You are on page 1of 278

Health Science Tool Kit

Content Resource

Assembled by:
Ranjna Kapoor
Glenn Harrison
David Simcock
Kate Miller
Pam Megaw
Daniela Johnson
January 2018
1 | AN INTRODUCTION TO
THE HUMAN BODY

Figure 1.1 Blood Pressure A proficiency in anatomy and physiology is fundamental to any career in the health
professions. (credit: Bryan Mason/flickr)

After studying this chapter, you will be able to:


• Distinguish between anatomy and physiology, and identify several branches of each
• Identify the different organ systems of the human body
• Use appropriate anatomical terminology to identify key body structures, body
regions, and directions in the body
• Understand the logic behind basic medical terminology
• Compare and contrast at least four medical imagining techniques in terms of their
function and use in medicine
Though you may approach a course in anatomy and physiology strictly as a requirement for
your field of study, the knowledge you gain in this course will serve you well in many aspects
of your life. An understanding of anatomy and physiology is not only fundamental to any
career in the health professions, but it can also benefit your own health. Familiarity with the
human body can help you make healthful choices and prompt you to take appropriate
action when signs of illness arise. Your knowledge in this field will help you understand news
about nutrition, medications, medical devices, and procedures and help you understand
genetic or infectious diseases.
This chapter begins with an overview of anatomy and physiology and a preview of the body
regions and functions. It then introduces a set of standard terms for body structures and
positions in the body, as well as common medical terms that will serve as a foundation for
more comprehensive information covered later in your degree courses. The chapter
concludes with a brief descriptions and uses of four medical imaging techniques.

1.1 | Overview of Anatomy and Physiology


Human anatomy is the scientific study of the body’s structures. In the past, anatomy has
primarily been studied via observing injuries, and later by the dissection of anatomical
structures of cadavers, but in the past century, computer-assisted imaging techniques have
allowed clinicians to look inside the living body. Physiology explains how the structures of
the body work together to maintain life. It is difficult to study structure (anatomy) without
knowledge of function (physiology). The two disciplines are typically studied together
because form and function are closely related in all living things.
Human anatomy is the scientific study of the body’s structures. Some of these structures
are very small and can only be observed and analysed with the assistance of a microscope.
Other larger structures can readily be seen, manipulated, measured, and weighed. The word
“anatomy” comes from a Greek root that means “to cut apart.” Human anatomy was first
studied by observing the exterior of the body and observing the wounds of soldiers and
other injuries. Later, physicians were allowed to dissect bodies of the dead to augment their
knowledge. When a body is dissected, its structures are cut apart in order to observe their
physical attributes and their relationships to one another. Dissection is still used in medical
schools, anatomy courses, and in pathology labs. In order to observe structures in living
people, however, a number of imaging techniques have been developed. These techniques
allow clinicians to visualise structures inside the living body such as a cancerous tumour or a
fractured bone.
Like most scientific disciplines, anatomy has areas of specialisation. Gross anatomy is the
study of the larger structures of the body, those visible without the aid of magnification
(Figure 1.2a).
Figure 1.2 Gross and Microscopic Anatomy (a) Gross anatomy considers large structures such as the brain.
(b) Microscopic anatomy can deal with the same structures, though at a different scale. This is a micrograph of
nerve cells from the brain. LM × 1600. (credit a: “WriterHound”/Wikimedia Commons; credit b: Micrograph
provided by the Regents of University of Michigan Medical School © 2012)

