.
 Secretory glands are derived from epithelial tissue and are specialized for
secretion, which is the release of specific products from cells in response
to stimulation. Glands are formed during embryonic development and can
be classified as either exocrine or endocrine. Exocrine glands secrete
through ducts to the outside of the organism, while endocrine glands
release their secretions into the bloodstream.

2. CONNECTIVE TISSUE is characterized by having relatively few cells dispersed


within a large amount of extracellular material that they secrete. It connects,
supports, and anchors different parts of the body, including loose connective
tissue, tendons, bones, and blood. The cells in connective tissue produce
specific molecules that they release into the extracellular spaces between cells,
such as the elastic protein elastin, which allows structures to stretch and recoil.

3. MUSCULAR TISSUE is composed of cells that are specialized for contraction


and force generation. There are three types of muscle tissue in vertebrates:
skeletal muscle, which causes movement of the skeleton, cardiac muscle,
which pumps blood out of the heart, and smooth muscle, which controls the
movement of contents through hollow organs.
General Physiology

4. NERVOUS TISSUE consists of cells that initiate and transmit electrical


impulses, which serve as signals to relay information between parts of the
organism. Nervous tissue can be found in the brain, spinal cord, and monitoring
linings, as well as muscles, glands, and effector organs.

These primary tissues are the building blocks of organs and are referred to as
classical tissues. The term tissue is also used to describe the aggregate of
cellular and extracellular components that make up a specific organ, such as
lung tissue or liver tissue.

The Organ Level


Organs are composed of two or more types of primary tissue that work
together to fulfill a specific function or functions. An example of such an organ
is the stomach, which is made up of four primary tissue types, and functions to
store ingested food, start the digestion process, and move food further into the
digestive tract. The walls of the vertebrate stomach are lined with epithelial
tissue, which prevents harsh digestive chemicals and undigested food from
entering the bloodstream. The gland cells in the stomach, derived from the
epithelial tissue, include exocrine cells that secrete digestive juices into the
lumen and endocrine cells that release hormones that regulate the stomach's
exocrine secretion and muscle contraction. The walls of the stomach are also
composed of smooth muscle tissue, which contracts to mix food with digestive
juices and propel the mixture into the intestine, and nervous tissue, which
controls muscle contraction and gland secretion with hormones. All these
tissues are held together by connective tissue.

The Organ System Level


Organs are further grouped into organ systems, which are collections of
organs that perform related functions and interact to achieve a common task
that is crucial for the survival of the entire body. For instance, the digestive
system, in vertebrates, comprises of the mouth, salivary glands, pharynx,
esophagus, stomach, pancreas, liver, gallbladder, small intestine, and large
intestine (and sometimes additional organs). These organs collectively break
down food into small absorbable nutrient molecules. The chapter concludes
with a more in-depth examination of these systems.

The Organismal Level


The entire animal body, which is a single living entity, is composed of
various organ systems that are structurally and functionally linked together to
form a separate entity from the external environment. An animal is made up of
living cells organized into life-sustaining systems.

Functions of Human Life

The different organ systems each have different functions and therefore
unique roles to perform in physiology. These many functions can be
summarized in terms of a few that we might consider definitive of human life:
General Physiology

organization, metabolism, responsiveness, movement, development, and


reproduction.

Organization

A human body consists of trillions of cells organized in a way that


maintains distinct internal compartments. These compartments keep body cells
separated from external environmental threats and keep the cells moist and
nourished. They also separate internal body fluids from the countless
microorganisms that grow on body surfaces, including the lining of certain
passageways that connect to the outer surface of the body. The intestinal tract,
for example, is home to more bacterial cells than the total of all human cells in
the body, yet these bacteria are outside the body and cannot be allowed to
circulate freely inside the body.

Cells, for example, have a cell membrane (also referred to as the plasma
membrane) that keeps the intracellular environment—the fluids and organelles
—separate from the extracellular environment. Blood vessels keep blood inside a
closed circulatory system, and nerves and muscles are wrapped in connective
tissue sheaths that separate them from surrounding structures. In the chest
and abdomen, a variety of internal membranes keep major organs such as the
lungs, heart, and kidneys separate from others.

The body’s largest organ system is the integumentary system, which


includes the skin and its associated structures, such as hair and nails. The
surface tissue of skin is a barrier that protects internal structures and fluids
from potentially harmful microorganisms and other toxins.

Metabolism

The first law of thermodynamics


holds that energy can neither be created
nor destroyed—it can only change form.
Your basic function as an organism is to
consume (ingest) energy and molecules in
the foods you eat, convert some of it into
fuel for movement, sustain your body
functions, and build and maintain your
body structures. There are two types of
reactions that accomplish this: anabolism
and catabolism.

 Anabolism is the process whereby


smaller, simpler molecules are Anabolic reactions are building reactions,
combined into larger, more complex and they consume energy. Catabolic
substances. Your body can assemble, reactions break materials down and release
by utilizing energy, the complex energy. Metabolism includes both anabolic
and catabolic reactions.
General Physiology

chemicals it needs by combining small molecules derived from the foods


you eat.
 Catabolism is the process by which larger more complex substances are
broken down into smaller simpler molecules. Catabolism releases energy.
The complex molecules found in foods are broken down so the body can
use their parts to assemble the structures and substances needed for life.

Taken together, these two processes are called metabolism. Metabolism is


the sum of all anabolic and catabolic reactions that take place in the body. Both
anabolism and catabolism occur simultaneously and continuously to keep you
alive.

Every cell in your body makes use of a chemical compound, adenosine


triphosphate (ATP), to store and release energy. The cell stores energy in the
synthesis (anabolism) of ATP, then moves the ATP molecules to the location
where energy is needed to fuel cellular activities. Then the ATP is broken down
(catabolism) and a controlled amount of energy is released, which is used by the
cell to perform a particular job.

Responsiveness

Responsiveness is the ability of an


organism to adjust to changes in its internal
and external environments. An example of
responsiveness to external stimuli could
include moving toward sources of food and
water and away from perceived dangers.
Changes in an organism’s internal
environment, such as increased body
temperature, can cause the responses of
sweating and the dilation of blood vessels in
the skin to decrease body temperature.
Runners demonstrate two characteristics of
living humans—responsiveness and
Movement movement. Anatomic structures and
physiological processes allow runners to
Human movement includes not only coordinate the action of muscle groups and
actions at the joints of the body, but also the sweat in response to rising internal body
motion of individual organs and even temperature. (credit: Phil Roeder/flickr)
individual cells. As you read these words, red
General Physiology

and white blood cells are moving throughout your body, muscle cells are
contracting and relaxing to maintain your posture and to focus your vision, and
glands are secreting chemicals to regulate body functions. Your body is
coordinating the action of entire muscle groups to enable you to move air into
and out of your lungs, to push blood throughout your body, and to propel the
food you have eaten through your digestive tract. Consciously, of course, you
contract your skeletal muscles to move the bones of your skeleton to get from
one place to another, and to carry out all of the activities of your daily life.

Development, growth and reproduction

Development is all the changes the body goes through in life. Development
includes the process of differentiation, in which unspecialized cells become
specialized in structure and function to perform certain tasks in the body.
Development also includes the processes of growth and repair, both of which
involve cell differentiation.

Growth is the increase in body size. Humans, like all multicellular organisms,
grow by increasing the number of existing cells, increasing the amount of non-
cellular material around cells (such as mineral deposits in bone), and, within
very narrow limits, increasing the size of existing cells.

Reproduction is the formation of a new organism from parent organisms. In


humans, reproduction is carried out by the male and female reproductive
systems. Because death will come to all complex organisms, without
reproduction, the line of organisms would end.

Requirements of Human Life

Humans have been acclimating to life on Earth for at least the past
200,000 years. Earth and its atmosphere have provided us with air to breathe,
water to drink, and food to eat, but these are not the only requirements for
survival. Although you may rarely think about it, you also cannot live outside of
a certain range of temperature and pressure that the surface of our planet and
its atmosphere provides. The next sections explore these four requirements of
life.

Oxygen
Atmospheric air is only about 20 percent oxygen, but that oxygen is a key
component of the chemical reactions that keep the body alive, including the
reactions that produce ATP. Brain cells are especially sensitive to lack of oxygen
because of their requirement for a high-and-steady production of ATP. Brain
damage is likely within five minutes without oxygen, and death is likely within
ten minutes.

Nutrients
General Physiology

A nutrient is a substance in foods and beverages that is essential to


human survival. The three basic classes of nutrients are water, the energy-
yielding and body-building nutrients, and the micronutrients (vitamins and
minerals).

The most critical nutrient is water. Depending on the environmental


temperature and our state of health, we may be able to survive for only a few
days without water. The body’s functional chemicals are dissolved and
transported in water, and the chemical reactions of life take place in water.
Moreover, water is the largest component of cells, blood, and the fluid between
cells, and water makes up about 70 percent of an adult’s body mass. Water also
helps regulate our internal temperature and cushions, protects, and lubricates
joints and many other body structures.

The energy-yielding nutrients are primarily carbohydrates and lipids,


while proteins mainly supply the amino acids that are the building blocks of the
body itself. You ingest these in plant and animal foods and beverages, and the
digestive system breaks them down into molecules small enough to be
absorbed. The breakdown products of carbohydrates and lipids can then be
used in the metabolic processes that convert them to ATP. Although you might
feel as if you are starving after missing a single meal, you can survive without
consuming the energy-yielding nutrients for at least several weeks.

Water and the energy-yielding nutrients are also referred to as


macronutrients because the body needs them in large amounts. In contrast,
micronutrients are vitamins and minerals. These elements and compounds
participate in many essential chemical reactions and processes, such as nerve
impulses, and some, such as calcium, also contribute to the body’s structure.
Your body can store some of the micronutrients in its tissues, and draw on
those reserves if you fail to consume them in your diet for a few days or weeks.
Some others micronutrients, such as vitamin C and most of the B vitamins, are
water-soluble and cannot be stored, so you need to consume them every day or
two.

Narrow Range of Temperature


You have probably seen news stories about athletes who died of heat
stroke, or hikers who died of exposure to cold. Such deaths occur because the
chemical reactions upon which the body depends can only take place within a
narrow range of body temperature, from just below to just above 37°C (98.6°F).
When body temperature rises well above or drops well below normal, certain
proteins (enzymes) that facilitate chemical reactions lose their normal structure
and their ability to function and the chemical reactions of metabolism cannot
proceed.

That said, the body can respond effectively to short-term exposure to heat
or cold. One of the body’s responses to heat is, of course, sweating. As sweat
evaporates from skin, it removes some thermal energy from the body, cooling it.
Adequate water (from the extracellular fluid in the body) is necessary to produce
sweat, so adequate fluid intake is essential to balance that loss during the sweat
General Physiology

response. Not surprisingly, the sweat response is much less effective in a humid
environment because the air is already saturated with water. Thus, the sweat on
the skin’s surface is not able to evaporate, and internal body temperature can
get dangerously high.

The body can also respond effectively to short-term exposure to cold. One
response to cold is shivering, which is random muscle movement that generates
heat. Another response is increased breakdown of stored energy to generate
heat. When that energy reserve is depleted, however, and the core temperature
begins to drop significantly, red blood cells will lose their ability to give up
oxygen, denying the brain of this critical component of ATP production. This
lack of oxygen can cause confusion, lethargy, and eventually loss of
consciousness and death. The body responds to cold by reducing blood
circulation to the extremities, the hands, and feet, in order to prevent blood
from cooling there and so that the body’s core can stay warm. Even when core
body temperature remains stable, however, tissues exposed to severe cold,
especially the fingers and toes, can develop frostbite when blood flow to the
extremities has been much reduced. This form of tissue damage can be
permanent and lead to gangrene, requiring amputation of the affected region.

Narrow Range of Atmospheric Pressure


Pressure is a force exerted by a substance that is in contact with another
substance. Atmospheric pressure is pressure exerted by the mixture of gases
(primarily nitrogen and oxygen) in the Earth’s atmosphere. Although you may
not perceive it, atmospheric pressure is constantly pressing down on your body.
This pressure keeps gases within your body, such as the gaseous nitrogen in
body fluids, dissolved. If you were suddenly ejected from a spaceship above
Earth’s atmosphere, you would go from a situation of normal pressure to one of
very low pressure. The pressure of the nitrogen gas in your blood would be
much higher than the pressure of nitrogen in the space surrounding your body.
As a result, the nitrogen gas in your blood would expand, forming bubbles that
could block blood vessels and even cause cells to break apart.

Atmospheric pressure does more than just keep blood gases dissolved. Your
ability to breathe—that is, to take in oxygen and release carbon dioxide—also
depends upon a precise atmospheric pressure. Altitude sickness occurs in part
because the atmosphere at high altitudes exerts less pressure, reducing the
exchange of these gases, and causing shortness of breath, confusion, headache,
lethargy, and nausea. Mountain climbers carry oxygen to reduce the effects of
both low oxygen levels and low barometric pressure at higher altitudes.

The dynamic pressure of body fluids is also important to human survival.


For example, blood pressure, which is the pressure exerted by blood as it flows
within blood vessels, must be great enough to enable blood to reach all body
tissues, and yet low enough to ensure that the delicate blood vessels can
withstand the friction and force of the pulsating flow of pressurized blood.

Application:
General Physiology

Refer to the presented pictures and answer the following questions.

1. How does having multiple levels of organization


benefit a multicellular organism?

2. Describe how structure is related to function in an


organism.

3. Explain why the failure or one organ or organ system


can affect the function of other body systems in an
organism.

Assessment:

Read the following questions and choose the choices that


corresponds to the best answer.
1. A student is observing the chlorophyll of a leaf under a microscope. What
level of organization are they observing?
a. Cell b. Tissue c. Organ d. Organ System

2. A doctor is performing a surgery on a patient's heart. What level of


organization is the doctor working on?
a. Cell b. Tissue c. Organ d. Organ System
General Physiology

3. An ecologist is studying the interactions between different species in a forest


ecosystem. What level of organization are they studying?
a. Cell b. Tissue c. Organ d. Organ System

4. A biology teacher is explaining the structure and function of a human hand.


What level of organization is the teacher explaining?
a. Cell b. Tissue c. Organ d. Organ System

5. A veterinarian is treating a sick dog. What is the highest level of organization


that the veterinarian is considering?
a. Cell b. Tissue c. Organ d. Organ System

6. Your body has just been subjected to extreme cold and you are shivering to
warm up. This is an example of what function of human life?
a. Movement
b. Organization
c. Responsiveness
d. Metabolism

7. You just ate a big meal and now your body is breaking down the food into
smaller molecules to use for energy. This is an example of what function of
human life?
a. Movement
b. Organization
c. Responsiveness
d. Metabolism

8. You just grew an inch taller! This is an example of what function of human
life?
a. Movement
b. Organization
c. Responsiveness
d. Development, growth, and reproduction

9. You just cut your finger and now the cells at the site of the injury are working
to repair the tissue. This is an example of what function of human life?
a. Movement
b. Organization
c. Responsiveness
d. Development, growth, and reproduction

10. The cells in your skin are working to keep the internal structures and fluids
separated from the external environment. This is an example of what function of
human life?
a. Movement
b. Organization
General Physiology

c. Responsiveness
d. Development, growth, and reproduction

11. Which of the following is the most critical nutrient for human survival?
a. Protein
b. Carbohydrates
c. Water
d. Vitamin C

12. How much of an adult's body mass is made up of water?


a. 50%
b. 60%
c. 70%
d. 80%

13. What are the three basic classes of nutrients?


a. Proteins, carbohydrates, and lipids
b. Vitamins, minerals, and water
c. Energy-yielding, micronutrients, and body-building nutrients
d. Proteins, vitamins, and water
14. What is atmospheric pressure?
a. Pressure exerted by the Earth's atmosphere.
b. Pressure exerted by water.
c. Pressure exerted by the sun.
d. Pressure exerted by the moon.

15. What happens if a person is suddenly ejected from a spaceship above


Earth's atmosphere?
a. They will experience an increase in pressure.
b. They will experience no change in pressure.
c. They will experience a decrease in pressure.
d. They will experience an increase in temperature.

Lesson 3
Homeostasis and Evolutionary Adaptation
Objectives:
At the end of the lesson the students shall be able to:
 Discuss the role of homeostasis in healthy functioning;
 Compare and Contrast negative and positive feedback, giving
one physiologic example of each mechanism; and
 Show the body’s mechanism in maintaining internal balance
through a simulation.
General Physiology

Activity:

Virtual Simulation

Go to this link:
https://pbslm-contrib.s3.amazonaws.com/WGBH/conv16/conv16-int-bcc/
index.html

After clicking the link, you will be routed to a virtual simulation. In this
activity, it’s up to you, not the body, to maintain homeostasis in a virtual
person. The gauges in the simulation will show current conditions in the body.
To achieve homeostasis, click the left or right arrows to adjust heart rate,
respiration, blood vessel dilation, perspiration, or blood sugar hormone release.
Your goal is to keep the dials centered. You can also click the button for each
body system feature to learn more.

picture of that shows what is inside the simulation.

Analysis: From the activity, How does the body maintain homeostasis in
response to changes in the internal or external environment?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________
How does diet and exercise affect homeostasis in the body?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________
How does homeostasis relate to the concept of "disequilibrium"
in the body?
General Physiology

Abstraction:

Biological adaptations have two levels of explanation: the


mechanistic (or proximate) and the evolutionary (or ultimate)

Physiologists view organisms as complex systems whose


functions can be explained through sequences of physical and chemical
processes. However, biological traits such as fur color or visual acuity are
unique because they have evolved over millions of years through random
variation and natural selection. In this process, individuals with genetic
characteristics that increase their survival and reproductive success will pass
on those genes to future generations, leading to changes in the genetic makeup
of a species over time.

Therefore, scientists must understand that biological phenomena have


two levels of explanation:

Mechanistic or Proximate Explanation: This answer to the question "How does it


work?" and focuses on the function or structure of a trait and its underlying
mechanisms. This is the core of physiological and medical sciences.

Evolutionary or Ultimate Explanation: This answer to the question "How did it


get to be this way?" and recognizes that biological traits are the result of
evolutionary history, shaped by environmental pressures, mating patterns, and
other selective forces. The interplay between genetic variation and natural
selection has led to the unique features that we observe in different species
today.