Macro- means “large,” thus, gross anatomy is also referred to as macroscopic anatomy. In
contrast, micro- means “small,” and microscopic anatomy is the study of structures that can
be observed only with the use of a microscope or other magnification devices (Figure 1.2b).
Microscopic anatomy includes cytology, the study of cells and histology, the study of tissues.
As the technology of microscopes has advanced, anatomists have been able to observe
smaller and smaller structures of the body, from slices of large structures like the heart to
neuronal cells. Even the minute structures of an individual cell can be viewed by electron
microscopes, as will be discussed in the next chapter.
Anatomists take two general approaches to the study of the body’s structures: regional and
systemic. Regional anatomy is the study of the interrelationships of all of the structures in a
specific body region, such as the abdomen. Studying regional anatomy helps us appreciate
the interrelationships of body structures, such as how muscles, nerves, blood vessels, and
other structures work together to serve a particular body region. In contrast, systemic
anatomy is the study of the structures that make up a discrete body system—that is, a
group of structures that work together to perform a unique body function. For example, a
systemic anatomical study of the digestive system (an organ system) would consider all of
the organs involved in digestion in the human body.
Whereas anatomy is about structure, physiology is about function. Human physiology is the
scientific study of the chemistry and physics of the structures of the body and the ways in
which they work together to support the functions of life. Much of the study of physiology
centres on the body’s tendency toward homeostasis. Homeostasis is the state of steady
internal conditions maintained by living things. This steady state for conditions, such as
temperature and concentration of diverse chemicals, is necessary for the body to function
properly.
The study of physiology certainly includes observation, both with the naked eye and with
microscopes, as well as manipulations and measurements. However, current advances in
physiology usually depend on carefully designed laboratory experiments that reveal the
functions of the many structures and chemical compounds that make up the human body.
Like anatomists, physiologists typically specialise in a particular branch of physiology. For
example, neurophysiology is the study of the brain, spinal cord, and nerves and how these
work together to perform functions as complex and diverse as vision, movement, and
thinking. Physiologists may work from the organ level (exploring, for example, what
different parts of the brain do) to the molecular level (such as exploring how an
electrochemical signal travels along nerves).
Form (anatomy) is closely related to function (physiology) in all living things. For example,
the thin flap of your eyelid can snap down to clear away dust particles and almost
instantaneously slide back up to allow you to see again. At the microscopic level, the
arrangement and function of the nerves and muscles that serve the eyelid allow for its quick
action and retreat. At an even smaller level of analysis, the function of these nerves and
muscles likewise relies on the interactions of specific molecules and ions. Even the three-
dimensional structure of certain molecules is essential to their function.
Your study of anatomy and physiology will make more sense if you continually relate the
form of the structures you are studying to their function. In fact, it can be somewhat
frustrating to attempt to study anatomy without an understanding of the physiology that a
body structure supports. Imagine, for example, trying to appreciate the unique arrangement
of the bones of the human hand (that is, its form) if you had no conception of the function
of the hand. Fortunately, your understanding of how the human hand manipulates tools—
from pens to cell phones—helps you appreciate the unique alignment of the thumb in
opposition to the four fingers, making your hand a structure that allows you to pinch and
grasp objects and type text messages.

1.2 | Organ systems of the human body


Figure 1.3 represents the 11 distinct body systems. Assigning organs to organ systems can
be imprecise since organs that “belong” to one system can also have functions integral to
another system. In fact, most organs contribute to more than one system.
Figure 1.3 Organ Systems of the Human Body Organs that work together are grouped into organ systems.
Figure 1.3 Organ Systems of the Human Body Organs that work together are grouped into organ systems.
The organism level is the highest level of organisation. An organism is a living being that has
a cellular structure and that can independently perform all physiologic functions necessary
for life. In multicellular organisms, including humans, all cells, tissues, organs, and organ
systems of the body work together to maintain the life and health of the organism.

1.3 | Anatomical Terminology


Anatomists and health care providers use terminology that can be bewildering to the
uninitiated. However, the purpose of this language is not to confuse, but rather to increase
precision and reduce medical errors. For example, is a scar “above the wrist” located on the
forearm two or three inches away from the hand? Or is it at the base of the hand? Is it on
the palm side or back-side? By using precise anatomical terminology, we eliminate
ambiguity. Anatomical terms derive from ancient Greek and Latin words.

Anatomical Position
To further increase precision, anatomists standardise the way in which they view the body.
Just as maps are normally oriented with north at the top, the standard body “map,”
or anatomical position, is that of the body standing upright, with the feet at shoulder width
and parallel, toes forward. The upper limbs are held out to each side, and the palms of the
hands face forward as illustrated in Figure 1.4. Using this standard position reduces
confusion and therefore,
• Anterior denotes towards the front
• Posterior denotes towards the back
It does not matter how the body being described is oriented, the terms are used as if it is in
anatomical position. For example, a scar in the “anterior (front) carpal (wrist) region” would
be present on the palm side of the wrist. The term “anterior” would be used even if the
hand were palm down on a table.
A body that is lying down is described as either prone or supine. Prone describes a face-
down orientation, and supine describes a face up orientation. These terms are sometimes
used in describing the position of the body during specific physical examinations or surgical
procedures.
Figure 1.4 Regions of the Human Body. The human body is shown in anatomical position in an (a) anterior
view and a (b) posterior view

Directional Terms
Certain directional anatomical terms appear throughout this and any other anatomy
textbooks (Figure 1.5). These terms are essential for describing the relative locations of
different body structures. For instance, an anatomist might describe one band of tissue as
“inferior to” another or a physician might describe a tumour as “superficial to” a deeper
body structure. Commit these terms to memory to avoid confusion when you are studying
or describing the locations of particular body parts.
• Anterior (or ventral) Describes the front or direction toward the front of the body.
The toes are anterior to the foot.
• Posterior (or dorsal) Describes the back or direction toward the back of the body.
The popliteus is posterior to the patella (refer to Figure 1.4).
• Superior (or cranial) describes a position above or higher than another part of the
body proper. The orbits of the eyes are superior to the oral cavity (mouth).
• Inferior (or caudal) describes a position below or lower than another part of the
body proper; near or toward the tail (in humans, the coccyx, or lowest part of the
spinal column). The pelvis is inferior to the abdomen.
• Lateral describes the side or direction toward the side of the body. The thumb
(pollex) is lateral to the digits.
• Medial describes the middle or direction toward the middle of the body. The rib
cage is medial to the arm (left or right).
• Proximal describes a position in a limb that is nearer to the point of attachment or
the trunk of the body. The right elbow joint is proximal to the right hand.
• Distal describes a position in a limb that is farther from the point of attachment or
the trunk of the body. The right hand is distal to the femur.
• Superficial describes a position closer to the surface of the body. The skin is
superficial to the bones.
• Deep describes a position farther from the surface of the body. The brain is deep to
the skull.

In addition, the following directional prefixes are also useful for understanding some of the
content presented in this text book:
• Epi: upper above
• Endo: inner, within
• Meso: middle, halfway, intermediate
• Trans: across
• Intra: within
• Inter: between
Figure 1.5 Directional Terms Applied to the Human Body Paired directional terms are shown as applied to
the human body.

Body Planes
A section is a two-dimensional surface of a three-dimensional structure that has been cut.
Modern medical imaging devices enable clinicians to obtain “virtual sections” of living
bodies. We call these scans. Body sections and scans can be correctly interpreted, however,
only if the viewer understands the plane along which the section was made. A plane is
an imaginary two-dimensional surface that passes through the body. There are three planes
commonly referred to in anatomy and medicine, as illustrated in Figure 1.6.
• The sagittal plane is the plane that divides the body or an organ vertically into right
and left sides. If this vertical plane runs directly down the middle of the body, it is
called the midsagittal or median plane. If it divides the body into unequal right and
left sides, it is called a parasagittal plane or less commonly a longitudinal section.
• The frontal plane is the plane that divides the body or an organ into an anterior
(front) portion and a posterior (rear) portion. The frontal plane is often referred to as
a coronal plane. (“Corona” is Latin for “crown.”)
• The transverse plane is the plane that divides the body or organ horizontally into
upper and lower portions. Transverse planes produce images referred to as cross
sections.

Figure 1.6 Planes of the Body The three planes most commonly used in anatomical and medical imaging are
the sagittal, frontal (or coronal), and transverse plane.
Body Cavities
The body maintains its internal organisation by means of membranes, sheaths, and other
structures that separate compartments. The dorsal (posterior) cavity and the ventral
(anterior) cavity are the largest body compartments (Figure 1.7). These cavities contain and
protect delicate internal organs, and the ventral cavity allows for significant changes in the
size and shape of the organs as they perform their functions. The lungs, heart, stomach, and
intestines, for example, can expand and contract without distorting other tissues or
disrupting the activity of nearby organs.

Figure 1.7 Dorsal and Ventral Body Cavities The ventral cavity includes the thoracic and abdominopelvic
cavities and their subdivisions. The dorsal cavity includes the cranial and spinal cavities.