Physiologists believe that the process of natural selection has led to the
evolution of adaptations in organisms, which are beneficial traits that enhance
their survival. When conducting evolutionary studies, biologists distinguish
between two types of adaptations: homologous and analogous. Homologous
traits are those that are related through common ancestry, such as bird wings
and human arms, while analogous traits are those that have similar structures
and functions but evolved independently, such as bird and insect wings.

Natural selection has resulted in the evolution of sophisticated


adaptations, but it is also limited by the past. Evolution occurs through
modifications of previously evolved features and variations of these features,
which may not always result in the most efficient or optimal design.

Homeostasis

Maintaining homeostasis requires that the body continuously monitor its


internal conditions. From body temperature to blood pressure to levels of certain
General Physiology

nutrients, each physiological condition has a particular set point. A set point is
the physiological value around which the normal range fluctuates. A normal
range is the restricted set of values that is optimally healthful and stable. For
example, the set point for normal human body temperature is approximately
37°C (98.6°F) Physiological parameters, such as body temperature and blood
pressure, tend to fluctuate within a normal range a few degrees above and below
that point. Control centers in the brain and other parts of the body monitor and
react to deviations from homeostasis using negative feedback. Negative feedback
is a mechanism that reverses a deviation from the set point. Therefore, negative
feedback maintains body parameters within their normal range. The
maintenance of homeostasis by negative feedback always goes on throughout
the body, and an understanding of negative feedback is thus fundamental to an
understanding of human physiology.

Negative Feedback
A negative feedback system has
three basic components. A sensor,
also referred to a receptor, is a
component of a feedback system that
monitors a physiological value. This
value is reported to the control
center. The control center is the
component in a feedback system that
compares the value to the normal
range. If the value deviates too much
from the set point, then the control
center activates an effector. An effector is the component in a feedback system
that causes a change to reverse the situation and return the value to the normal
range.

To set the system in motion, a stimulus must drive a physiological


parameter beyond its normal range (that is, beyond homeostasis). This stimulus
is “heard” by a specific sensor. For example, in the control of blood glucose,
specific endocrine cells in the pancreas detect excess glucose (the stimulus) in
the bloodstream. These pancreatic beta cells respond to the increased level of
blood glucose by releasing the hormone insulin into the bloodstream. The
insulin signals skeletal muscle fibers, fat cells (adipocytes), and liver cells to
take up the excess glucose, removing it from the bloodstream. As glucose
concentration in the bloodstream drops, the decrease in concentration—the
actual negative feedback—is detected by pancreatic alpha cells, and insulin
release stops. This prevents blood sugar levels from continuing to drop below
the normal range.

Humans have a similar temperature regulation feedback system that


works by promoting either heat loss or heat gain. When the brain’s temperature
regulation center receives data from the sensors indicating that the body’s
temperature exceeds its normal range, it stimulates a cluster of brain cells
referred to as the “heat-loss center.” This stimulation has three major effects:
General Physiology

Blood vessels in the skin begin to dilate allowing more blood from the
body core to flow to the surface of the skin allowing the heat to radiate into the
environment.
As blood flow to the skin increases, sweat glands are activated to increase their
output. As the sweat evaporates from the skin surface into the surrounding air,
it takes heat with it.

The depth of respiration increases, and a person may breathe through an


open mouth instead of through the nasal passageways. This further increases
heat loss from the lungs.
In contrast, activation of the brain’s heat-gain center by exposure to cold
reduces blood flow to the skin, and blood returning from the limbs is diverted
into a network of deep veins. This arrangement traps heat closer to the body
core and restricts heat loss. If heat loss is severe, the brain triggers an increase
in random signals to skeletal muscles, causing them to contract and producing
shivering. The muscle contractions of shivering release heat while using up ATP.
The brain triggers the thyroid gland in the endocrine system to release thyroid
hormone, which increases metabolic activity and heat production in cells
throughout the body. The brain also signals the adrenal glands to release
epinephrine (adrenaline), a hormone that causes the breakdown of glycogen into
glucose, which can be used as an energy source. The breakdown of glycogen
into glucose also results in increased metabolism and heat production.

Positive Feedback
Positive feedback intensifies a
change in the body’s physiological
condition rather than reversing it. A
deviation from the normal range results
in more change, and the system moves
farther away from the normal range.
Positive feedback in the body is normal
only when there is a definite end point.
Childbirth and the body’s response to
blood loss are two examples of positive
feedback loops that are normal but are
activated only when needed.

Childbirth at full term is an


example of a situation in which the
maintenance of the existing body state is not desired. Enormous changes in the
mother’s body are required to expel the baby at the end of pregnancy. And the
events of childbirth, once begun, must progress rapidly to a conclusion or the
life of the mother and the baby are at risk. The extreme muscular work of labor
and delivery are the result of a positive feedback system.

The first contractions of labor (the stimulus) push the baby toward the
cervix (the lowest part of the uterus). The cervix contains stretch-sensitive nerve
cells that monitor the degree of stretching (the sensors). These nerve cells send
messages to the brain, which in turn causes the pituitary gland at the base of
General Physiology

the brain to release the hormone oxytocin into the bloodstream. Oxytocin
causes stronger contractions of the smooth muscles in of the uterus (the
effectors), pushing the baby further down the birth canal. This causes even
greater stretching of the cervix. The cycle of stretching, oxytocin release, and
increasingly more forceful contractions stops only when the baby is born. At this
point, the stretching of the cervix halts, stopping the release of oxytocin.

A second example of positive feedback centers on reversing extreme


damage to the body. Following a penetrating wound, the most immediate threat
is excessive blood loss. Less blood circulating means reduced blood pressure
and reduced perfusion (penetration of blood) to the brain and other vital organs.
If perfusion is severely reduced, vital organs will shut down and the person will
die. The body responds to this potential catastrophe by releasing substances in
the injured blood vessel wall that begin the process of blood clotting. As each
step of clotting occurs, it stimulates the release of more clotting substances.
This accelerates the processes of clotting and sealing off the damaged area.
Clotting is contained in a local area based on the tightly controlled availability of
clotting proteins. This is an adaptive, life-saving cascade of events.

Application:

Homeostatic Imbalance Scenario Analysis:


General Physiology

Group yourself into five. Read the scenario below and answer the
following questions below. Present your answer to the class.

A person has a high-salt diet and doesn't drink enough water, causing their
body to become dehydrated. In response, the body's osmoregulatory system works
to maintain homeostasis by retaining water and releasing antidiuretic hormone
(ADH) to reduce urine output. If the body's response to dehydration is inadequate,
the person may experience symptoms such as increased thirst, dry mouth, and
dark yellow urine. In severe cases, dehydration can lead to serious health
problems such as heat stroke, kidney failure, and even death.

1. What is the cause of the homeostatic imbalance in this scenario?

2. What is the role of the osmoregulatory system in maintaining homeostasis


in this scenario?

3. How does the body respond to dehydration in order to maintain


homeostasis?

4. What happens if the body's response to dehydration is inadequate?

5. How does a high-salt diet contribute to dehydration and the homeostatic


imbalance in this scenario?
6. What are the potential consequences of an inadequate response to
dehydration?

7. How does the body's response to dehydration impact other physiological


systems and processes?

8. How might this scenario be prevented or managed to maintain


homeostasis?

Assessment:

Read the following questions and choose the choices that


corresponds to the best answer.
General Physiology

1. Which of the following is a good definition of homeostasis?


A. Homeostasis means that our body monitors the value of several parameters
and prevents them from ever changing.
B. Homeostasis means that our body doesn't regulate the value of different body
parameters.
C. Homeostasis means that our body monitors the value of several parameters
and restores them to their correct value after any alteration.
D. Homeostasis means that we stand still all the time.

2. Each of the following are the variables do our bodies regulate EXCEPT.
A.Salt concentration in blood
B.Glucose concentration in blood
C.Body temperature
D.Blood volume

3. Which of the following is an example of positive feedback to indirectly


maintain homeostasis?
A. the chemical reactions involved in blood clotting.
B. your blood vessels dilate and you begin to sweat in response to elevated body
temperature.
C. your blood vessels constrict and you begin to shiver in response to low body
temperature.
D. your coach compliments you on your performance in practice.

4. When glucose levels in the blood rise, your brain sends a signal to your
pancreas. The pancreas releases insulin, which opens channels in cell
membranes to allow glucose to enter the cell, lowering blood sugar levels. What
kind of mechanism this is under homeostasis?
A. Negative feedback mechanism
B. Positive feedback mechanism
C. Excitatory mechanism
D. Inhibitory mechanism

5. What happens if the body's response to a stressor is inadequate?


a. The body maintains homeostasis.
b. The body experiences a homeostatic imbalance.
c. The body experiences an increase in growth.
d. The body experiences a decrease in growth.

General Physiology
Module 2: Animal
Physiology
General Physiology

Welcome to the Module 2!


The human body is a complex
organism composed of various
interrelated systems that work together
to maintain its overall functioning. In
this module, we will be exploring four of
these systems, namely the nervous
system, cardiovascular system, digestive
system, and reproductive system.
The nervous system is responsible
for transmitting electrical signals
between the brain and different parts of
the body.
The cardiovascular system
consists of the heart and blood vessels
and is responsible for distributing
oxygen, hormones, and nutrients to the
body's cells.
The digestive system breaks down
food into smaller components that can
be absorbed and used for energy and cell
growth.
The reproductive system is
responsible for producing and
transporting sperm and eggs, and for
supporting the development of a fetus.
By the end of this module, you
will have a deeper understanding of
these systems, how they work
individually and together to maintain the
health and proper functioning of the
human body.

Lesson 1
Regulation, Integration, and Control
Objectives:
General Physiology

At the end of the lesson the students shall be able to:

 Distinguish the major functions of the nervous system: sensation,


integration, and response;
 List the sequence of events in a simple sensory receptor–motor
response pathway;
 Describe the components of the membrane that establish the
resting membrane potential;
 Describe the changes that occur to the membrane that result in
the action potential;
 Explain the differences between the types of graded potentials;
and
 Categorize the major neurotransmitters by chemical type and
effect.

Activity:
“Label the Puzzle”
Look for information about Reflex Arc. Label the reflex arc puzzle
from the information you have gathered and present it to the
class.

Analysis: From the activity, how does information travel through the
reflex arc to produce a reflex response?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________
What role does the central nervous system play in a reflex arc?
General Physiology

Abstraction:

The Function of Nervous Tissue

Having looked at the components of nervous tissue, and the


basic anatomy of the nervous system, next comes an
understanding of how nervous tissue is capable of communicating within the
nervous system. Before getting to the nuts and bolts of how this works, an
illustration of how the components come together will be helpful.

Imagine you are about to take a


shower in the morning before going to
school. You have turned on the faucet
to start the water as you prepare to get
in the shower. After a few minutes, you
expect the water to be a temperature
that will be comfortable to enter. So
you put your hand out into the spray
of water. What happens next depends
on how your nervous system interacts
with the stimulus of the water
temperature and what you do in
response to that stimulus.
Found in the skin of your fingers
or toes is a type of sensory receptor that
is sensitive to temperature, called
a thermoreceptor. When you place
your hand under the shower, the cell
membrane of the thermoreceptors
changes its electrical state (voltage). The
General Physiology

amount of change is dependent on the strength of the stimulus (how hot the
water is). This is called a graded potential. If the stimulus is strong, the voltage
of the cell membrane will change enough to generate an electrical signal that
will travel down the axon. You have learned about this type of signaling before,
with respect to the interaction of nerves and muscles at the neuromuscular
junction. The voltage at which such a signal is generated is called
the threshold, and the resulting electrical signal is called an action potential.
In this example, the action potential travels—a process known as propagation—
along the axon from the axon hillock to the axon terminals and into the synaptic
end bulbs. When this signal reaches the end bulbs, it causes the release of a
signaling molecule called a neurotransmitter.

The neurotransmitter diffuses across the short distance of the synapse


and binds to a receptor protein of the target neuron. When the molecular signal
binds to the receptor, the cell membrane of the target neuron changes its
electrical state and a new graded potential begins. If that graded potential is
strong enough to reach threshold, the second neuron generates an action
potential at its axon hillock. The target of this neuron is another neuron in
the thalamus of the brain, the part of the CNS that acts as a relay for sensory
information. At another synapse, neurotransmitter is released and binds to its
receptor. The thalamus then sends the sensory information to the cerebral
cortex, the outermost layer of gray matter in the brain, where conscious
perception of that water temperature begins.

Within the cerebral cortex,


information is processed among many
neurons, integrating the stimulus of the
water temperature with other sensory
stimuli, with your emotional state (you
just aren't ready to wake up; the bed is
calling to you), memories (perhaps of the
lab notes you have to study before a
quiz). Finally, a plan is developed about
what to do, whether that is to turn the
temperature up, turn the whole shower
off and go back to bed, or step into the
shower. To do any of these things, the
cerebral cortex has to send a command out to your body to move muscles.

A region of the cortex is specialized for sending signals down to the spinal
cord for movement. The upper motor neuron is in this region, called
the precentral gyrus of the frontal cortex, which has an axon that extends all
the way down the spinal cord. At the level of the spinal cord at which this axon
makes a synapse, a graded potential occurs in the cell membrane of a lower
motor neuron. This second motor neuron is responsible for causing muscle
fibers to contract. In the manner described in the chapter on muscle tissue, an
action potential travels along the motor neuron axon into the periphery. The
axon terminates on muscle fibers at the neuromuscular junction. Acetylcholine
General Physiology

is released at this specialized synapse, which causes the muscle action potential
to begin, following a large potential known as an end plate potential. When the
lower motor neuron excites the muscle fiber, it contracts. All of this occurs in a
fraction of a second, but this story is the basis of how the nervous system
functions.

The Action Potential

Electrically Active Cell Membranes

Most cells in the body make use of charged particles, ions, to build up a
charge across the cell membrane. Previously, this was shown to be a part of how
muscle cells work. For skeletal muscles to contract, based on excitation–
contraction coupling, requires input from a neuron. Both cells make use of the
cell membrane to regulate ion movement between the extracellular fluid and
cytosol.

As you learned in the chapter on cells, the cell membrane is primarily


responsible for regulating what can cross the membrane and what stays on only
one side. The cell membrane
is a phospholipid bilayer, so
only substances that can
pass directly through the
hydrophobic core can diffuse
through unaided. Charged
particles, which are
hydrophilic by definition,
cannot pass through the cell
membrane without ass.

Transmembrane
proteins, specifically channel proteins, make this possible. Several passive
transport channels, as well as active transport pumps, are necessary to
generate a transmembrane potential and an action potential. Of special interest
is the carrier protein referred to as the sodium/potassium pump that moves
sodium ions (Na+) out of a cell and potassium ions (K +) into a cell, thus
regulating ion concentration on both sides of the cell membrane.

The sodium/potassium pump requires energy in the form of adenosine


triphosphate (ATP), so it is also referred to as an ATPase. As was explained in
the cell chapter, the concentration of Na + is higher outside the cell than inside,
and the concentration of K + is higher inside the cell than outside. That means
that this pump is moving the ions against the concentration gradients for
sodium and potassium, which is why it requires energy. In fact, the pump
basically maintains those concentration gradients.

Ion channels are pores that allow specific charged particles to cross the
membrane in response to an existing concentration gradient. Proteins are
General Physiology

capable of spanning the cell membrane, including its hydrophobic core, and can
interact with the charge of ions because of the varied properties of amino acids
found within specific domains or regions of the protein channel. Hydrophobic
amino acids are found in the domains that are apposed to the hydrocarbon tails
of the phospholipids. Hydrophilic amino acids are exposed to the fluid
environments of the extracellular fluid and cytosol. Additionally, the ions will
interact with the hydrophilic amino acids, which will be selective for the charge
of the ion. Channels for cations (positive ions) will have negatively charged side
chains in the pore. Channels for anions (negative ions) will have positively
charged side chains in the pore. This is called electrochemical exclusion,
meaning that the channel pore is charge-specific.

Ion channels can also be specified by the diameter of the pore. The
distance between the amino acids will be specific for the diameter of the ion
when it dissociates from the water molecules surrounding it. Because of the
surrounding water molecules, larger pores are not ideal for smaller ions because
the water molecules will interact, by hydrogen bonds, more readily than the
amino acid side chains. This is called size exclusion. Some ion channels are
selective for charge but not necessarily for size, and thus are called
a nonspecific channel. These nonspecific channels allow cations—particularly
Na+, K+, and Ca2+—to cross the membrane, but exclude anions.

Ion channels do not always freely allow ions to diffuse across the
membrane. Some are opened by certain events, meaning the channels
are gated. So another way that channels can be categorized is on the basis of
how they are gated. Although these classes of ion channels are found primarily
in the cells of nervous or muscular tissue, they also can be found in the cells of
epithelial and connective tissues.

A ligand-gated channel opens because a signaling molecule, a ligand,


binds to the extracellular
region of the channel. This
type of channel is also known
as an ionotropic
receptor because when the
ligand, known as a
neurotransmitter in the
nervous system, binds to the
protein, ions cross the
membrane changing its
charge.

A mechanically gated channel opens because of a physical distortion of the


cell membrane. Many channels associated with the sense of touch (somato-
sensation) are mechanically
gated. For example, as
pressure is applied to the
skin, these channels open
and allow ions to enter the
General Physiology

cell. Similar to this type of channel would be the channel that opens on the
basis of temperature changes, as in testing the water in the shower.

A voltage-gated channel is a channel that responds to changes in the electrical


properties of the membrane in
which it is embedded.
Normally, the inner portion of
the membrane is at a negative
voltage. When that voltage
becomes less negative, the
channel begins to allow ions to
cross the membrane.

A leakage channel is
randomly gated, meaning that it opens and closes at random, hence the
reference to leaking. There is
no actual event that opens
the channel; instead, it has
an intrinsic rate of switching
between the open and closed
states. Leakage channels
contribute to the resting
transmembrane voltage of the
excitable membrane.

The Membrane Potential

The electrical state of the cell


membrane can have several
variations. These are all variations in
the membrane potential. A potential
is a distribution of charge across the
General Physiology

cell membrane, measured in millivolts (mV). The standard is to compare the


inside of the cell relative to the outside, so the membrane potential is a value
representing the charge on the intracellular side of the membrane based on the
outside being zero, relatively speaking.