Subdivisions of the Posterior (Dorsal) and Anterior (Ventral) Cavities


The posterior (dorsal) and anterior (ventral) cavities are each subdivided into smaller
cavities. In the posterior (dorsal) cavity, the cranial cavity houses the brain, and the spinal
cavity (or vertebral cavity) encloses the spinal cord. Just as the brain and spinal cord make
up a continuous, uninterrupted structure, the cranial and spinal cavities that house them
are also continuous. The brain and spinal cord are protected by the bones of the skull and
vertebral column and by cerebrospinal (cerebro=brain) fluid, a colourless fluid produced by
the brain, which cushions the brain and spinal cord within the posterior (dorsal) cavity.
The anterior (ventral) cavity has two main subdivisions: the thoracic cavity and the
abdominopelvic cavity (see Figure 1.7). The thoracic cavity is the more superior subdivision
of the anterior cavity, and it is enclosed by the rib cage. The thoracic cavity contains the
lungs and the heart, which is located in the mediastinum. The diaphragm forms the floor of
the thoracic cavity and separates it from the more inferior abdominopelvic cavity.
The abdominopelvic cavity is the largest cavity in the body. Although no membrane
physically divides the abdominopelvic cavity, it can be useful to distinguish between the
abdominal cavity, the division that houses the digestive organs, and the pelvic cavity, the
division that houses the organs of reproduction.

Basic medical terminology


In addition to anatomical terms described above, it is useful to discuss the origin of other
types of medical terms. Again, most medical terms usually have Latin or Greek origin. These
terms are made up of roots, prefixes, and suffixes. The root of a term often refers to an
organ, tissue, or condition, whereas the prefix or suffix often describes the root. For
example, in the disorder hypertension, the prefix “hyper-” means “high” or “over,” and the
root word “tension” refers to pressure, so the word “hypertension” refers to
abnormally high blood pressure. Similarly, in the condition of hypothyroidism, "hypo"
means "decreased" or "below", therefore hypothyroidism denotes an underactive thyroid
gland.
Many root words denote the specific structure or organ (Table 1.1), for example, the Latin
"ren/o"means kidney, therefore "renal failure" describes kidney failure; further, the heart is
represented by the term "Cardi/o", therefore cardiac arrest describes heart failure.
Table 1.1 Some examples of how the root words for organs are used in the construction of
medical terms

Organ Root word Example of medical term Meaning

lungs pneum/o pneumonia infection of the lungs

derm skin dermatitis inflammation of the skin

arthr/o joint arthritis disease: painful inflammation of joints

pod/o foot podiatry care of the human foot

hem/a blood hemolysis rupture of red blood cells

rhin/o nose rhinoplasty plastic surgery on nose

gastr/o stomach gastritis inflammation of the stomach lining


1.4 | Medical Imaging
By the end of this section, you will be able to discuss the uses and drawbacks of X-ray
imaging and identify four modern medical imaging techniques and how they are used
For thousands of years, fear of the dead and legal sanctions limited the ability of anatomists
and physicians to study the internal structures of the human body. An inability to control
bleeding, infection, and pain made surgeries infrequent, and those that were performed—
such as wound suturing, amputations, tooth and tumor removals, skull drilling, and
cesarean birth—did not greatly advance knowledge about internal anatomy. Theories about
the function of the body and about disease were therefore largely based on external
observations and imagination. During the fourteenth and fifteenth centuries, however, the
detailed anatomical drawings of Italian artist and anatomist Leonardo da Vinci and Flemish
anatomist Andreas Vesalius were published, and interest in human anatomy began to
increase. Medical schools began to teach anatomy using human dissection; although some
resorted to grave robbing to obtain corpses. Laws were eventually passed that
enabled students to dissect the corpses of criminals and those who donated their bodies for
research. Still, it was not until the late nineteenth century that medical researchers
discovered non-surgical methods to look inside the living body.

X-Rays
German physicist Wilhelm Röntgen (1845–1923) was experimenting with electrical current
when he discovered that a mysterious and invisible “ray” would pass through his flesh but
leave an outline of his bones on a screen coated with a metal compound. In 1895, Röntgen
made the first durable record of the internal parts of a living human: an “X-ray” image (as
it came to be called) of his wife’s hand. Scientists around the world quickly began their own
experiments with X-rays, and by 1900, X-rays were widely used to detect a variety of injuries
and diseases. In 1901, Röntgen was awarded the first Nobel Prize for physics for his work in
this field.
The X-ray is a form of high energy electromagnetic radiation with a short wavelength
capable of penetrating solids and ionising gases. As they are used in medicine, X-rays are
emitted from an X-ray machine and directed toward a specially treated metallic plate placed
behind the patient’s body. The beam of radiation results in darkening of the X-ray plate. X-
rays are slightly impeded by soft tissues, which show up as gray on the X-ray plate, whereas
hard tissues, such as bone, largely block the rays, producing a light-toned “shadow.” Thus, X-
rays are best used to visualise hard body structures such as teeth and bones (Figure 1.8).
Like many forms of high energy radiation, however, X-rays are capable of damaging cells
and initiating changes that can lead to cancer. This danger of excessive exposure to X-rays
was not fully appreciated for many years after their widespread use.
Figure 1.8 X-Ray of a Hand High energy electromagnetic radiation allows the internal structures of the body,
such as bones, to be seen in X-rays like these. (credit: Trace Meek/flickr)

Refinements and enhancements of X-ray techniques have continued throughout the


twentieth and twenty-first centuries. Although often supplanted by more sophisticated
imaging techniques, the X-ray remains a “workhorse” in medical imaging, especially for
viewing fractures and for dentistry. The disadvantage of harmful irradiation to the patient
and the operator is now attenuated by proper shielding and by limiting exposure.
Modern Medical Imaging
X-rays can depict a two-dimensional image of a body region, and only from a single angle. In
contrast, more recent medical imaging technologies produce data that is integrated and
analysed by computers to produce three-dimensional images or images that reveal aspects
of body functioning.

Computed Tomography
The term tomography (tomo= cutting of section, graphy=recording, in this case picture)
refers to imaging by sections. Computed tomography (CT) is a noninvasive imaging
technique as no actual cutting takes place and only "virtual cutting" is used. Computers are
used to analyse several cross-sectional X-rays in order to reveal minute details about
structures in the body (Figure 1.9a). The technique was invented in the 1970s and is based
on the principle that, as X-rays pass through the body, they are absorbed or reflected at
different levels. In the technique, a patient lies on a motorised platform while a
computerised axial tomography (CAT) scanner rotates 360 degrees around the patient,
taking X-ray images. A computer combines these images into a two-dimensional view of the
scanned area, or “slice.”
Since 1970, the development of more powerful computers and more sophisticated software
has made CT scanning routine for many types of diagnostic evaluations. It is especially useful
for soft tissue scanning, such as of the brain and the thoracic and abdominal viscera. Its level
of detail is so precise that it can allow physicians to measure the size of a mass down to
a millimetre. The main disadvantage of CT scanning is that it exposes patients to a dose of
radiation many times higher than that of X-rays. In fact, children who undergo CT scans are
at increased risk of developing cancer, as are adults who have multiple CT scans.
A CT or CAT scan relies on a circling scanner that revolves around the patient’s body. Watch
this video (http://openstaxcollege.org/l/CATscan) to learn more about CT and CAT scans.
What type of radiation does a CT scanner use?

Magnetic Resonance Imaging


Magnetic resonance imaging (MRI) is a noninvasive medical imaging technique based on a
phenomenon of nuclear physics discovered in the 1930s, in which matter exposed to
magnetic fields and radio waves was found to emit radio signals. In 1970, a physician and
researcher named Raymond Damadian noticed that malignant (cancerous) tissue gave off
different signals than normal body tissue. He applied for a patent for the first MRI scanning
device, which was in use clinically by the early 1980s. The early MRI scanners were crude,
but advances in digital computing and electronics led to their advancement over any other
technique for precise imaging, especially to discover tumours. MRI also has the
major advantage of not exposing patients to radiation.
Figure 1.9 Medical Imaging Techniques (a) The results of a CT scan of the head are shown as
successive transverse sections. (b) An MRI machine generates a magnetic field around a patient. (c) PET scans
use radiopharmaceuticals to create images of active blood flow and physiologic activity of the organ or organs
being targeted. (d) Ultrasound technology is used to monitor pregnancies because it is the least invasive of
imaging techniques and uses no electromagnetic radiation. (credit a: Akira Ohgaki/flickr; credit b: “Digital
Cate”/flickr; credit c: “Raziel”/Wikimedia Commons; credit d: “Isis”/Wikimedia Commons)