The concentration of ions in extracellular and intracellular fluids is largely


balanced, with a net neutral charge. However, a slight difference in charge
occurs right at the membrane surface, both internally and externally. It is the
difference in this very limited region that has all the power in neurons (and
muscle cells) to generate electrical signals, including action potentials.

Before these electrical signals can be described, the resting state of the
membrane must be explained. When the cell is at rest, and the ion channels are
closed (except for leakage channels which randomly open), ions are distributed
across the membrane in a very predictable way. The concentration of
Na+ outside the cell is 10 times greater than the concentration inside. Also, the
concentration of K+ inside the cell is greater than outside. The cytosol contains a
high concentration of anions, in the form of phosphate ions and negatively
charged proteins. Large anions are a component of the inner cell membrane,
including specialized phospholipids and proteins associated with the inner
leaflet of the membrane (leaflet is a term used for one side of the lipid bilayer
membrane). The negative charge is localized in the large anions.

With the ions distributed across the membrane at these concentrations,


the difference in charge is measured at -70 mV, the value described as
the resting membrane potential. The exact value measured for the resting
membrane potential varies between cells, but -70 mV is most commonly used as
this value. This voltage would actually be much lower except for the
contributions of some important proteins in the membrane. Leakage channels
allow Na+ to slowly move into the cell or K + to slowly move out, and the
Na+/K+ pump restores them. This may appear to be a waste of energy, but each
has a role in maintaining the membrane potential.

The Action Potential

Resting membrane potential describes the steady state of the cell, which
is a dynamic process that is balanced by ion leakage and ion pumping. Without
any outside influence, it will not change. To get an electrical signal started, the
membrane potential has to change.

This starts with a channel opening for Na+ in the membrane. Because the
concentration of Na+ is higher outside the cell than inside the cell by a factor of
10, ions will rush into the cell that are driven largely by the concentration
gradient. Because sodium is a positively charged ion, it will change the relative
voltage immediately inside the cell relative to immediately outside. The resting
potential is the state of the membrane at a voltage of -70 mV, so the sodium
cation entering the cell will cause it to become less negative. This is known
as depolarization, meaning the membrane potential moves toward zero.
General Physiology

The concentration gradient for Na + is so strong that it will continue to


enter the cell even after the membrane potential has become zero, so that the
voltage immediately around the pore begins to become positive. The electrical
gradient also plays a role, as negative proteins below the membrane attract the
sodium ion. The membrane potential will reach +30 mV by the time sodium has
entered the cell.

As the membrane potential reaches +30 mV, other voltage-gated channels


are opening in the membrane. These channels are specific for the potassium
ion. A concentration gradient acts on K +, as well. As K+ starts to leave the cell,
taking a positive charge with it, the membrane potential begins to move back
toward its resting voltage. This is called repolarization, meaning that the
membrane voltage moves back toward the -70 mV value of the resting
membrane potential.

Repolarization returns the membrane potential to the -70 mV value that


indicates the resting potential, but it actually overshoots that value. Potassium
ions reach equilibrium when the membrane voltage is below -70 mV, so a period
of hyperpolarization occurs while the K + channels are open. Those K+ channels
are slightly delayed in closing, accounting for this short overshoot.

What has been described here is the action potential, which is presented
as a graph of voltage over time in. It is the electrical signal that nervous tissue
generates for communication. The change in the membrane voltage from -70 mV
at rest to +30 mV at the end of depolarization is a 100 mV change. That can also
be written as a 0.1 V change. To put that value in perspective, think about a
battery. An AA battery that you might find in a television remote has a voltage of
1.5 V, or a 9 V battery (the rectangular battery with two posts on one end) is,
obviously, 9 V. The change
seen in the action potential is
one or two orders of
magnitude less than the
charge in these batteries. In
fact, the membrane potential
can be described as a battery.
A charge is stored across the
membrane that can be
released under the correct
conditions. A battery in your
remote has stored a charge
that is “released” when you
push a button.

The question is, now, what initiates the action potential? The description
above conveniently glosses over that point. But it is vital to understanding what
is happening. The membrane potential will stay at the resting voltage until
something changes. The description above just says that a Na + channel opens.
Now, to say “a channel opens” does not mean that one individual
transmembrane protein changes. Instead, it means that one kind of channel
General Physiology

opens. There are a few different types of channels that allow Na + to cross the
membrane. A ligand-gated Na+ channel will open when a neurotransmitter binds
to it and a mechanically gated Na+ channel will open when a physical stimulus
affects a sensory receptor (like pressure applied to the skin compresses a touch
receptor). Whether it is a neurotransmitter binding to its receptor protein or a
sensory stimulus activating a sensory receptor cell, some stimulus gets the
process started. Sodium starts to enter the cell and the membrane becomes less
negative.

A third type of channel that is an important part of depolarization in the


action potential is the voltage-gated Na + channel. The channels that start
depolarizing the membrane because of a stimulus help the cell to depolarize
from -70 mV to -55 mV. Once the membrane reaches that voltage, the voltage-
gated Na+ channels open. This is what is known as the threshold. Any
depolarization that does not change the membrane potential to -55 mV or higher
will not reach threshold and thus will not result in an action potential. Also, any
stimulus that depolarizes the membrane to -55 mV or beyond will cause a large
number of channels to open and an action potential will be initiated.

Because of the threshold, the action potential can be likened to a digital


event—it either happens or it does not. If the threshold is not reached, then no
action potential occurs. If depolarization reaches -55 mV, then the action
potential continues and runs all the way to +30 mV, at which K + causes
repolarization, including the hyperpolarizing overshoot. Also, those changes are
the same for every action potential, which means that once the threshold is
reached, the exact same thing happens. A stronger stimulus, which might
depolarize the membrane well past threshold, will not make a “bigger” action
potential. Action potentials are “all or none.” Either the membrane reaches the
threshold and everything occurs as described above, or the membrane does not
reach the threshold and nothing else happens. All action potentials peak at the
same voltage (+30 mV), so one action potential is not bigger than another.
Stronger stimuli will initiate multiple action potentials more quickly, but the
individual signals are not bigger. Thus, for example, you will not feel a greater
sensation of pain, or have a stronger muscle contraction, because of the size of
the action potential because they are not different sizes.

As we have seen, the depolarization and repolarization of an action


potential are dependent on two types of channels (the voltage-gated Na + channel
and the voltage-gated K+ channel). The voltage-gated Na+ channel actually has
two gates. One is the activation gate, which opens when the membrane
potential crosses -55 mV. The other gate is the inactivation gate, which closes
after a specific period of time—on the order of a fraction of a millisecond. When
a cell is at rest, the activation gate is closed and the inactivation gate is open.
However, when the threshold is reached, the activation gate opens, allowing
Na+ to rush into the cell. Timed with the peak of depolarization, the inactivation
gate closes. During repolarization, no more sodium can enter the cell. When the
membrane potential passes -55 mV again, the activation gate closes. After that,
the inactivation gate re-opens, making the channel ready to start the whole
process over again.
General Physiology

The voltage-gated K+ channel has only one gate, which is sensitive to a


membrane voltage of -50 mV. However, it does not open as quickly as the
voltage-gated Na+ channel does. It might take a fraction of a millisecond for the
channel to open once that voltage has been reached. The timing of this coincides
exactly with when the Na+ flow peaks, so voltage-gated K+ channels open just as
the voltage-gated Na+ channels are being inactivated. As the membrane potential
repolarizes and the voltage passes -50 mV again, the channel closes—again,
with a little delay. Potassium continues to leave the cell for a short while and
the membrane potential becomes more negative, resulting in the hyperpolarizing
overshoot. Then the channel closes again and the membrane can return to the
resting potential because of the ongoing activity of the non-gated channels and
the Na+/K+ pump.

All of this takes place within


approximately 2 milliseconds. While an
action potential is in progress, another one
cannot be initiated. That effect is referred to
as the refractory period. There are two
phases of the refractory period:
the absolute refractory period and
the relative refractory period. During the
absolute phase, another action potential
will not start. This is because of the inactivation gate of the voltage-gated
Na+ channel. Once that channel is back to its resting conformation (less than -
55 mV), a new action potential could be started, but only by a stronger stimulus
than the one that initiated the current action potential. This is because of the
flow of K+ out of the cell. Because that ion is rushing out, any Na + that tries to
enter will not depolarize the cell, but will only keep the cell from hyperpolarizing.

Propagation of the Action Potential

The action potential is initiated at the beginning of the axon, at what is


called the initial segment. There is a high density of voltage-gated Na + channels
so that rapid depolarization can take place here. Going down the length of the
axon, the action potential is propagated because more voltage-gated
Na+ channels are opened as the depolarization spreads. This spreading occurs
because Na+ enters through the channel and moves along the inside of the cell
membrane. As the Na+ moves, or flows, a short distance along the cell
membrane, its positive charge depolarizes a little more of the cell membrane. As
that depolarization spreads, new voltage-gated Na+ channels open and more ions
rush into the cell, spreading the depolarization a little farther.

Because voltage-gated Na+ channels are inactivated at the peak of the


depolarization, they cannot be opened again for a brief time—the absolute
refractory period. Because of this, depolarization spreading back toward
previously opened channels has no effect. The action potential must propagate
General Physiology

toward the axon terminals; as a result, the polarity of the neuron is maintained,
as mentioned above.

Propagation, as described above, applies to unmyelinated axons. When


myelination is present, the action potential propagates differently. Sodium ions
that enter the cell at the initial segment start to spread along the length of the
axon segment, but there is no voltage-gated Na + channels until the first node of
Ranvier. Because there is not constant opening of these channels along the
axon segment, the depolarization spreads at an optimal speed. The distance
between nodes is the optimal distance to keep the membrane still depolarized
above threshold at the next node. As Na + spreads along the inside of the
membrane of the axon segment, the charge starts to dissipate. If the node were
any farther down the axon, that depolarization would have fallen off too much
for voltage-gated Na+ channels to be activated at the next node of Ranvier. If the
nodes were any closer together, the speed of propagation would be slower.

Propagation along an unmyelinated axon is referred to as continuous


conduction; along the length of a myelinated axon, it is saltatory conduction.
Continuous conduction is slow because there are always voltage-gated
Na+ channels opening, and more and more Na+ is rushing into the cell. Saltatory
conduction is faster because the action potential basically jumps from one node
to the next (saltare = “to leap”), and the new influx of Na + renews the depolarized
membrane. Along with the myelination of the axon, the diameter of the axon can
influence the speed of conduction. Much as water runs faster in a wide river
than in a narrow creek, Na+-based depolarization spreads faster down a wide
axon than down a narrow one. This concept is known as resistance and is
generally true for electrical wires or plumbing, just as it is true for axons,
although the specific conditions are different at the scales of electrons or ions
versus water in a river.

Communication Between Neurons


Graded Potentials

Local changes in the membrane potential are called graded potentials and
are usually associated with the dendrites of a neuron. The amount of change in
the membrane potential is determined by the size of the stimulus that causes it.
In the example of testing the temperature of the shower, slightly warm water
would only initiate a small change in a thermoreceptor, whereas hot water
would cause a large amount of change in the membrane potential.
General Physiology

Graded potentials can be of two sorts, either they are depolarizing or


hyperpolarizing. For a membrane
at the resting potential, a graded
potential represents a change in
that voltage either above -70 mV
or below -70 mV. Depolarizing
graded potentials are often the
result of Na+ or Ca2+ entering the
cell. Both of these ions have
higher concentrations outside the
cell than inside; because they
have a positive charge, they will
move into the cell causing it to
become less negative relative to
the outside. Hyperpolarizing graded potentials can be caused by K + leaving the
cell or Cl- entering the cell. If a positive charge moves out of a cell, the cell
becomes more negative; if a negative charge enters the cell, the same thing
happens.

Types of Graded Potentials

For the unipolar cells of sensory neurons—both those with free nerve
endings and those within encapsulations—graded potentials develop in the
dendrites that influence the generation of an action potential in the axon of the
same cell. This is called a generator potential. For other sensory receptor cells,
such as taste cells or photoreceptors of the retina, graded potentials in their
membranes result in the release of neurotransmitters at synapses with sensory
neurons. This is called a receptor potential.

A postsynaptic potential (PSP) is the graded potential in the dendrites of


a neuron that is receiving synapses from other cells. Postsynaptic potentials can
be depolarizing or hyperpolarizing. Depolarization in a postsynaptic potential is
called an excitatory postsynaptic potential (EPSP) because it causes the
membrane potential to move toward threshold. Hyperpolarization in a
postsynaptic potential is an inhibitory postsynaptic potential (IPSP) because
it causes the membrane potential to move away from threshold.

Summation

All types of graded potentials will


result in small changes of either
depolarization or hyperpolarization in the
voltage of a membrane. These changes can
lead to the neuron reaching threshold if the
changes add together or summate. The
combined effects of different types of
graded potentials are illustrated. If the total
change in voltage in the membrane is a
positive 15 mV, meaning that the
General Physiology

membrane depolarizes from -70 mV to -55 mV, then the graded potentials will
result in the membrane reaching threshold.

For receptor potentials, threshold is not a factor because the change in


membrane potential for receptor cells directly causes neurotransmitter release.
However, generator potentials can initiate action potentials in the sensory
neuron axon, and postsynaptic potentials can initiate an action potential in the
axon of other neurons. Graded potentials summate at a specific location at the
beginning of the axon to initiate the action potential, namely the initial segment.
For sensory neurons, which do not have a cell body between the dendrites and
the axon, the initial segment is directly adjacent to the dendritic endings. For all
other neurons, the axon hillock is essentially the initial segment of the axon,
and it is where summation takes place. These locations have a high density of
voltage-gated Na+ channels that initiate the depolarizing phase of the action
potential.

Summation can be spatial or temporal, meaning it can be the result of


multiple graded potentials at different locations on the neuron, or all at the
same place but separated in time. Spatial summation is related to associating
the activity of multiple inputs to a neuron with each other. Temporal
summation is the relationship of multiple action potentials from a single cell
resulting in a significant change in the membrane potential. Spatial and
temporal summation can act together, as well.

Synapses

There are two types of connections between electrically active cells,


chemical synapses and electrical synapses. In a chemical synapse, a chemical
signal—namely, a neurotransmitter—is released from one cell and it affects the
other cell. In an electrical synapse, there is a direct connection between the
two cells so that ions can pass directly from one cell to the next. If one cell is
depolarized in an electrical synapse, the joined cell also depolarizes because the
ions pass between the cells. Chemical synapses involve the transmission of
chemical information from one cell to the next. This section will concentrate on
the chemical type of synapse.

An example of a chemical synapse is the neuromuscular junction (NMJ)


described in the chapter on muscle tissue. In the nervous system, there are
many more synapses that are essentially the same as the NMJ. All synapses
have common characteristics, which can be summarized in this list:

 presynaptic element
 neurotransmitter (packaged in vesicles)
 synaptic cleft
 receptor proteins
 postsynaptic element
 neurotransmitter elimination or re-uptake
General Physiology

For the NMJ, these characteristics are as follows: the presynaptic element is
the motor neuron's axon terminals, the neurotransmitter is acetylcholine, the
synaptic cleft is the space between the cells where the neurotransmitter
diffuses, the receptor protein is the nicotinic acetylcholine receptor, the
postsynaptic element is the sarcolemma of the muscle cell, and the
neurotransmitter is eliminated by acetylcholinesterase. Other synapses are
similar to this, and the specifics are different, but they all contain the same
characteristics.

Neurotransmitter Release

When an action potential reaches the axon


terminals, voltage-gated Ca2+ channels in the
membrane of the synaptic end bulb open. The
concentration of Ca2+ increases inside the end
bulb, and the Ca2+ ion associates with proteins in
the outer surface of neurotransmitter vesicles.
The Ca2+ facilitates the merging of the vesicle with
the presynaptic membrane so that the
neurotransmitter is released through exocytosis
into the small gap between the cells, known as
the synaptic cleft.

Once in the synaptic cleft, the


neurotransmitter diffuses the short distance to the postsynaptic membrane and
can interact with neurotransmitter receptors. Receptors are specific for the
neurotransmitter, and the two fit together like a key and lock. One
neurotransmitter binds to its receptor and will not bind to receptors for other
neurotransmitters, making the binding a specific chemical event.

Neurotransmitter Systems

The important thing to remember


about neurotransmitters, and signaling
chemicals in general, is that the effect is
entirely dependent on the receptor.
Neurotransmitters bind to one of two classes
of receptors at the cell surface, ionotropic or
metabotropic. Ionotropic receptors are
ligand-gated ion channels, such as the
nicotinic receptor for acetylcholine or the
glycine receptor. A metabotropic
receptor involves a complex of proteins that
result in metabolic changes within the cell.
The receptor complex includes the
transmembrane receptor protein, a G
protein, and an effector protein. The
neurotransmitter, referred to as the first
messenger, binds to the receptor protein on
General Physiology

the extracellular surface of the cell, and the intracellular side of the protein
initiates activity of the G protein. The G protein is a guanosine triphosphate
(GTP) hydrolase that physically moves from the receptor protein to the effector
protein to activate the latter. An effector protein is an enzyme that catalyzes
the generation of a new molecule, which acts as the intracellular mediator of the
signal that binds to the receptor. This intracellular mediator is called the second
messenger.

Different receptors use different second messengers. Two common


examples of second messengers are cyclic adenosine monophosphate (cAMP)
and inositol triphosphate (IP3). The enzyme adenylate cyclase (an example of an
effector protein) makes cAMP, and phospholipase C is the enzyme that makes
IP3. Second messengers, after they are produced by the effector protein, cause
metabolic changes within the cell. These changes are most likely the activation
of other enzymes in the cell. In neurons, they often modify ion channels, either
opening or closing them. These enzymes can also cause changes in the cell,
such as the activation of genes in the nucleus, and therefore the increased
synthesis of proteins. In neurons, these kinds of changes are often the basis of
stronger connections between cells at the synapse and may be the basis of
learning and memory.

Application:

“Nervous System in Action”

Instructions:
1. By group, provide a scenario that requires the use of the nervous system,
such as reaching for a glass of water, typing on a keyboard, or playing a musical
instrument.

2. Identify the different components of the nervous system involved in the


scenario that you have given, including sensory receptors, afferent neurons, the
spinal cord, interneurons, efferent neurons, and effectors.

3. With your group, create a diagram of the nervous system involved in their
scenario by using a large poster board. The diagram should show the path of
the electrical signals from the sensory receptors to the effectors and should also
include information about the role of each component of the nervous system.