Drawbacks of MRI scans include their much higher cost, and patient discomfort with the
procedure. The MRI scanner subjects the patient to such powerful electromagnets that the
scan room must be shielded. The patient must be enclosed in a metal tube-like device for
the duration of the scan (see Figure 1.9b), sometimes as long as thirty minutes, which
can be uncomfortable and impractical for ill patients. The device is also so noisy that, even
with earplugs, patients can become anxious or even fearful. These problems have been
overcome somewhat with the development of “open” MRI scanning, which does not require
the patient to be entirely enclosed in the metal tube. Patients with iron-containing metallic
implants (internal sutures, some prosthetic devices, and so on) cannot undergo MRI
scanning because it can dislodge these implants.
Functional MRIs (fMRIs), which detect the concentration of blood flow in certain parts of the
body, are increasingly being used to study the activity in parts of the brain during various
body activities. This has helped scientists learn more about the locations of different brain
functions and more about brain abnormalities and diseases.
A patient undergoing an MRI is surrounded by a tube-shaped scanner. Watch this video to
learn more about MRIs. (http://openstaxcollege.org/l/MRI)

Positron emission tomography


Positron emission tomography (PET) is a medical imaging technique involving the use of so-
called radiopharmaceuticals, substances that emit radiation that is short-lived and therefore
relatively safe to administer to the body. Although the first PET scanner was introduced in
1961, it took 15 more years before radiopharmaceuticals were combined with
the technique and revolutionised its potential. The main advantage is that PET (see Figure
1.9c) can illustrate physiologic activity of the organ or organs being targeted, whereas CT
and MRI scans can only show static images. PET is widely used to diagnose a multitude of
conditions, such as heart disease, the spread of cancer, certain forms of infection, brain
abnormalities, bone disease, and thyroid disease.
PET relies on radioactive substances administered several minutes before the scan. Watch
this video (http://openstaxcollege.org/l/PET) to learn more about PET.

Ultrasonography
Ultrasonography is an imaging technique that uses the transmission of high-frequency
sound waves into the body to generate an echo signal that is converted by a computer into
a real-time image of anatomy and physiology (see Figure 1.9d). Ultrasonography is the least
invasive of all imaging techniques, and it is therefore used more freely in sensitive situations
such as pregnancy. The technology was first developed in the 1940s and 1950s.
Ultrasonography is used to study heart function, blood flow in the neck or extremities,
certain conditions such as gallbladder disease, and fetal growth and development. The main
disadvantages of ultrasonography are that the image quality is heavily operator-dependent
and that it is unable to penetrate bone and gas.
KEY TERMS
abdominopelvic cavity division of the anterior (ventral) cavity that houses the abdominal
and pelvic viscera
anatomical position standard reference position used for describing locations and
directions on the human body
anatomy science that studies the form and composition of the body’s structures
anterior describes the front or direction toward the front of the body; also referred to as
ventral
anterior cavity larger body cavity located anterior to the posterior (dorsal) body cavity;
includes the serous membrane-lined pleural cavities for the lungs, pericardial cavity for the
heart, and peritoneal cavity for the abdominal and pelvic organs; also referred to as ventral
cavity
caudal describes a position below or lower than another part of the body proper; near or
toward the tail (in humans, the coccyx, or lowest part of the spinal column); also referred to
as inferior
cell smallest independently functioning unit of all organisms; in animals, a cell contains
cytoplasm, composed of fluid and organelles
cytology the study of cells
computed tomography (CT) medical imaging technique in which a computer-enhanced
cross-sectional X-ray image is obtained
cranial describes a position above or higher than another part of the body proper; also
referred to as superior
cranial cavity division of the posterior (dorsal) cavity that houses the brain
deep describes a position farther from the surface of the body
distal describes a position farther from the point of attachment or the trunk of the body
dorsal describes the back or direction toward the back of the body; also referred to as
posterior
dorsal cavity posterior body cavity that houses the brain and spinal cord; also referred to
the posterior body cavity
frontal plane two-dimensional, vertical plane that divides the body or organ into anterior
and posterior portions
gross anatomy study of the larger structures of the body, typically with the unaided eye;
also referred to macroscopic anatomy
histology the study of tissues
homeostasis steady state of body systems that living organisms maintain
inferior describes a position below or lower than another part of the body proper; near or
toward the tail (in humans, the coccyx, or lowest part of the spinal column); also referred to
as caudal
lateral describes the side or direction toward the side of the body
magnetic resonance imaging (MRI) medical imaging technique in which a device generates
a magnetic field to obtain detailed sectional images of the internal structures of the body
medial describes the middle or direction toward the middle of the body
metabolism sum of all of the body’s chemical reactions
microscopic anatomy study of very small structures of the body using magnification
organ functionally distinct structure composed of two or more types of tissues
organ system group of organs that work together to carry out a particular function
organism living being that has a cellular structure and that can independently perform all
physiologic functions necessary for life
physiology science that studies the chemistry, biochemistry, and physics of the body’s
functions
plane imaginary two-dimensional surface that passes through the body
positron emission tomography (PET) medical imaging technique in which
radiopharmaceuticals are traced to reveal metabolic and physiological functions in tissues
posterior describes the back or direction toward the back of the body; also referred to as
dorsal
posterior cavity posterior body cavity that houses the brain and spinal cord; also referred to
as dorsal cavity
pressure force exerted by a substance in contact with another substance
prone face down
proximal describes a position nearer to the point of attachment or the trunk of the body
regional anatomy study of the structures that contribute to specific body regions
renewal process by which worn-out cells are replaced
sagittal plane two-dimensional, vertical plane that divides the body or organ into right and
left sides
section in anatomy, a single flat surface of a three-dimensional structure that has been cut
through
sensory information is collected from special detectors throughout the body, for example
the sense of taste is collected by detectors on the tongue
spinal cavity division of the dorsal cavity that houses the spinal cord; also referred to as
vertebral cavity
superficial describes a position nearer to the surface of the body
superior describes a position above or higher than another part of the body proper; also
referred to as cranial
supine face up
systemic anatomy study of the structures that contribute to specific body systems
thoracic cavity division of the anterior (ventral) cavity that houses the heart, lungs,
esophagus, and trachea
tissue group of similar or closely related cells that act together to perform a specific
function
transverse plane two-dimensional, horizontal plane that divides the body or organ into
superior and inferior portions
ultrasonography application of ultrasonic waves to visualise subcutaneous body structures
such as tendons and organs
ventral describes the front or direction toward the front of the body; also referred to as
anterior
ventral cavity larger body cavity located anterior to the posterior (dorsal) body cavity;
includes the serous membrane-lined pleural cavities for the lungs, pericardial cavity for the
heart, and peritoneal cavity for the abdominal and pelvic organs; also referred to as anterior
body cavity
X-ray form of high energy electromagnetic radiation with a short wavelength capable of
penetrating solids and ionising gases; used in medicine as a diagnostic aid to visualise body
structures such as bones
REVIEW QUESTIONS
Question 1.1
The smallest independently functioning unit of an organism is a (an)
____________________.
A cell
B molecule
C organ
D tissue