4. Once the diagrams are completed, present your group’s "Nervous System in
Action" simulation to the class.
General Physiology

Assessment:

Read the following questions and choose the choices that


corresponds to the best answer.
1. What type of sensory receptor is found in the skin of fingers and toes?
A. Photoreceptor
B. Thermoreceptor
C. Proprioreceptor
D. Nociceptor

2. What happens to the electrical state of the thermoreceptor's cell membrane


when a hot stimulus is applied?
A. The electrical state remains the same
B. The electrical state decreases
C. The electrical state increases
D. The electrical state oscillates

3. What is the name of the electrical signal generated by the graded potential
when the stimulus reaches a certain threshold?
A. Action potential
B. Propagation signal
C. Neurotransmitter
D. Receptor protein

4. What part of the brain acts as a relay for sensory information from the skin
receptors?
A. The thalamus
B. The cerebellum
C. The medulla
D. The pons

5. What is the final result of processing the sensory information in the cerebral
cortex?
A. The release of neurotransmitter
B. The generation of an action potential
C. A plan of movement
D. The contraction of muscle fibers

6. What type of ion channel opens because of a physical distortion of the cell
membrane?
General Physiology

A. Ligand-gated channel
B. Mechanically gated channel
C. Voltage-gated channel
D. Leakage channel

7. What is the main difference between a mechanically gated channel and a


voltage-gated channel?
A. A mechanically gated channel opens because of a physical distortion of the
cell membrane, while a voltage-gated channel opens because of changes in the
electrical properties of the membrane.
B. A mechanically gated channel opens in response to temperature changes,
while a voltage-gated channel opens in response to pressure.
C. A mechanically gated channel opens randomly, while a voltage-gated channel
opens in response to changes in voltage.
D. Both channels open in response to changes in the cell membrane.

8. What is the main function of the cell membrane in regulating ion movement?
A. It regulates the concentration of ions in the extracellular fluid and cytosol.
B. It selectively allows specific charged particles to cross the membrane.
C. It determines the voltage of the excitable membrane.
D. It opens and closes channels at random.

9. What type of transport channels are necessary to generate a transmembrane


potential and an action potential?
A. Passive transport channels
B. Active transport pumps
C. Ligand-gated channels
D. Both A and B

10. How do ion channels determine the type of ions that can cross the
membrane?
A. By the diameter of the pore
B. By the charge of the ion
C. By the size of the ion
D. By the voltage of the membrane
General Physiology

Lesson 2
Fluid and Transport
Objectives:
At the end of the lesson the students shall be able to:

 Distinguish between systolic pressure, diastolic pressure, pulse


pressure, and mean arterial pressure;
 Describe the clinical measurement of pulse and blood pressure;
 Identify and discuss five variables affecting arterial blood flow
and blood pressure;
 Discuss several factors affecting blood flow in the venous
system;
 Create a flow chart showing the major systemic veins through
which blood travels from the feet to the right atrium of the heart;
 Distinguish between capillary hydrostatic pressure and blood
colloid osmotic pressure, explaining the contribution of each to
net filtration pressure;
 Compare filtration and reabsorption; and
 Explain the fate of fluid that is not reabsorbed from the tissues
into the vascular capillaries.

Activity:
“Label the Puzzle”
Click this link and
watch the video
about the Arteries
of the Cardio- Vascular System,
and label the puzzle below.
https:// www.kenhub.com/
en/library/ learning-strategies/
cardiovascular-
system- diagrams-quizzes-
and-free- worksheets
General Physiology

Analysis: From the activity, what do you think are the factors affecting
blood pressure?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________
What role does cardiovascular system plays in transporting
materials all throughout the body?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________

Abstraction:

Blood Flow, Blood Pressure and Resistance

Blood flow refers to the movement of blood through a vessel,


tissue, or organ, and is usually expressed in terms of volume of blood per unit of
time. It is initiated by the contraction of the ventricles of the heart. Ventricular
contraction ejects blood into the major arteries, resulting in flow from regions of
higher pressure to regions of lower pressure, as blood encounters smaller
arteries and arterioles, then capillaries, then the venules and veins of the
venous system. This section discusses a number of critical variables that
contribute to blood flow throughout the body. It also discusses the factors that
impede or slow blood flow, a phenomenon known as resistance.
General Physiology

Hydrostatic pressure is the force exerted by a fluid due to gravitational


pull, usually against the wall of the container in which it is located. One form of
hydrostatic pressure is blood pressure, the force exerted by blood upon the
walls of the blood vessels or the chambers of the heart. Blood pressure may be
measured in capillaries and veins, as well as the vessels of the pulmonary
circulation; however, the term blood pressure without any specific descriptors
typically refers to systemic arterial blood pressure—that is, the pressure of
blood flowing in the arteries of the systemic circulation. In clinical practice, this
pressure is measured in mm Hg and is usually obtained using the brachial
artery of the arm.

Components of Arterial Blood Pressure


Arterial blood pressure in the larger vessels consists of several distinct
components: systolic and diastolic pressures, pulse pressure, and mean arterial
pressure.

 Systolic and Diastolic Pressures


When systemic arterial blood pressure is measured, it is recorded as a ratio of
two numbers (e.g., 120/80 is a normal adult blood pressure), expressed as
systolic pressure over diastolic pressure. The systolic pressure is the higher
value (typically around 120 mm Hg) and reflects the arterial pressure resulting
from the ejection of blood during ventricular contraction, or systole.
The diastolic pressure is the lower value (usually about 80 mm Hg) and
represents the arterial pressure of blood during ventricular relaxation, or
diastole.
 Pulse Pressure
The difference between the systolic
pressure and the diastolic pressure is
the pulse pressure. For example, an
individual with a systolic pressure of
120 mm Hg and a diastolic pressure of
80 mm Hg would have a pulse
pressure of 40 mmHg.

Generally, a pulse pressure


should be at least 25 percent of the
systolic pressure. A pulse pressure
below this level is described as low
or narrow. This may occur, for
example, in patients with a low stroke volume, which may be seen in congestive
heart failure, stenosis of the aortic valve, or significant blood loss following
trauma. In contrast, a high or wide pulse pressure is common in healthy people
following strenuous exercise, when their resting pulse pressure of 30–40 mm Hg
General Physiology

may increase temporarily to 100 mm Hg as stroke volume increases. A


persistently high pulse pressure at or above 100 mm Hg may indicate excessive
resistance in the arteries and can be caused by a variety of disorders. Chronic
high resting pulse pressures can degrade the heart, brain, and kidneys, and
warrant medical treatment.
 Mean Arterial Pressure
Mean arterial pressure (MAP) represents the “average” pressure of blood in the
arteries, that is, the average force driving blood into vessels that serve the
tissues. Mean is a statistical concept and is calculated by taking the sum of the
values divided by the number of values. Although complicated to measure
directly and complicated to calculate, MAP can be approximated by adding the
diastolic pressure to one-third of the pulse pressure or systolic pressure minus
the diastolic pressure:
MAP = diastolic BP + (systolic-diastolic BP)3MAP = diastolic BP + (systolic-
diastolic BP)3

this value is approximately 80 + (120 − 80) / 3, or 93.33. Normally, the MAP


falls within the range of 70–110 mm Hg. If the value falls below 60 mm Hg for
an extended time, blood pressure will not be high enough to ensure circulation
to and through the tissues, which results in ischemia, or insufficient blood
flow. A condition called hypoxia, inadequate oxygenation of tissues, commonly
accompanies ischemia. The term hypoxemia refers to low levels of oxygen in
systemic arterial blood. Neurons are especially sensitive to hypoxia and may die
or be damaged if blood flow and oxygen supplies are not quickly restored.

Pulse

After blood is ejected from the


heart, elastic fibers in the arteries
help maintain a high-pressure
gradient as they expand to
accommodate the blood, then
recoil. This expansion and recoiling
effect, known as the pulse, can be
palpated manually or measured
electronically. Although the effect
diminishes over distance from the
heart, elements of the systolic and
diastolic components of the pulse
are still evident down to the level of
the arterioles.

Because pulse indicates


heart rate, it is measured clinically
to provide clues to a patient’s state
of health. It is recorded as beats
General Physiology

per minute. Both the rate and the strength of the pulse are important clinically.
A high or irregular pulse rate can be caused by physical activity or other
temporary factors, but it may also indicate a heart condition. The pulse strength
indicates the strength of ventricular contraction and cardiac output. If the pulse
is strong, then systolic pressure is high. If it is weak, systolic pressure has
fallen, and medical intervention may be warranted.

Pulse can be palpated manually by placing the tips of the fingers across an
artery that runs close to the body surface and pressing lightly. While this
procedure is normally performed using the radial artery in the wrist or the
common carotid artery in the neck, any superficial artery that can be palpated
may be used. Common sites to find a pulse include temporal and facial arteries
in the head, brachial arteries in the upper arm, femoral arteries in the thigh,
popliteal arteries behind the knees, posterior tibial arteries near the medial
tarsal regions, and dorsalis pedis arteries in the feet. A variety of commercial
electronic devices are also available to measure pulse.

Capillary Exchange
The primary purpose of the cardiovascular system is to circulate gases,
nutrients, wastes, and other substances to and from the cells of the body. Small
molecules, such as gases, lipids, and lipid-soluble molecules, can diffuse
directly through the membranes of the endothelial cells of the capillary wall.
Glucose, amino acids, and ions—including sodium, potassium, calcium, and
chloride—use transporters to move through specific channels in the membrane
by facilitated diffusion. Glucose, ions, and larger molecules may also leave the
blood through intercellular clefts. Larger molecules can pass through the pores
of fenestrated capillaries, and even large plasma proteins can pass through the
great gaps in the sinusoids. Some large proteins in blood plasma can move into
and out of the endothelial cells packaged within vesicles by endocytosis and
exocytosis. Water moves by osmosis.
Bulk Flow
The mass movement of fluids into and out of capillary beds requires a
transport mechanism far more efficient than mere diffusion. This movement,
often referred to as bulk flow, involves two pressure-driven mechanisms:
Volumes of fluid move from an area of higher pressure in a capillary bed to an
area of lower pressure in the tissues via filtration. In contrast, the movement of
fluid from an area of higher pressure in the tissues into an area of lower
pressure in the capillaries is reabsorption. Two types of pressure interact to
drive each of these movements: hydrostatic pressure and osmotic pressure.
Hydrostatic Pressure
The primary force driving fluid transport between the capillaries and
tissues is hydrostatic pressure, which can be defined as the pressure of any
fluid enclosed in a space. Blood hydrostatic pressure is the force exerted by
the blood confined within blood vessels or heart chambers. Even more
General Physiology

specifically, the pressure exerted by blood against the wall of a capillary is


called capillary hydrostatic pressure (CHP), and is the same as capillary blood
pressure. CHP is the force that drives fluid out of capillaries and into the
tissues.
As fluid exits a capillary and moves into tissues, the hydrostatic pressure in the
interstitial fluid correspondingly rises. This opposing hydrostatic pressure is
called the interstitial fluid hydrostatic pressure (IFHP). Generally, the CHP
originating from the arterial pathways is considerably higher than the IFHP,
because lymphatic vessels are continually absorbing excess fluid from the
tissues. Thus, fluid generally moves out of the capillary and into the interstitial
fluid. This process is called filtration.

Osmotic Pressure
The net pressure that drives reabsorption—the movement of fluid from the
interstitial fluid back into the capillaries—is called osmotic pressure (sometimes
referred to as oncotic pressure). Whereas hydrostatic pressure forces fluid out of
the capillary, osmotic pressure draws fluid back in. Osmotic pressure is
determined by osmotic concentration gradients, that is, the difference in the
solute-to-water concentrations in the blood and tissue fluid. A region higher in
solute concentration (and lower in water concentration) draws water across a
semipermeable membrane from a region higher in water concentration (and
lower in solute concentration).
As we discuss osmotic pressure in blood and tissue fluid, it is important
to recognize that the formed elements of blood do not contribute to osmotic
concentration gradients. Rather, it is the plasma proteins that play the key role.
Solutes also move across the capillary wall according to their concentration
gradient, but overall, the concentrations should be similar and not have a
significant impact on osmosis. Because of their large size and chemical
structure, plasma proteins are not truly solutes, that is, they do not dissolve but
are dispersed or suspended in their fluid medium, forming a colloid rather than
a solution.
The pressure created by the concentration of colloidal proteins in the
blood is called the blood colloidal osmotic pressure (BCOP). Its effect on
capillary exchange accounts for the reabsorption of water. The plasma proteins
suspended in blood cannot move across the semipermeable capillary cell
membrane, and so they remain in the plasma. As a result, blood has a higher
colloidal concentration and lower water concentration than tissue fluid. It
therefore attracts water. We can also say that the BCOP is higher than
General Physiology

the interstitial fluid colloidal osmotic pressure (IFCOP), which is always very


low because interstitial fluid contains few proteins. Thus, water is drawn from
the tissue fluid back into the capillary, carrying dissolved molecules with it. This
difference in colloidal osmotic pressure accounts for reabsorption.
Interaction of Hydrostatic and Osmotic Pressures
The normal unit used to express pressures within the cardiovascular
system is millimeters of mercury (mm Hg). When blood leaving an arteriole first
enters a capillary bed, the CHP is quite high—about 35 mm Hg. Gradually, this
initial CHP declines as the blood moves through the capillary so that by the time
the blood has reached the venous end, the CHP has dropped to approximately
18 mm Hg. In comparison, the plasma proteins remain suspended in the blood,
so the BCOP remains fairly constant at about 25 mm Hg throughout the length
of the capillary and considerably above the osmotic pressure in the interstitial
fluid.
The net filtration pressure (NFP) represents the interaction of the
hydrostatic and osmotic pressures, driving fluid out of the capillary. It is equal
to the difference between the CHP and the BCOP. Since filtration is, by
definition, the movement of fluid out of the capillary, when reabsorption is
occurring, the NFP is a negative number.
NFP changes at different points in a capillary bed. Close to the arterial
end of the capillary, it is approximately 10 mm Hg, because the CHP of 35 mm
Hg minus the BCOP of 25 mm Hg equals 10 mm Hg. Recall that the hydrostatic
and osmotic pressures of the interstitial fluid are essentially negligible. Thus,
the NFP of 10 mm Hg drives a net movement of fluid out of the capillary at the
arterial end. At approximately the middle of the capillary, the CHP is about the
same as the BCOP of 25 mm Hg, so the NFP drops to zero. At this point, there is
no net change of volume: Fluid moves out of the capillary at the same rate as it
moves into the capillary. Near the venous end of the capillary, the CHP has
dwindled to about 18 mm Hg due to loss of fluid. Because the BCOP remains
steady at 25 mm Hg, water is drawn into the capillary, that is, reabsorption
occurs. Another way of expressing this is to say that at the venous end of the
capillary, there is an NFP of −7 mm Hg.
General Physiology

The Role of Lymphatic Capillaries


Since overall CHP is higher than BCOP, it is inevitable that more net fluid
will exit the capillary through filtration at the arterial end than enters through
reabsorption at the venous end. Considering all capillaries over the course of a
day, this can be quite a substantial amount of fluid: Approximately 24 liters per
day are filtered, whereas 20.4 liters are reabsorbed. This excess fluid is picked
up by capillaries of the lymphatic system. These extremely thin-walled vessels
have copious numbers of valves that ensure unidirectional flow through ever-
larger lymphatic vessels that eventually drain into the subclavian veins in the
neck. An important function of the lymphatic system is to return the fluid
(lymph) to the blood. Lymph may be thought of as recycled blood plasma.

Circulatory Pathways
Virtually every cell, tissue, organ, and system in the body is impacted by the
circulatory system. This includes the generalized and more specialized functions of
transport of materials, capillary exchange, maintaining health by transporting white
blood cells and various immunoglobulins (antibodies), hemostasis, regulation of body
temperature, and helping to maintain acid-base balance. In addition to these shared
functions, many systems enjoy a unique relationship with the circulatory system.
As you learn about the vessels of the systemic and pulmonary circuits, notice
that many arteries and veins share the same names, parallel one another throughout
the body, and are very similar on the right and left sides of the body. These pairs of
vessels will be traced through only one side of the body. Where differences occur in
branching patterns or when vessels are singular, this will be indicated. For example,
you will find a pair of femoral arteries and a pair of femoral veins, with one vessel on
each side of the body. In contrast, some vessels closer to the midline of the body, such
as the aorta, are unique. Moreover, some superficial veins, such as the great saphenous
vein in the femoral region, have no arterial counterpart. Another phenomenon that can
make the study of vessels challenging is that names of vessels can change with
location. Like a street that changes name as it passes through an intersection, an
artery or vein can change names as it passes an anatomical landmark. For example,
the left subclavian artery becomes the axillary artery as it passes through the body wall
and into the axillary region, and then becomes the brachial artery as it flows from the
axillary region into the upper arm (or brachium). You will also find examples of
anastomoses where two blood vessels that previously branched reconnect.
Anastomoses are especially common in veins, where they help maintain blood flow even
when one vessel is blocked or narrowed, although there are some important ones in the
arteries supplying the brain.

As you read about circular pathways, notice that there is an occasional,


very large artery referred to as a trunk, a term indicating that the vessel gives
rise to several smaller arteries. For example, the celiac trunk gives rise to the
left gastric, common hepatic, and splenic arteries.
General Physiology
General Physiology
General Physiology

Pulmonary Circulation

Recall that blood returning from


the systemic circuit enters the
right atrium via the superior and
inferior venae cavae and the
coronary sinus, which drains the
blood supply of the heart muscle.
These vessels will be described
more fully later in this section.
This blood is relatively low in
oxygen and relatively high in
carbon dioxide, since much of the
oxygen has been extracted for use
by the tissues and the waste gas
carbon dioxide was picked up to be transported to the lungs for elimination.
From the right atrium, blood moves into the right ventricle, which pumps it to
the lungs for gas exchange. This system of vessels is referred to as
the pulmonary circuit.

The single vessel exiting the right ventricle is the pulmonary trunk. At the base
of the pulmonary trunk is the pulmonary semilunar valve, which prevents
backflow of blood into the right ventricle during ventricular diastole. As the
pulmonary trunk reaches the superior surface of the heart, it curves posteriorly
and rapidly bifurcates (divides) into two branches, a left and a right pulmonary
artery. To prevent confusion between these vessels, it is important to refer to
the vessel exiting the heart as the pulmonary trunk, rather than also calling it a
pulmonary artery. The pulmonary arteries in turn branch many times within
the lung, forming a series of smaller arteries and arterioles that eventually lead
to the pulmonary capillaries. The pulmonary capillaries surround lung
structures known as alveoli that are the sites of oxygen and carbon dioxide
exchange.

Once gas exchange is completed, oxygenated blood flows from the pulmonary
capillaries into a series of pulmonary venules that eventually lead to a series of
larger pulmonary veins. Four pulmonary veins, two on the left and two on the
right, return blood to the left atrium. At this point, the pulmonary circuit is
complete. 
General Physiology

Application:

“Case Analysis”

Two men have had medical operations on their circulatory systems.


One has had a pacemaker and the other has had a defibrillator installed in
their chest. One of the men said that they are the same things, and the other
says they are not but he can’t remember what each one does. Settle their
disagreement by explaining what they do.
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________

Assessment:

Read the following questions and choose the choices that


corresponds to the best answer.

1. Blood leaving the left ventricle enters the ________.


a. Pulmonary trunk b. Pulmonary artery
c. inferior vena cava d. aorta

2. The right ventricle pumps blood into the __________.


a. systemic circuit b. lungs
c. left atrium d. right atrium

3. The visceral pericardium is the same as the _________.


a. epicardium b. endocardium
c. myocardium d. parietal pericardium

4. What is blood flow?


a. The movement of blood through a vessel, tissue, or organ
b. The movement of blood through the heart
c. The movement of blood through the liver
d. The movement of blood through the brain
General Physiology

5. What initiates blood flow?


a. The contraction of the ventricles of the heart
b. The relaxation of the ventricles of the heart
c. The contraction of the atria of the heart
d. The relaxation of the atria of the heart

6. What is hydrostatic pressure?


a. The force exerted by a fluid due to gravitational pull.
b. The force exerted by blood upon the walls of the blood vessels.
c. The pressure of blood flowing in the veins.
d. The pressure of blood flowing in the lungs.

7. What does systolic pressure reflect?


a. The arterial pressure resulting from ventricular relaxation.
b. The arterial pressure resulting from ventricular contraction.
c. The arterial pressure in the veins.
d. The arterial pressure in the capillaries.

8. What is diastolic pressure?


a. The higher value of the blood pressure ratio.
b. The lower value of the blood pressure ratio.
c. The force exerted by blood upon the walls of the heart.
d. The force exerted by blood upon the walls of the lungs.

9. What is pulse pressure?


a. The pressure of blood flowing in the systemic circulation
b. The difference between the systolic pressure and the diastolic pressure
c. The average pressure of blood in the arteries
d. The force exerted by a fluid due to gravitational pull

10. What is mean arterial pressure?


a. The average pressure of blood in the veins
b. The average pressure of blood in the capillaries
c. The average force driving blood into vessels that serve the tissues
d. The pressure of blood flowing in the systemic circulation

11. What is the normal range of mean arterial pressure?


a. 60-70 mm Hg
b. 70-80 mm Hg
c. 80-90 mm Hg
d. 70-110 mm Hg

12. What is the result of low mean arterial pressure for an extended time?
a. Normal circulation to and through the tissues
b. Insufficient blood flow
c. High levels of oxygen in systemic arterial blood
d. Adequate oxygenation of tissues
General Physiology

13. What does pulse indicate?


a. The strength of ventricular contraction and cardiac output
b. The state of health of a patient
c. The rate of physical activity
d. The amount of blood in the body.

14. What is the difference between systolic and diastolic pressures?


a. Systolic pressure is the pressure of blood in the arteries while diastolic
pressure is the pressure of blood in the veins.
b. Systolic pressure is the pressure of blood during ventricular contraction while
diastolic pressure is the pressure of blood during ventricular relaxation.
c. Systolic pressure is the average pressure of blood in the arteries while
diastolic pressure is the pulse pressure.
d. Systolic pressure is the pressure of blood in the capillaries while diastolic
pressure is the pressure of blood in the venules and veins.

15. What is mean arterial pressure (MAP)?


a. The pressure of blood in the arteries during ventricular relaxation.
b. The average pressure of blood in the veins.
c. The difference between the systolic and diastolic pressures.
d. The average force driving blood into vessels that serve the tissues.
General Physiology

Lesson 3
Energy, Maintenance, and Environmental Exchange

Objectives:
At the end of the lesson the students shall be able to:

 Discuss six fundamental activities of the digestive


system, giving an example of each;
 Compare and contrast the neural and hormonal controls
involved in digestion;
 Identify the locations and primary secretions involved in the
chemical digestion of carbohydrates, proteins, lipids, and
nucleic acids; and
 Compare and contrast absorption of the hydrophilic and
hydrophobic nutrients.

Activity:
“Build the Flow”
Create your own flowchart or diagram that shows the journey of food
through the digestive system, highlighting the different organs and
their functions.
General Physiology

Analysis: From the activity, what were the main steps in the digestive
process that you simulated in the activity?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________
How did the different parts of the digestive system work together
to digest the food you simulated in the activity?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________
Which parts of the digestive system do you think are the most
important and why?
_______________________________________________________________
_______________________________________________________________
_______________________________________________________________

Abstraction:

The Digestive System

The digestive system uses mechanical and chemical activities to


break food down into absorbable substances during its journey
through the digestive system.

Digestive Process

The processes of digestion include six activities: ingestion, propulsion,


mechanical or physical digestion, chemical digestion, absorption, and
defecation.
General Physiology
General Physiology

The first of these processes, ingestion, refers to the entry of food into the
alimentary canal through the mouth. There, the food is chewed and mixed with
saliva, which contains enzymes that begin breaking down the carbohydrates in
the food plus some lipid digestion via lingual lipase. Chewing increases the
surface area of the food and allows an appropriately sized bolus to be produced.

Food leaves the mouth when


the tongue and pharyngeal muscles
propel it into the esophagus. This act
of swallowing, the last voluntary act
until defecation, is an example
of propulsion, which refers to the
movement of food through the
digestive tract. It includes both the
voluntary process of swallowing and
the involuntary process of
peristalsis. Peristalsis consists of
sequential, alternating waves of
contraction and relaxation of
alimentary wall smooth muscles,
which act to propel food. These waves also play a role in mixing food with
digestive juices. Peristalsis is so powerful that foods and liquids you swallow
enter your stomach even if you are standing on your head.

Digestion includes both mechanical and chemical processes. Mechanical


digestion is a purely physical process that does not change the chemical nature
of the food. Instead, it makes the food smaller to increase both surface area and
mobility. It includes mastication, or chewing, as well as tongue movements that
help break food into smaller bits and mix food with saliva. Although there may
be a tendency to think that mechanical digestion is limited to the first steps of
the digestive process, it occurs after the food leaves the mouth, as well. The
mechanical churning of food in the stomach serves to further break it apart and
expose more of its surface area to digestive juices, creating an acidic “soup”
called chyme. Segmentation, which occurs mainly in the small intestine,
consists of localized contractions of circular muscle of the muscularis layer of
the alimentary canal. These contractions isolate small sections of the intestine,
moving their contents back and forth while continuously subdividing, breaking
up, and mixing the contents. By moving food back and forth in the intestinal
lumen, segmentation mixes food with digestive juices and facilitates absorption.

In chemical digestion, starting in the mouth, digestive secretions break


down complex food molecules into their chemical building blocks (for example,
proteins into separate amino acids). These secretions vary in composition, but
typically contain water, various enzymes, acids, and salts. The process is
completed in the small intestine.
General Physiology

Food that has been broken down is of no value to the body unless it
enters the bloodstream, and its nutrients are put to work. This occurs through
the process of absorption, which takes place primarily within the small
intestine. There, most nutrients are absorbed from the lumen of the alimentary
canal into the bloodstream through the epithelial cells that make up the
mucosa. Lipids are absorbed into
lacteals and are transported via the
lymphatic vessels to the bloodstream
(the subclavian veins near the heart).
The details of these processes will be
discussed later.

In defecation, the final step in


digestion, undigested materials are
removed from the body as feces.

In some cases, a single organ is


in charge of a digestive process. For
example, ingestion occurs only in the
mouth and defecation only in the
anus. However, most digestive
processes involve the interaction of
several organs and occur gradually as
food moves through the alimentary canal.

Some chemical digestion occurs in the mouth. Some absorption can occur
in the mouth and stomach, for example, alcohol and aspirin.

Regulatory Mechanisms

Neural and endocrine regulatory mechanisms work to maintain the


optimal conditions in the lumen needed for digestion and absorption. These
regulatory mechanisms, which stimulate digestive activity through mechanical
and chemical activity, are controlled both extrinsically and intrinsically.

Neural Controls

The walls of the alimentary canal contain a variety of sensors that help
regulate digestive functions. These include mechanoreceptors, chemoreceptors,
and osmoreceptors, which are capable of detecting mechanical, chemical, and
osmotic stimuli, respectively. For example, these receptors can sense when the
presence of food has caused the stomach to expand, whether food particles have
been sufficiently broken down, how much liquid is present, and the type of
nutrients in the food (lipids, carbohydrates, and/or proteins). Stimulation of
these receptors provokes an appropriate reflex that furthers the process of
digestion. This may entail sending a message that activates the glands that
secrete digestive juices into the lumen, or it may mean the stimulation of
muscles within the alimentary canal, thereby activating peristalsis and
segmentation that move food along the intestinal tract.
General Physiology

The walls of the entire alimentary canal are embedded with nerve
plexuses that interact with the central nervous system and other nerve plexuses
—either within the same digestive organ or in different ones. These interactions
prompt several types of reflexes. Extrinsic nerve plexuses orchestrate long
reflexes, which involve the central and autonomic nervous systems and work in
response to stimuli from outside the digestive system. Short reflexes, on the
other hand, are orchestrated by intrinsic nerve plexuses within the alimentary
canal wall. These two plexuses and their connections were introduced earlier as
the enteric nervous system. Short reflexes regulate activities in one area of the
digestive tract and may coordinate local peristaltic movements and stimulate
digestive secretions. For example, the sight, smell, and taste of food initiate long
reflexes that begin with a sensory neuron delivering a signal to the medulla
oblongata. The response to the signal is to stimulate cells in the stomach to
begin secreting digestive juices in preparation for incoming food. In contrast,
food that distends the stomach initiates short reflexes that cause cells in the
stomach wall to increase their secretion of digestive juices.

Hormonal Controls

A variety of hormones are involved in the digestive process. The main


digestive hormone of the stomach is gastrin, which is secreted in response to
the presence of food. Gastrin stimulates the secretion of gastric acid by the
parietal cells of the stomach mucosa. Other GI hormones are produced and act
upon the gut and its accessory organs. Hormones produced by the duodenum
include secretin, which stimulates a watery secretion of bicarbonate by the
pancreas; cholecystokinin (CCK), which stimulates the secretion of pancreatic
enzymes and bile from the liver and release of bile from the gallbladder; and
gastric inhibitory peptide, which inhibits gastric secretion and slows gastric
emptying and motility. These GI hormones are secreted by specialized epithelial
cells, called endocrinocytes, located in the mucosal epithelium of the stomach
and small intestine. These hormones then enter the bloodstream, through which
they can reach their target organs.

Carbohydrate Digestion

The average American diet is about 50 percent carbohydrates, which may


be classified according to the number of monomers they contain of simple
sugars (monosaccharides and disaccharides) and/or complex sugars
(polysaccharides). Glucose, galactose, and fructose are the three
monosaccharides that are commonly consumed and are readily absorbed. Your
digestive system is also able to break down the disaccharide sucrose (regular
table sugar: glucose + fructose), lactose (milk sugar: glucose + galactose), and
maltose (grain sugar: glucose + glucose), and the polysaccharides glycogen and
starch (chains of monosaccharides). Your bodies do not produce enzymes that
can break down most fibrous polysaccharides, such as cellulose. While
indigestible polysaccharides do not provide any nutritional value, they do
provide dietary fiber, which helps propel food through the alimentary canal.
General Physiology

In the small intestine, pancreatic amylase does the ‘heavy lifting’ for


starch and carbohydrate digestion. After amylases break down starch into
smaller fragments, the brush border enzyme α-dextrinase starts working on α-
dextrin, breaking off one glucose unit at a time. Three brush border enzymes
hydrolyze sucrose, lactose, and maltose into monosaccharides. Sucrase splits
sucrose into one molecule of fructose and one molecule of
glucose; maltase breaks down maltose and maltotriose into two and three
glucose molecules, respectively; and lactase breaks down lactose into one
molecule of glucose and one molecule of galactose. Insufficient lactase can lead
to lactose

intolerance.

Protein Digestion
General Physiology

Proteins are polymers


composed of amino acids linked by
peptide bonds to form long chains.
Digestion reduces them to their
constituent amino acids. You usually
consume about 15 to 20 percent of
your total calorie intake as protein.
The digestion of protein starts in the
stomach, where HCl and pepsin
break proteins into smaller
polypeptides, which then travel to the
small intestine. Chemical digestion in
the small intestine is continued by
pancreatic enzymes, including
chymotrypsin and trypsin, each of
which act on specific bonds in amino
acid sequences. At the same time,
the cells of the brush border secrete enzymes such
as aminopeptidase and dipeptidase, which further break down peptide chains.
This results in molecules small enough to enter the bloodstream.
Lipid Digestion
A healthy diet limits lipid intake to 35 percent of total calorie intake. The
most common dietary lipids are triglycerides, which are made up of a glycerol
molecule bound to three fatty acid chains. Small amounts of dietary cholesterol
and phospholipids are also consumed.
The three lipases responsible for lipid digestion are lingual lipase, gastric lipase,
and pancreatic lipase. However, because the pancreas is the only
consequential source of lipase, virtually all lipid digestion occurs in the small
intestine. Pancreatic lipase breaks down each triglyceride into two free fatty
acids and a monoglyceride. The fatty acids include both short-chain (less than
10 to 12 carbons) and long-chain fatty acids.
Nucleic Acid Digestion
The nucleic acids DNA and RNA are found in most of the foods you eat.
Two types of pancreatic nuclease are responsible for their
digestion: deoxyribonuclease, which digests DNA, and ribonuclease, which
digests RNA. The nucleotides produced by this digestion are further broken
down by two intestinal brush border enzymes (nucleosidase and phosphatase)
into pentoses, phosphates, and nitrogenous bases, which can be absorbed
through the alimentary canal wall. The large food molecules that must be
broken down into subunits.
Absorption
General Physiology

The mechanical and


digestive processes have one
goal: to convert food into
molecules small enough to be
absorbed by the epithelial
cells of the intestinal villi.
The absorptive capacity of
the alimentary canal is
almost endless. Each day,
the alimentary canal
processes up to 10 liters of
food, liquids, and GI
secretions, yet less than one
liter enters the large
intestine. Almost all ingested
food, 80 percent of
electrolytes, and 90 percent
of water are absorbed in the
small intestine. Although the
entire small intestine is
involved in the absorption of
water and lipids, most
absorption of carbohydrates
and proteins occurs in the
jejunum. Notably, bile salts and vitamin B 12 are absorbed in the terminal ileum.
By the time chyme passes from the ileum into the large intestine, it is
essentially indigestible food residue (mainly plant fibers like cellulose), some
water, and millions of bacteria.
Absorption can occur through five mechanisms: (1) active transport, (2)
passive diffusion, (3) facilitated diffusion, (4) co-transport (or secondary active
transport), and (5) endocytosis. As you will recall from Chapter 3, active
transport refers to the movement of a substance across a cell membrane going
from an area of lower concentration to an area of higher concentration (up the
concentration gradient). In this type of transport, proteins within the cell
membrane act as “pumps,” using cellular energy (ATP) to move the substance.
Passive diffusion refers to the movement of substances from an area of higher
concentration to an area of lower concentration, while facilitated diffusion refers
to the movement of substances from an area of higher to an area of lower
concentration using a carrier protein in the cell membrane. Co-transport uses
the movement of one molecule through the membrane from higher to lower
concentration to power the movement of another from lower to higher. Finally,
endocytosis is a transportation process in which the cell membrane engulfs
material. It requires energy, generally in the form of ATP.
General Physiology

Because the cell’s plasma membrane is made up of hydrophobic


phospholipids, water-soluble nutrients must use transport molecules embedded
in the membrane to enter cells. Moreover, substances cannot pass between the
epithelial cells of the intestinal mucosa because these cells are bound together
by tight junctions. Thus, substances can only enter blood capillaries by passing
through the apical surfaces of epithelial cells and into the interstitial fluid.
Water-soluble nutrients enter the capillary blood in the villi and travel to the
liver via the hepatic portal vein.
In contrast to the water-soluble nutrients, lipid-soluble nutrients can diffuse
through the plasma membrane. Once inside the cell, they are packaged for
transport via the base of the cell and then enter the lacteals of the villi to be
transported by lymphatic vessels to the systemic circulation via the thoracic
duct. The absorption of most nutrients through the mucosa of the intestinal villi
requires active transport fueled by ATP.
Carbohydrate Absorption
All carbohydrates are absorbed in the form of monosaccharides. The small
intestine is highly efficient at this, absorbing monosaccharides at an estimated
rate of 120 grams per hour. All normally digested dietary carbohydrates are
absorbed; indigestible fibers are eliminated in the feces. The monosaccharides
glucose and galactose are transported into the epithelial cells by common
protein carriers via secondary active transport (that is, co-transport with sodium
ions). The monosaccharides leave these cells via facilitated diffusion and enter
the capillaries through intercellular clefts. The monosaccharide fructose (which
is in fruit) is absorbed and transported by facilitated diffusion alone. The
monosaccharides combine with the transport proteins immediately after the
disaccharides are broken down.

Protein Absorption

Active transport mechanisms, primarily in the duodenum and jejunum,


absorb most proteins as their breakdown products, amino acids. Almost all (95
to 98 percent) protein is digested and absorbed in the small intestine. The type
of carrier that transports an amino acid varies. Most carriers are linked to the
active transport of sodium. Short chains of two amino acids (dipeptides) or three
amino acids (tripeptides) are also transported actively. However, after they enter
the absorptive epithelial cells, they are broken down into their amino acids
before leaving the cell and entering the capillary blood via diffusion.

Lipid Absorption

About 95 percent of lipids are absorbed in the small intestine. Bile salts
not only speed up lipid digestion, they are also essential to the absorption of the
end products of lipid digestion. Short-chain fatty acids are relatively water
soluble and can enter the absorptive cells (enterocytes) directly. The small size
of short-chain fatty acids enables them to be absorbed by enterocytes via simple
General Physiology

diffusion, and then take the same path as monosaccharides and amino acids
into the blood capillary of a villus.

The large and hydrophobic long-chain fatty acids and monoacylglycerides


are not so easily suspended in the watery intestinal chyme. However, bile salts
and lecithin resolve this issue by enclosing them in a micelle, which is a tiny
sphere with polar (hydrophilic) ends facing the watery environment and
hydrophobic tails turned to the interior, creating a receptive environment for the
long-chain fatty acids. The core also includes cholesterol and fat-soluble
vitamins. Without micelles, lipids would sit on the surface of chyme and never
come in contact with the absorptive surfaces of the epithelial cells. Micelles can
easily squeeze between microvilli and get very near the luminal cell surface. At
this point, lipid substances exit the micelle and are absorbed via simple
diffusion.

The free fatty acids and monoacylglycerides that enter the epithelial cells
are reincorporated into triglycerides. The triglycerides are mixed with
phospholipids and cholesterol and surrounded with a protein coat. This new
complex, called a chylomicron, is a water-soluble lipoprotein. After being
processed by the Golgi apparatus, chylomicrons are released from the cell. Too
big to pass through the basement membranes of blood capillaries, chylomicrons
instead enter the large pores of lacteals. The lacteals come together to form the
lymphatic vessels. The chylomicrons are transported in the lymphatic vessels
and empty through the thoracic duct into the subclavian vein of the circulatory
system. Once in the bloodstream, the enzyme lipoprotein lipase breaks down
the triglycerides of the chylomicrons into free fatty acids and glycerol. These
breakdown products then pass through capillary walls to be used for energy by
cells or stored in adipose tissue as fat. Liver cells combine the remaining
chylomicron remnants with proteins, forming lipoproteins that transport
cholesterol in the blood.
General Physiology

Nucleic Acid Absorption

The products of nucleic acid digestion—pentose sugars, nitrogenous


bases, and phosphate ions—are transported by carriers across the villus
epithelium via active transport. These products then enter the bloodstream.

Mineral Absorption

The electrolytes absorbed by the small intestine are from both GI


secretions and ingested foods. Since electrolytes dissociate into ions in water,
most are absorbed via active transport throughout the entire small intestine.
General Physiology

During absorption, co-transport mechanisms result in the accumulation of


sodium ions inside the cells, whereas anti-port mechanisms reduce the
potassium ion concentration inside the cells. To restore the sodium-potassium
gradient across the cell membrane, a sodium-potassium pump requiring ATP
pumps sodium out and potassium in.

In general, all minerals that enter the intestine are absorbed, whether you
need them or not. Iron and calcium are exceptions; they are absorbed in the
duodenum in amounts that meet the body’s current requirements, as follows:

 Iron—The ionic iron needed for the production of hemoglobin is absorbed


into mucosal cells via active transport. Once inside mucosal cells, ionic
iron binds to the protein ferritin, creating iron-ferritin complexes that
store iron until needed. When the body has enough iron, most of the
stored iron is lost when worn-out epithelial cells slough off. When the
body needs iron because, for example, it is lost during acute or chronic
bleeding, there is increased uptake of iron from the intestine and
accelerated release of iron into the bloodstream. Since women experience
significant iron loss during menstruation, they have around four times as
many iron transport proteins in their intestinal epithelial cells as do men.
 Calcium—Blood levels of ionic calcium determine the absorption of dietary
calcium. When blood levels of ionic calcium drop, parathyroid hormone
(PTH) secreted by the parathyroid glands stimulates the release of calcium
ions from bone matrices and increases the reabsorption of calcium by the
kidneys. PTH also upregulates the activation of vitamin D in the kidney,
which then facilitates intestinal calcium ion absorption.

Vitamin Absorption

The small intestine absorbs the vitamins that occur naturally in food and
supplements. Fat-soluble vitamins (A, D, E, and K) are absorbed along with
dietary lipids in micelles via simple diffusion. This is why you are advised to eat
some fatty foods when you take fat-soluble vitamin supplements. Most water-
soluble vitamins (including most B vitamins and vitamin C) also are absorbed
by simple diffusion. An exception is vitamin B 12, which is a very large molecule.
Intrinsic factor secreted in the stomach binds to vitamin B 12, preventing its
digestion and creating a complex that binds to mucosal receptors in the
terminal ileum, where it is taken up by endocytosis.

Water Absorption

Each day, about nine liters of fluid enter the small intestine. About 2.3
liters are ingested in foods and beverages, and the rest is from GI secretions.
About 90 percent of this water is absorbed in the small intestine. Water
absorption is driven by the concentration gradient of the water: The
concentration of water is higher in chyme than it is in epithelial cells. Thus,
water moves down its concentration gradient from the chyme into cells. As
noted earlier, much of the remaining water is then absorbed in the colon.
General Physiology

Application:

“Healthy Meal Plan”

Design a healthy meal plan: Based on what you have learned about the
digestive system, create a healthy meal plan for one week. Include a variety of
foods that are high in fiber, protein, and other essential nutrients, and
explain why each food is important for a healthy digestive system.
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________
____________________________________________________________________________

Assessment:

Read the following questions and choose the choices that


corresponds to the best answer.
1. What is the first process in digestion?
a) Propulsion
b) Absorption
c) Ingestion
d) Chemical digestion

2. What does the act of swallowing represent in the digestive process?


a) Chemical digestion
b) Ingestion
c) Propulsion
d) Absorption

3. What is the purpose of mechanical digestion?


a) To change the chemical nature of food
b) To make food smaller to increase surface area and mobility
General Physiology

c) To mix food with digestive juices


d) To remove undigested materials from the body

4. Which process involves breaking down complex food molecules into their
chemical building blocks?
a) Mechanical digestion
b) Ingestion
c) Chemical digestion
d) Defecation

5. Where does most absorption take place in the digestive system?


a) Mouth and stomach
b) Small intestine
c) Large intestine
d) Esophagus

6. What is the final step in digestion?


a) Absorption
b) Propulsion
c) Chemical digestion
d) Defecation

7. What is the main function of segmentation in the small intestine?


a) Breaking down complex food molecules
b) Mixing food with digestive juices and facilitating absorption
c) Removing undigested materials from the body
d) Increasing the surface area of the food

8. Which of the following is NOT an example of mechanical digestion?


a) Chewing
b) Tongue movements
c) Absorption of alcohol
d) The mechanical churning of food in the stomach

9. What is the acidic “soup” called that is formed in the stomach after the
mechanical churning of food?
a) Chyme
b) Lacteals
c) Subclavian veins
d) Feces

10. What happens in the process of defecation?


a) Undigested materials are removed from the body as feces
b) Food is propelled through the digestive tract
c) Food is broken down into its chemical building blocks
d) Nutrients are absorbed into the bloodstream.
General Physiology

Male Reproductive System


Unique for its role in human reproduction, a gamete is a specialized sex
cell carrying 23 chromosomes—one half the number in body cells. At
fertilization, the chromosomes in one male gamete, called a sperm (or
spermatozoon), combine with the chromosomes in one female gamete, called an
oocyte. The function of the male reproductive system is to produce sperm and
transfer them to the female reproductive tract. The paired testes are a crucial
component in this process, as they produce both sperm and androgens, the
hormones that support male reproductive physiology. In humans, the most
important male androgen is testosterone. Several accessory organs and ducts
aid the process of sperm maturation and transport the sperm and other seminal
components to the penis, which delivers sperm to the female reproductive tract.
In this section, we examine each of these different structures, and discuss the
process of sperm production and transport.
General Physiology

Scrotum
The testes are located in a skin-covered, highly pigmented, muscular sack
called the scrotum that extends from the body behind the penis. This location
is important in sperm production, which occurs within the testes, and proceeds
more efficiently when the testes are kept 2 to 4°C below core body temperature.
The dartos muscle makes up the subcutaneous muscle layer of the
scrotum. It continues internally to make up the scrotal septum, a wall that
divides the scrotum into two compartments, each housing one testis.
Descending from the internal oblique muscle of the abdominal wall are the two
cremaster muscles, which cover each testis like a muscular net. By contracting
simultaneously, the dartos and cremaster muscles can elevate the testes in cold
weather (or water), moving the testes closer to the body and decreasing the
surface area of the scrotum to retain heat. Alternatively, as the environmental
General Physiology

temperature increases, the scrotum relaxes, moving the testes farther from the
body core and increasing scrotal surface area, which promotes heat loss.
Externally, the scrotum has a raised medial thickening on the surface called the
raphae.

Germ Cells
The least mature cells, the spermatogonia (singular = spermatogonium),
line the basement membrane inside the tubule. Spermatogonia are the stem
cells of the testis, which means that they are still able to differentiate into a
variety of different cell types throughout adulthood. Spermatogonia divide to
produce primary and secondary spermatocytes, then spermatids, which finally
produce formed sperm. The process that begins with spermatogonia and
concludes with the production of sperm is called spermatogenesis.
Spermatogenesis
As just noted, spermatogenesis occurs in the seminiferous tubules that
form the bulk of each testis. The process begins at puberty, after which time
sperm are produced constantly throughout a man’s life. One production cycle,
from spermatogonia through formed sperm, takes approximately 64 days. A new
cycle starts approximately every 16 days, although this timing is not
synchronous across the seminiferous tubules. Sperm counts—the total number
of sperm a man produces—slowly decline after age 35, and some studies
suggest that smoking can lower sperm counts irrespective of age.
The process of spermatogenesis begins with mitosis of the diploid
spermatogonia. Because these cells are diploid (2n), they each have a complete
copy of the father’s genetic material, or 46 chromosomes. However, mature
gametes are haploid (1n), containing 23 chromosomes—meaning that daughter
cells of spermatogonia must undergo a second cellular division through the
process of meiosis.
General Physiology

Two identical diploid cells result from spermatogonia mitosis. One of these
cells remains a spermatogonium, and the other becomes a
primary spermatocyte, the next stage in the process of spermatogenesis. As in
mitosis, DNA is replicated in a primary spermatocyte, before it undergoes a cell
division called meiosis I. During meiosis I each of the 23 pairs of chromosomes
separates. This results in two cells, called secondary spermatocytes, each with
only half the number of chromosomes. Now a second round of cell division
(meiosis II) occurs in both of the secondary spermatocytes. During meiosis II
each of the 23 replicated chromosomes divides, similar to what happens during
mitosis. Thus, meiosis results in separating the chromosome pairs. This second
meiotic division results in a total of four cells with only half of the number of
chromosomes. Each of these new cells is a spermatid. Although haploid, early
spermatids look very similar to cells in the earlier stages of spermatogenesis,
with a round shape, central nucleus, and large amount of cytoplasm. A process
called spermiogenesis transforms these early spermatids, reducing the
cytoplasm, and beginning the formation of the parts of a true sperm. The fifth
stage of germ cell formation—spermatozoa, or formed sperm—is the end result
of this process, which occurs in the portion of the tubule nearest the lumen.
Eventually, the sperm are released into the lumen and are moved along a series
of ducts in the testis toward a structure called the epididymis for the next step
of sperm maturation.
Sperm Transport
General Physiology

To fertilize an egg, sperm must be moved from the seminiferous tubules in


the testes, through the epididymis, and—later during ejaculation—along the
length of the penis and out into the female reproductive tract.
Role of the Epididymis
From the lumen of the seminiferous tubules, the immotile sperm are
surrounded by testicular fluid and moved to the epididymis (plural =
epididymides), a coiled tube attached to the testis where newly formed sperm
continue to mature. Though the epididymis does not take up much room in its
tightly coiled state, it would be approximately 6 m (20 feet) long if straightened.
It takes an average of 12 days for sperm to move through the coils of the
epididymis, with the shortest recorded transit time in humans being one day.
Sperm enter the head of the epididymis and are moved along predominantly by
the contraction of smooth muscles lining the epididymal tubes. As they are
moved along the length of the epididymis, the sperm further mature and acquire
the ability to move under their own power. Once inside the female reproductive
tract, they will use this ability to move independently toward the unfertilized
egg. The more mature sperm are then stored in the tail of the epididymis (the
final section) until ejaculation occurs.
Duct System
During ejaculation, sperm exit the tail of the epididymis and are pushed
by smooth muscle contraction to the ductus deferens (also called the vas
deferens). The ductus deferens is a thick, muscular tube that is bundled
together inside the scrotum with connective tissue, blood vessels, and nerves
into a structure called the spermatic cord. Because the ductus deferens is
physically accessible within the scrotum, surgical sterilization to interrupt
sperm delivery can be performed by cutting and sealing a small section of the
ductus (vas) deferens. This procedure is called a vasectomy, and it is an effective
form of male birth control. Although it may be possible to reverse a vasectomy,
clinicians consider the procedure permanent, and advise men to undergo it only
if they are certain they no longer wish to father children.
From each epididymis, each ductus deferens extends superiorly into the
abdominal cavity through the inguinal canal in the abdominal wall. From here,
the ductus deferens continues posteriorly to the pelvic cavity, ending posterior
to the bladder where it dilates in a region called the ampulla (meaning “flask”).

Sperm make up only 5 percent of the final volume of semen, the thick,
milky fluid that the male ejaculates. The bulk of semen is produced by three
critical accessory glands of the male reproductive system: the seminal vesicles,
the prostate, and the bulbourethral glands.
Seminal Vesicles
General Physiology

As sperm pass through the ampulla of the ductus deferens at ejaculation,


they mix with fluid from the associated seminal vesicle. The paired seminal
vesicles are glands that contribute approximately 60 percent of the semen
volume. Seminal vesicle fluid contains large amounts of fructose, which is used
by the sperm mitochondria to generate ATP to allow movement through the
female reproductive tract.
The fluid, now containing both sperm and seminal vesicle secretions, next
moves into the associated ejaculatory duct, a short structure formed from the
ampulla of the ductus deferens and the duct of the seminal vesicle. The paired
ejaculatory ducts transport the seminal fluid into the next structure, the
prostate gland.
Prostate Gland
The centrally located prostate gland sits anterior to the rectum at the
base of the bladder surrounding the prostatic urethra (the portion of the urethra
that runs within the prostate). About the size of a walnut, the prostate is formed
of both muscular and glandular tissues. It excretes an alkaline, milky fluid to
the passing seminal fluid—now called semen—that is critical to first coagulate
and then decoagulate the semen following ejaculation. The temporary
thickening of semen helps retain it within the female reproductive tract,
providing time for sperm to utilize the fructose provided by seminal vesicle
secretions. When the semen regains its fluid state, sperm can then pass farther
into the female reproductive tract.
Bulbourethral Glands
The final addition to semen is made by two bulbourethral glands (or
Cowper’s glands) that release a thick, salty fluid that lubricates the end of the
urethra and the vagina and helps to clean urine residues from the penile
urethra. The fluid from these accessory glands is released after the male
becomes sexually aroused, and shortly before the release of the semen. It is
therefore sometimes called pre-ejaculate. It is important to note that, in addition
to the lubricating proteins, it is possible for bulbourethral fluid to pick up sperm
already present in the urethra, and therefore it may be able to cause pregnancy.

The Penis
The penis is the male organ of copulation (sexual intercourse). It is flaccid
for non-sexual actions, such as urination, and turgid and rod-like with sexual
arousal. When erect, the stiffness of the organ allows it to penetrate into the
vagina and deposit semen into the female reproductive tract.
General Physiology

The shaft of the penis surrounds the urethra. The shaft is composed of
three column-like chambers of erectile tissue that span the length of the shaft.
Each of the two larger lateral chambers is called a corpus cavernosum (plural =
corpora cavernosa). Together, these make up the bulk of the penis. The corpus
spongiosum, which can be felt as a raised ridge on the erect penis, is a smaller
chamber that surrounds the spongy, or penile, urethra. The end of the penis,
called the glans penis, has a high concentration of nerve endings, resulting in
very sensitive skin that influences the likelihood of ejaculation. The skin from
the shaft extends down over the glans and forms a collar called the prepuce (or
foreskin). The foreskin also contains a dense concentration of nerve endings,
and both lubricate and protect the sensitive skin of the glans penis. A surgical
procedure called circumcision, often performed for religious or social reasons,
removes the prepuce, typically within days of birth.
Both sexual arousal and REM sleep (during which dreaming occurs) can
induce an erection. Penile erections are the result of vasocongestion, or
engorgement of the tissues because of more arterial blood flowing into the penis
than is leaving in the veins. During sexual arousal, nitric oxide (NO) is released
from nerve endings near blood vessels within the corpora cavernosa and
spongiosum. Release of NO activates a signaling pathway that results in
relaxation of the smooth muscles that surround the penile arteries, causing
General Physiology

them to dilate. This dilation increases the amount of blood that can enter the
penis and induces the endothelial cells in the penile arterial walls to also secrete
NO and perpetuate the vasodilation. The rapid increase in blood volume fills the
erectile chambers, and the increased pressure of the filled chambers compresses
the thin-walled penile venules, preventing venous drainage of the penis. The
result of this increased blood flow to the penis and reduced blood return from
the penis is erection. Depending on the flaccid dimensions of a penis, it can
increase in size slightly or greatly during erection, with the average length of an
erect penis measuring approximately 15 cm.
Testosterone
Testosterone, an androgen, is a steroid hormone produced by Leydig
cells. The alternate term for Leydig cells, interstitial cells, reflects their location
between the seminiferous tubules in the testes. In male embryos, testosterone is
secreted by Leydig cells by the seventh week of development, with peak
concentrations reached in the second trimester. This early release of
testosterone results in the anatomical differentiation of the male sexual organs.
In childhood, testosterone concentrations are low. They increase during
puberty, activating characteristic physical changes and initiating
spermatogenesis.
Functions of Testosterone
The continued presence of testosterone is necessary to keep the male
reproductive system working properly, and Leydig cells produce approximately 6
to 7 mg of testosterone per day. Testicular steroidogenesis (the manufacture of
androgens, including testosterone) results in testosterone concentrations that
are 100 times higher in the testes than in the circulation. Maintaining these
normal concentrations of testosterone promotes spermatogenesis, whereas low
levels of testosterone can lead to infertility. In addition to intratesticular
secretion, testosterone is also released into the systemic circulation and plays
an important role in muscle development, bone growth, the development of
secondary sex characteristics, and maintaining libido (sex drive) in both males
and females. In females, the ovaries secrete small amounts of testosterone,
although most is converted to estradiol. A small amount of testosterone is also
secreted by the adrenal glands in both sexes.
Control of Testosterone
The regulation of testosterone concentrations throughout the body is
critical for male reproductive function.
General Physiology

The regulation of Leydig cell production of testosterone begins outside of the


testes. The hypothalamus and the pituitary gland in the brain integrate external
and internal signals to control testosterone synthesis and secretion. The
regulation begins in the hypothalamus. Pulsatile release of a hormone
called gonadotropin-releasing hormone (GnRH) from the hypothalamus
stimulates the endocrine release of hormones from the pituitary gland. Binding
of GnRH to its receptors on the anterior pituitary gland stimulates release of the
two gonadotropins: luteinizing hormone (LH) and follicle-stimulating hormone
(FSH). These two hormones are critical for reproductive function in both men
and women. In men, FSH binds predominantly to the Sertoli cells within the
seminiferous tubules to promote spermatogenesis. FSH also stimulates the
Sertoli cells to produce hormones called inhibins, which function to inhibit FSH
release from the pituitary, thus reducing testosterone secretion. These
polypeptide hormones correlate directly with Sertoli cell function and sperm
number; inhibin B can be used as a marker of spermatogenic activity. In men,
LH binds to receptors on Leydig cells in the testes and upregulates the
production of testosterone.
A negative feedback loop predominantly controls the synthesis and secretion of
both FSH and LH. Low blood concentrations of testosterone stimulate the
hypothalamic release of GnRH. GnRH then stimulates the anterior pituitary to
secrete LH into the bloodstream. In the testis, LH binds to LH receptors on
Leydig cells and stimulates the release of testosterone. When concentrations of
testosterone in the blood reach a critical threshold, testosterone itself will bind
to androgen receptors on both the hypothalamus and the anterior pituitary,
inhibiting the synthesis and secretion of GnRH and LH, respectively. When the
blood concentrations of testosterone once again decline, testosterone no longer
General Physiology

interacts with the receptors to the same degree and GnRH and LH are once
again secreted, stimulating more testosterone production. This same process
occurs with FSH and inhibin to control spermatogenesis.
Female Reproductive System
The female
reproductive system
functions to produce
gametes and reproductive
hormones, just like the
male reproductive
system; however, it also
has the additional task of
supporting the
developing fetus and
delivering it to the
outside world. Unlike its
male counterpart, the
female reproductive
system is located
primarily inside the
pelvic cavity. The gamete
they produce is called
an oocyte.
The Ovarian Cycle
The ovarian
cycle is a set of
predictable changes in a
female’s oocytes and
ovarian follicles. During a
woman’s reproductive
years, it is a roughly 28-
day cycle that can be correlated with, but is not the same as, the menstrual
cycle (discussed shortly). The cycle includes two interrelated processes:
oogenesis (the production of female gametes) and folliculogenesis (the growth
and development of ovarian follicles).
Oogenesis
Gametogenesis in females is called oogenesis. The process begins with the
ovarian stem cells, or oogonia. Oogonia are formed during fetal development,
and divide via mitosis, much like spermatogonia in the testis. Unlike
spermatogonia, however, oogonia form primary oocytes in the fetal ovary prior to
birth. These primary oocytes are then arrested in this stage of meiosis I, only to
resume it years later, beginning at puberty and continuing until the woman is
near menopause (the cessation of a woman’s reproductive functions). The
General Physiology

number of primary oocytes present in the ovaries declines from one to two
million in an infant, to approximately 400,000 at puberty, to zero by the end of
menopause.
The initiation of ovulation—the release of an oocyte from the ovary—
marks the transition from puberty into reproductive maturity for women. From
then on, throughout a woman’s reproductive years, ovulation occurs
approximately once every 28 days. Just prior to ovulation, a surge of luteinizing
hormone triggers the resumption of meiosis in a primary oocyte. This initiates
the transition from primary to secondary oocyte. However, as you can see in,
this cell division does not result in two identical cells. Instead, the cytoplasm is
divided unequally, and one daughter cell is much larger than the other. This
larger cell, the secondary oocyte, eventually leaves the ovary during ovulation.
The smaller cell, called the first polar body, may or may not complete meiosis
and produce second polar bodies; in either case, it eventually disintegrates.
Therefore, even though oogenesis produces up to four cells, only one survives.

How does the diploid secondary oocyte become an ovum—the haploid


female gamete? Meiosis of a secondary oocyte is completed only if a sperm
succeeds in penetrating its barriers. Meiosis II then resumes, producing one
haploid ovum that, at the instant of fertilization by a (haploid) sperm, becomes
the first diploid cell of the new offspring (a zygote). Thus, the ovum can be
thought of as a brief, transitional, haploid stage between the diploid oocyte and
diploid zygote.
General Physiology

The larger amount of cytoplasm contained in the female gamete is used to


supply the developing zygote with nutrients during the period between
fertilization and implantation into the uterus. Interestingly, sperm contribute
only DNA at fertilization —not cytoplasm. Therefore, the cytoplasm and all of the
cytoplasmic organelles in the developing embryo are of maternal origin. This
includes mitochondria, which contain their own DNA. Scientific research in the
1980s determined that mitochondrial DNA was maternally inherited, meaning
that you can trace your mitochondrial DNA directly to your mother, her mother,
and so on back through your female ancestors.
Folliculogenesis
Again, ovarian follicles are oocytes and their supporting cells. They grow
and develop in a process called folliculogenesis, which typically leads to
ovulation of one follicle approximately every 28 days, along with death to
multiple other follicles. The death of ovarian follicles is called atresia, and can
occur at any point during follicular development. Recall that, a female infant at
birth will have one to two million oocytes within her ovarian follicles, and that
this number declines throughout life until menopause, when no follicles remain.
As you’ll see next, follicles progress from primordial, to primary, to secondary
and tertiary stages prior to ovulation—with the oocyte inside the follicle
remaining as a primary oocyte until right before ovulation.
Folliculogenesis begins with follicles in a resting state. These
small primordial follicles are present in newborn females and are the prevailing
follicle type in the adult ovary. Primordial follicles have only a single flat layer of
support cells, called granulosa cells, that surround the oocyte, and they can
stay in this resting state for years—some until right before menopause.
After puberty, a few primordial follicles will respond to a recruitment
signal each day, and will join a pool of immature growing follicles called primary
follicles. Primary follicles start with a single layer of granulosa cells, but the
granulosa cells then become active and transition from a flat or squamous
shape to a rounded, cuboidal shape as they increase in size and proliferate. As
the granulosa cells divide, the follicles—now called secondary follicles—increase
in diameter, adding a new outer layer of connective tissue, blood vessels,
and theca cells—cells that work with the granulosa cells to produce estrogens.
Within the growing secondary follicle, the primary oocyte now secretes a
thin acellular membrane called the zona pellucida that will play a critical role in
fertilization. A thick fluid, called follicular fluid, that has formed between the
granulosa cells also begins to collect into one large pool, or antrum. Follicles in
which the antrum has become large and fully formed are considered tertiary
follicles (or antral follicles). Several follicles reach the tertiary stage at the same
time, and most of these will undergo atresia. The one that does not die will
continue to grow and develop until ovulation, when it will expel its secondary
oocyte surrounded by several layers of granulosa cells from the ovary. Keep in
mind that most follicles don’t make it to this point. In fact, roughly 99 percent of
General Physiology

the follicles in the ovary will undergo atresia, which can occur at any stage of
folliculogenesis.

The Menstrual Cycle


Now that we have discussed the maturation of the cohort of tertiary
follicles in the ovary, the build-up and then shedding of the endometrial lining
in the uterus, and the function of the uterine tubes and vagina, we can put
everything together to talk about the three phases of the menstrual cycle—the
General Physiology

series of changes in which the uterine lining is shed, rebuilds, and prepares for
implantation.
The timing of the menstrual cycle starts with the first day of menses,
referred to as day one of a woman’s period. Cycle length is determined by
counting the days between the onset of bleeding in two subsequent cycles.
Because the average length of a woman’s menstrual cycle is 28 days, this is the
time period used to identify the timing of events in the cycle. However, the
length of the menstrual cycle varies among women, and even in the same
woman from one cycle to the next, typically from 21 to 32 days.
Just as the hormones produced by the granulosa and theca cells of the
ovary “drive” the follicular and luteal phases of the ovarian cycle, they also
control the three distinct phases of the menstrual cycle. These are the menses
phase, the proliferative phase, and the secretory phase.
Menses Phase
The menses phase of the menstrual cycle is the phase during which the lining
is shed; that is, the days that the woman menstruates. Although it averages
approximately five days, the menses phase can last from 2 to 7 days, or longer.
The menses phase occurs during the early days of the follicular phase of the
ovarian cycle, when progesterone, FSH, and LH levels are low. Recall that
progesterone concentrations decline as a result of the degradation of the corpus
luteum, marking the end of the luteal phase. This decline in progesterone
triggers the shedding of the stratum functionalis of the endometrium.
General Physiology

Proliferative Phase
General Physiology

Once menstrual flow ceases, the endometrium begins to proliferate again,


marking the beginning of the proliferative phase of the menstrual cycle. It
occurs when the granulosa and theca cells of the tertiary follicles begin to
produce increased amounts of estrogen. These rising estrogen concentrations
stimulate the endometrial lining to rebuild.
Recall that the high estrogen concentrations will eventually lead to a decrease in
FSH as a result of negative feedback, resulting in atresia of all but one of the
developing tertiary follicles. The switch to positive feedback—which occurs with
the elevated estrogen production from the dominant follicle—then stimulates the
LH surge that will trigger ovulation. In a typical 28-day menstrual cycle,
ovulation occurs on day 14. Ovulation marks the end of the proliferative phase
as well as the end of the follicular phase.
Secretory Phase
In addition to prompting the LH surge, high estrogen levels increase the uterine
tube contractions that facilitate the pick-up and transfer of the ovulated oocyte.
High estrogen levels also slightly decrease the acidity of the vagina, making it
more hospitable to sperm. In the ovary, the luteinization of the granulosa cells
of the collapsed follicle forms the progesterone-producing corpus luteum,
marking the beginning of the luteal phase of the ovarian cycle. In the uterus,
progesterone from the corpus luteum begins the secretory phase of the
menstrual cycle, in which the endometrial lining prepares for implantation. Over
the next 10 to 12 days, the endometrial glands secrete a fluid rich in glycogen. If
fertilization has occurred, this fluid will nourish the ball of cells now developing
from the zygote. At the same time, the spiral arteries develop to provide blood to
the thickened stratum functionalis.
If no pregnancy occurs within approximately 10 to 12 days, the corpus luteum
will degrade into the corpus albicans. Levels of both estrogen and progesterone
will fall, and the endometrium will grow thinner. Prostaglandins will be secreted
that cause constriction of the spiral arteries, reducing oxygen supply. The
endometrial tissue will die, resulting in menses—or the first day of the next
cycle.
Fertilization occurs when a sperm and an oocyte (egg) combine and their
nuclei fuse. Because each of these reproductive cells is a haploid cell containing
half of the genetic material needed to form a human being, their combination
forms a diploid cell. This new single cell, called a zygote, contains all of the
genetic material needed to form a human—half from the mother and half from
the father.
Transit of Sperm
Fertilization is a numbers game. During ejaculation, hundreds of millions of
sperm (spermatozoa) are released into the vagina. Almost immediately, millions
of these sperm are overcome by the acidity of the vagina (approximately pH 3.8),
and millions more may be blocked from entering the uterus by thick cervical
General Physiology

mucus. Of those that do enter, thousands are destroyed by phagocytic uterine


leukocytes. Thus, the race into the uterine tubes, which is the most typical site
for sperm to encounter the oocyte, is reduced to a few thousand contenders.
Their journey—thought to be facilitated by uterine contractions—usually takes
from 30 minutes to 2 hours. If the sperm do not encounter an oocyte
immediately, they can survive in the uterine tubes for another 3–5 days. Thus,
fertilization can still occur if intercourse takes place a few days before ovulation.
In comparison, an oocyte can survive independently for only approximately 24
hours following ovulation. Intercourse more than a day after ovulation will
therefore usually not result in fertilization.
During the journey, fluids in the female reproductive tract prepare the sperm for
fertilization through a process called capacitation, or priming. The fluids
improve the motility of the spermatozoa. They also deplete cholesterol molecules
embedded in the membrane of the head of the sperm, thinning the membrane in
such a way that will help facilitate the release of the lysosomal (digestive)
enzymes needed for the sperm to penetrate the oocyte’s exterior once contact is
made. Sperm must undergo the process of capacitation in order to have the
“capacity” to fertilize an oocyte. If they reach the oocyte before capacitation is
complete, they will be unable to penetrate the oocyte’s thick outer layer of cells.
Contact Between Sperm and Oocyte
Upon ovulation, the oocyte released by the ovary is swept into—and along—the
uterine tube. Fertilization must occur in the distal uterine tube because an
unfertilized oocyte cannot survive the 72-hour journey to the uterus. As you will
recall from your study of the oogenesis, this oocyte (specifically a secondary
oocyte) is surrounded by two protective layers. The corona radiata is an outer
layer of follicular (granulosa) cells that form around a developing oocyte in the
ovary and remain with it upon ovulation. The underlying zona
pellucida (pellucid = “transparent”) is a transparent, but thick, glycoprotein
membrane that surrounds the cell’s plasma membrane.
As it is swept along the distal uterine tube, the oocyte encounters the surviving
capacitated sperm, which stream toward it in response to chemical attractants
released by the cells of the corona radiata. To reach the oocyte itself, the sperm
must penetrate the two protective layers. The sperm first burrow through the
cells of the corona radiata. Then, upon contact with the zona pellucida, the
sperm bind to receptors in the zona pellucida. This initiates a process called
the acrosomal reaction in which the enzyme-filled “cap” of the sperm, called
the acrosome, releases its stored digestive enzymes. These enzymes clear a path
through the zona pellucida that allows sperm to reach the oocyte. Finally, a
single sperm makes contact with sperm-binding receptors on the oocyte’s
plasma membrane. The plasma membrane of that sperm then fuses with the
oocyte’s plasma membrane, and the head and mid-piece of the “winning” sperm
enter the oocyte interior.
General Physiology

How do sperm penetrate the corona radiata? Some sperm undergo a


spontaneous acrosomal reaction, which is an acrosomal reaction not triggered
by contact with the zona pellucida. The digestive enzymes released by this
reaction digest the extracellular matrix of the corona radiata. As you can see,
the first sperm to reach the oocyte is never the one to fertilize it. Rather,
hundreds of sperm cells must undergo the acrosomal reaction, each helping to
degrade the corona radiata and zona pellucida until a path is created to allow
one sperm to contact and fuse with the plasma membrane of the oocyte. If you
consider the loss of millions of sperm between entry into the vagina and
degradation of the zona pellucida, you can understand why a low sperm count
can cause male infertility.

When the first sperm fuses with the oocyte, the oocyte deploys two mechanisms
to prevent polyspermy, which is penetration by more than one sperm. This is
critical because if more than one sperm were to fertilize the oocyte, the resulting
zygote would be a triploid organism with three sets of chromosomes. This is
incompatible with life.
The first mechanism is the fast block, which involves a near instantaneous
change in sodium ion permeability upon binding of the first sperm, depolarizing
the oocyte plasma membrane and preventing the fusion of additional sperm
cells. The fast block sets in almost immediately and lasts for about a minute,
during which time an influx of calcium ions following sperm penetration triggers
the second mechanism, the slow block. In this process, referred to as
the cortical reaction, cortical granules sitting immediately below the oocyte
plasma membrane fuse with the membrane and release zonal inhibiting proteins
and mucopolysaccharides into the space between the plasma membrane and
the zona pellucida. Zonal inhibiting proteins cause the release of any other
General Physiology

attached sperm and destroy the oocyte’s sperm receptors, thus preventing any
more sperm from binding. The mucopolysaccharides then coat the nascent
zygote in an impenetrable barrier that, together with hardened zona pellucida, is
called a fertilization membrane.
The Zygote
Recall that at the point of fertilization, the oocyte has not yet completed meiosis;
all secondary oocytes remain arrested in metaphase of meiosis II until
fertilization. Only upon fertilization does the oocyte complete meiosis. The
unneeded complement of genetic material that results is stored in a second
polar body that is eventually ejected. At this moment, the oocyte has become an
ovum, the female haploid gamete. The two haploid nuclei derived from the
sperm and oocyte and contained within the egg are referred to as pronuclei.
They decondense, expand, and replicate their DNA in preparation for mitosis.
The pronuclei then migrate toward each other, their nuclear envelopes
disintegrate, and the male- and female-derived genetic material intermingles.
This step completes the process of fertilization and results in a single-celled
diploid zygote with all the genetic instructions it needs to develop into a human.
Most of the time, a woman releases a single egg during an ovulation cycle.
However, in approximately 1 percent of ovulation cycles, two eggs are released
and both are fertilized. Two zygotes form, implant, and develop, resulting in the
birth of dizygotic (or fraternal) twins. Because dizygotic twins develop from two
eggs fertilized by two sperm, they are no more identical than siblings born at
different times.
Much less commonly, a zygote can divide into two separate offspring during
early development. This results in the birth of monozygotic (or identical) twins.
Although the zygote can split as early as the two-cell stage, splitting occurs
most commonly during the early blastocyst stage, with roughly 70–100 cells
present. These two scenarios are distinct from each other, in that the twin
embryos that separated at the two-cell stage will have individual placentas,
whereas twin embryos that form from separation at the blastocyst stage will
share a placenta and a chorionic cavity.
Sexual Differentiation
Sexual differentiation does not begin until the fetal period, during weeks 9–12.
Embryonic males and females, though genetically distinguishable, are
morphologically identical. Bipotential gonads, or gonads that can develop into
male or female sexual organs, are connected to a central cavity called the cloaca
via Müllerian ducts and Wolffian ducts. (The cloaca is an extension of the
primitive gut.) Several events lead to sexual differentiation during this period.
During male fetal development, the bipotential gonads become the testes and
associated epididymis. The Müllerian ducts degenerate. The Wolffian ducts
become the vas deferens, and the cloaca becomes the urethra and rectum.
General Physiology

During female fetal development, the bipotential gonads develop into ovaries.
The Wolffian ducts degenerate. The Müllerian ducts become the uterine tubes
and uterus, and the cloaca divides and develops into a vagina, a urethra, and a
rectum.

The Fetal Circulatory System


During prenatal development, the fetal circulatory system is integrated with the
placenta via the umbilical cord so that the fetus receives both oxygen and
nutrients from the placenta. However, after childbirth, the umbilical cord is
severed, and the newborn’s circulatory system must be reconfigured. When the
heart first forms in the embryo, it exists as two parallel tubes derived from
mesoderm and lined with endothelium, which then fuse together. As the embryo
develops into a fetus, the tube-shaped heart folds and further differentiates into
General Physiology

the four chambers present in a mature heart. Unlike a mature cardiovascular


system, however, the fetal cardiovascular system also includes circulatory
shortcuts, or shunts. A shunt is an anatomical (or sometimes surgical)
diversion that allows blood flow to bypass immature organs such as the lungs
and liver until childbirth.
The placenta provides the fetus with necessary oxygen and nutrients via the
umbilical vein. (Remember that veins carry blood toward the heart. In this case,
the blood flowing to the fetal heart is oxygenated because it comes from the
placenta. The respiratory system is immature and cannot yet oxygenate blood
on its own.) From the umbilical vein, the oxygenated blood flows toward the
inferior vena cava, all but bypassing the immature liver, via the ductus
venosus shunt. The liver receives just a trickle of blood, which is all that it
needs in its immature, semifunctional state. Blood flows from the inferior vena
cava to the right atrium, mixing with fetal venous blood along the way.
Although the fetal liver is semifunctional, the fetal lungs are nonfunctional. The
fetal circulation therefore bypasses the lungs by shifting some of the blood
through the foramen ovale, a shunt that directly connects the right and left
atria and avoids the pulmonary trunk altogether. Most of the rest of the blood is
pumped to the right ventricle, and from there, into the pulmonary trunk, which
splits into pulmonary arteries. However, a shunt within the pulmonary artery,
the ductus arteriosus, diverts a portion of this blood into the aorta. This
ensures that only a small volume of oxygenated blood passes through the
immature pulmonary circuit, which has only minor metabolic requirements.
Blood vessels of uninflated lungs have high resistance to flow, a condition that
encourages blood to flow to the aorta, which presents much lower resistance.
The oxygenated blood moves through the foramen ovale into the left atrium,
where it mixes with the now deoxygenated blood returning from the pulmonary
circuit. This blood then moves into the left ventricle, where it is pumped into the
aorta. Some of this blood moves through the coronary arteries into the
myocardium, and some moves through the carotid arteries to the brain.
The descending aorta carries partially oxygenated and partially deoxygenated
blood into the lower regions of the body. It eventually passes into the umbilical
arteries through branches of the internal iliac arteries. The deoxygenated blood
collects waste as it circulates through the fetal body and returns to the
umbilical cord. Thus, the two umbilical arteries carry blood low in oxygen and
high in carbon dioxide and fetal wastes. This blood is filtered through the
placenta, where wastes diffuse into the maternal circulation. Oxygen and
nutrients from the mother diffuse into the placenta and from there into the fetal
blood, and the process repeats.
General Physiology

Other Organ Systems


During weeks 9–12 of fetal development, the brain continues to expand, the
body elongates, and ossification continues. Fetal movements are frequent
during this period, but are jerky and not well-controlled. The bone marrow
begins to take over the process of erythrocyte production—a task that the liver
performed during the embryonic period. The liver now secretes bile. The fetus
circulates amniotic fluid by swallowing it and producing urine. The eyes are
well-developed by this stage, but the eyelids are fused shut. The fingers and toes
begin to develop nails. By the end of week 12, the fetus measures approximately
9 cm (3.5 in) from crown to rump.
Weeks 13–16 are marked by sensory organ development. The eyes move closer
together; blinking motions begin, although the eyes remain sealed shut. The lips
exhibit sucking motions. The ears move upward and lie flatter against the head.
The scalp begins to grow hair. The excretory system is also developing: the
kidneys are well-formed, and meconium, or fetal feces, begins to accumulate in
the intestines. Meconium consists of ingested amniotic fluid, cellular debris,
mucus, and bile.
During approximately weeks 16–20, as the fetus grows and limb movements
become more powerful, the mother may begin to feel quickening, or fetal
movements. However, space restrictions limit these movements and typically
force the growing fetus into the “fetal position,” with the arms crossed and the
General Physiology

legs bent at the knees. Sebaceous glands coat the skin with a waxy, protective
substance called vernix caseosa that protects and moisturizes the skin and
may provide lubrication during childbirth. A silky hair called lanugo also covers
the skin during weeks 17–20, but it is shed as the fetus continues to grow.
Extremely premature infants sometimes exhibit residual lanugo.
Developmental weeks 21–30 are characterized by rapid weight gain, which is
important for maintaining a stable body temperature after birth. The bone
marrow completely takes over erythrocyte synthesis, and the axons of the spinal
cord begin to be myelinated, or coated in the electrically insulating glial cell
sheaths that are necessary for efficient nervous system functioning. (The
process of myelination is not completed until adolescence.) During this period,
the fetus grows eyelashes. The eyelids are no longer fused and can be opened
and closed. The lungs begin producing surfactant, a substance that reduces
surface tension in the lungs and assists proper lung expansion after birth.
Inadequate surfactant production in premature newborns may result in
respiratory distress syndrome, and as a result, the newborn may require
surfactant replacement therapy, supplemental oxygen, or maintenance in a
continuous positive airway pressure (CPAP) chamber during their first days or
weeks of life. In male fetuses, the testes descend into the scrotum near the end
of this period. The fetus at 30 weeks measures 28 cm (11 in) from crown to
rump and exhibits the approximate body proportions of a full-term newborn,
but still is much leaner.
The fetus continues to lay down subcutaneous fat from week 31 until birth. The
added fat fills out the hypodermis, and the skin transitions from red and
wrinkled to soft and pink. Lanugo is shed, and the nails grow to the tips of the
fingers and toes. Immediately before birth, the average crown-to-rump length is
35.5–40.5 cm (14–16 in), and the fetus weighs approximately 2.5–4 kg (5.5–8.8
lbs). Once born, the newborn is no longer confined to the fetal position, so
subsequent measurements are made from head-to-toe instead of from crown-to-
rump. At birth, the average length is approximately 51 cm (20 in).
General Physiology

General Physiology
Module 3: Plant
Physiology

Welcome to the Module 1!


In this unit, you will focus on
the functions and requirements of
life, homeostasis and regulation of
hormones, transport of substances,
and energy metabolism. These are
essential concepts that underpin our
understanding of how the human
body works. You will explore how
different systems and organs interact
to maintain homeostasis, regulate
hormones, transport essential
substances, and generate and use
energy.
The principles you learn here
will provide a strong foundation for
your further studies in human
physiology and will help you
understand the complex mechanisms
involved in maintaining health and
preventing disease. Get ready to
delve into the exciting world of
human physiology and learn about
the wonders of the human body!
General Physiology

Lesson 1: Water Balance of Plant

Water in plant life Water plays a crucial role in the life of plant. It is the most
abundant constituents of most organisms. Water typically accounts for more
than 70 percent by weight of non-woody plant parts. The water content of plants
is in a continual state of flux. The constant flow of water through plants is a
matter of considerable significance to their growth and survival. The uptake of
water by cells generates a pressure
known as turgor. Photosynthesis
requires that plants draw carbon
dioxide from the atmosphere, and at
the same time exposes them to water
loss. To prevent leaf desiccation, water
must be absorbed by the roots, and
transported through the plant body.
Balancing the uptake, transport, and
loss of water represents an important
challenge for land plants. The thermal
properties of water contribute to
temperature regulation, helping to ensure that plants do not cool down or heat
up too rapidly. Water has excellent solvent properties. Many of the biochemical
reactions occur in water and water is itself either a reactant or a product in a
large number of those reactions.

The practice of crop irrigation reflects the fact that water is a key resource
limiting agricultural productivity. Water availability likewise limits the
productivity of natural ecosystems. Plants use water in huge amounts, but only
small part of that remains in the plant to supply growth. About 97% of water
taken up by plants is lost to the atmosphere, 2% is used for volume increase or
cell expansion, and 1% for metabolic processes, predominantly photosynthesis.
Water loss to the atmosphere appears to be an inevitable consequence of
carrying out photosynthesis. The uptake of CO2 is coupled to the loss of water.
Because the driving gradient for water loss from leaves is much larger than that
for CO2 uptake, as many as 400 water molecules are lost for every CO2
molecule gained.

Water potential

 The structure and properties of water

Water is composed of an oxygen atom that is covalently bonded to two hydrogen


atoms. The oxygen atom has a partial negative charge, and the partial positive
charge is shared between the two hydrogen atoms, resulting in a polar molecule.
This polar structure leads to strong hydrogen bonding between adjacent water
molecules and with other polar molecules, making water a good solvent for ionic
substances and for molecules like sugars and proteins.
General Physiology

Water has a high specific heat capacity and a high latent heat of vaporization
due to its hydrogen bonding. Additionally, liquid water has high thermal
conductivity, which allows it to absorb and distribute large amounts of heat
energy rapidly. This makes water effective in dissipating the heat from
biochemical reactions in cells. The good solvent properties of water are also due
to its polar character, as measured by its high dielectric constant of 78.4.

The cohesive nature of water molecules due to hydrogen bonding results in


water having high surface tension and adhesion. This leads to capillarity, where
water rises in capillary tubes, and is important in maintaining the continuity of
water columns in plants. Additionally, hydrogen bonding gives water a high
tensile strength, enabling it to withstand negative pressures of up to -20 MPa.
These unique properties of water play critical roles in various biological
processes, such as water transport in plants and the stability of proteins in
cells.

 Water movement by diffusion, osmosis, and bulk flow

Water moves in plants through the xylem


tissues in two ways: bulk flow and diffusion.
Bulk flow is the movement of materials driven
by external forces, such as pressure or gravity,
resulting in all the molecules of a substance
moving in mass. On the other hand, diffusion is
the movement of molecules from high
concentration to low concentration driven by
the natural tendency for systems to reach an
even distribution of molecules and towards the
lowest possible energy state. This tendency can
be described by Fick's first law and the second law of thermodynamics.
However, diffusion is only effective over short distances and is too slow for long
distances.

Osmosis is the net movement of water across a selectively permeable barrier,


such as the membrane of plant cells. This movement is facilitated by
aquaporins, which are proteins that form water-selective channels across the
membrane. Osmosis is driven by the maximization of entropy, with the volume
of solvent diffusing through the membrane to dilute the solute. This process can
be demonstrated using an osmometer, where a sugar solution in a thistle tube
will increase in volume over time until the hydrostatic pressure balances the
force driving the water into the solution.

Water Potential
Water potential is a measure of the potential energy in water. Plant physiologists
are not interested in the energy in any one particular aqueous system, but are
very interested in water movement between two systems. In practical terms,
therefore, water potential is the difference in potential energy between a given
water sample and pure water (at atmospheric pressure and ambient
temperature). Water potential is denoted by the Greek letter ψ (psi) and is
General Physiology

expressed in units of pressure (pressure is a form of energy) called megapascals


(MPa). The potential of pure water (Ψwpure H2O) is, by convenience of
definition, designated a value of zero (even though pure water contains plenty of
potential energy, that energy is ignored). Water potential values for the water in
a plant root, stem, or leaf are therefore expressed relative to Ψwpure H2O.

The water potential in plant solutions is influenced by solute concentration,


pressure, gravity, and factors called matrix effects. Water potential can be
broken down into its individual components using the following equation:

Ψsystem = Ψtotal = Ψs + Ψp + Ψg + Ψm

where Ψs, Ψp, Ψg, and Ψm refer to the solute, pressure, gravity, and matric
potentials, respectively. “System” can refer to the water potential of the soil
water (Ψsoil), root water (Ψroot), stem water (Ψstem), leaf water (Ψleaf) or the
water in the atmosphere (Ψatmosphere): whichever aqueous system is under
consideration. As the individual components change, they raise or lower the
total water potential of a system. When this happens, water moves to
equilibrate, moving from the system or compartment with a higher water
potential to the system or compartment with a lower water potential. This brings
the difference in water potential between the two systems (ΔΨ) back to zero (ΔΨ
= 0). Therefore, for water to move through the plant from the soil to the air (a
process called transpiration),Ψsoil must be >Ψroot >Ψstem
>Ψleaf>Ψatmosphere.

Water only moves in response to ΔΨ, not in response to the individual


components. However, because the individual components influence the
totalΨsystem, by manipulating the individual components (especially Ψs), a
plant can control water movement.

Absorption by Roots

 Water in the soil


The water content and the rate of water movement in soils depend to a large
extent on soil type and soil structure. Like the water potential of the plant cells,
the water potential of soils may be dissected into three components: the osmotic
potential, the hydrostatic pressure and the gravitational potential. The osmotic
potential (Ψs) of soil water is generally negligible. The second component of soil
water potential is hydrostatic pressure (Ψp). For wet soils, Ψp is very close to
zero. As soil dries out Ψp decreases and can become quite negative. As the
water content of the soil decreases, the water recedes into the interstices
between soil particles, forming air-water surfaces whose curvature represents
the balance between the tendency to minimize the surface area of the air-water
interface and the attraction of the water for the soil particles. Water under a
curved surface develops a negative pressure (like in leaf mesophyll). As soil dries
out, water is first removed from the largest spaces between soil particles. The
value of Ψp may easily reach -1 to -2 MPa as the air-water interface recedes into
General Physiology

the smaller spaces between clay particles. The third component is gravitational
potential (Ψg). Gravity plays an important role in drainage.

 Water absorption by roots

Intimate contact between the surface of


root and the soil is essential for effective
water absorption. Root hairs are
filamentous outgrowths of root epidermal
cells that greatly increase the surface area
of the root, thus providing greater capacity
for absorption of ions and water from the
soil. Water enters the root most readily
near the root tip. The intimate contact
between the soil and the root surface is
easily ruptured when the soil is disturbed. It is for this reason that newly
transplanted seedlings and plants need to be protected from water loss for the
first few days after transplantation.

From the epidermis to the endodermis


of the root, there are three pathways
through which water can flow: the
apoplast, the symplast and the
transmembrane pathway.

1. The apoplast is the continuous system


of cell walls and intercellular air spaces.
In this pathway water moves without
crossing any membranes as it travels
across the root cortex.

2. The symplast consists of the entire network of cell cytoplasm interconnected


by plasmodesmata. In this pathway, water travels across the root cortex via the
plasmodesmata.

3. The transmembrane pathway is the route by which water enters a cell on one
side, exits the cell on the other side, enters the next in the series, and so on. In
this pathway, water crosses the plasma membrane of each cell in its path twice.

Water moves through three pathways in plants, but its movement is


determined by gradients and resistances, not just a single pathway. The
Casparian strip in the endodermis blocks the apoplast pathway and forces water
and solutes to pass through the plasma membrane, ensuring that water moves
General Physiology

symplastically across the endodermis. This requirement highlights the


importance of aquaporins in root permeability.

Root permeability can be affected by changes in temperature and oxygen


levels. Low temperature and anaerobic conditions can reduce the rate of
respiration, causing an increase in intracellular pH and altering the
conductance of aquaporins, making roots less permeable to water.

Root pressure is a phenomenon seen in some plants where sap is exuded


from cut xylem and positive pressures can be measured. This often results in
liquid droplets on the edges of leaves, known as guttation, which is most visible
when transpiration is suppressed and relative humidity is high, such as at
night.

Transpiration

Water movement is determined by differences in water potential. It can be


assumed that the driving force for transpiration is the difference in water
potential between the substomatal air space and the external atmosphere.
However, because the problem is now concerned with the diffusion of water
vapour rather than liquid water, it will be more convenient to think in terms of
vapour systems. We can say that when a gas phase has reached equilibrium
and is saturated with water vapour, the system will have achieved its saturation
vapour pressure.

The vapour pressure over a solution at atmospheric pressure is influenced


by solute concentration and mainly by temperature. In principle we can assume
that the substomatal air space of leaf is normally saturated or very nearly
saturated with water vapour. On the other hand, the atmosphere that
surrounds the leaf is usually unsaturated and may often have a very low water
content. This difference in water vapour pressure between the internal air
spaces of the leaf and the surrounding air is the driving force of transpiration.

On its way from the leaf to the atmosphere, water is pulled from the xylem
into the cell walls of the mesophyll, where it evaporates into the air spaces of the
leaf. The water vapor than exits the leaf through the stomatal pore. The
movement of liquid water through the living tissues of the leaf is controlled by
gradients in water potential. However, transport in the vapor phase is by
diffusion, so the final part of the transpiration stream is controlled by the
concentration gradient of water vapor. Almost all of the water lost from leaves is
lost by diffusion of water vapour through the tiny stomatal pores. The stomatal
transpiration accounts for 90 to 95% of water loss from leaves. The remaining 5
to 10% is accounted for by cuticular transpiration. In most herbaceous species,
stomata are present in both the upper and lower surfaces of the leaf, usually
more abundant on the lower surface. In many tree species, stomata are located
only on the lower surface of the leaf.

 The driving force for transpiration is the difference in water vapour


concentration.
General Physiology

Transpiration from the leaf depends on two major factors: (1) the difference in
water vapor concentration between the leaf air spaces and the external bulk air
and (2) the diffusional resistance of this pathway. Air space volume is about
10% in corn leaves, 30% in barley, and 40% in tobacco leaves. In contrast to the
volume of the air space, the internal surface area from which water evaporates
may be from 7 to 30 times the external leaf area. The air space in the leaf is
close to water potential equilibrium with the cell wall surfaces from which liquid
water is evaporating. The concentration of water vapor changes at various
points along the transpiration pathway from the cell wall surface to the bulk air
outside the leaf.

The second important factor governing water loss from the leaf is the diffusional
resistance of the transpiration pathway, which consists of two varying
components:

1. The resistance associated with diffusion through the stomatal pore, the leaf
stomatal resistance.

2. The resistance due to the layer of unstirred air next to the leaf surface
through which water vapor must diffuse to reach the turbulent air of the
atmosphere. This second resistance is called the leaf boundary layer resistance.

You might also like