Question 1.2
The body system responsible for structural support and movement is the
____________________.
A cardiovascular system
B endocrine system
C muscular system
D skeletal system

Question 1.3
What is the position of the body when it is in the “normal anatomical position?”
A The person is prone with upper limbs, including palms, touching sides and lower limbs
touching at sides.
B The person is standing facing the observer, with upper limbs extended out at a ninety-
degree angle from the torso and lower limbs in a wide stance with feet pointing laterally
C The person is supine with upper limbs, including palms, touching sides and lower limbs
touching at sides.
D None of the above
Question 1.4
The lumbar region is ____________________.
A inferior to the gluteal region
B inferior to the umbilical region
C superior to the cervical region
D superior to the popliteal region

Question 1.5
The right foot and the right arm are ___ structures
A contralateral
B lateral
C medial
D ipsilateral

Question 1.6
Which of these structures is lateral to the rib cage?
A Spine
B Arms
C Legs
D Head

Question 1.7
If point A is closest to the midline in comparison to point B, then point A is ___ to point B.
A contralateral
B ipsilateral
C lateral
D medial
Question 1.8
Which of the following imaging techniques would be best to use to study the uptake of
nutrients by rapidly multiplying cancer cells?
A CT
B MRI
C PET
D ultrasonography

Question 1.9
Which of the following imaging studies can be used most safely during pregnancy?
A CT scans
B PET scans
C ultrasounds
D X-rays

Question 1.10
What are two major disadvantages of MRI scans?
A release of radiation and poor quality images
B high cost and the need for shielding from the magnetic signals
C can only view metabolically active tissues and inadequate availability of equipment
D release of radiation and the need for a patient to be confined to metal tube for up to 30
minutes
ANSWERS
Question 1.1
A

Question 1.2
D

Question 1.3
D

Question 1.4
D

Question 1.5
D

Question 1.6
A

Question 1.7
D

Question 1.8
C

Question 1.9
C

Question 1.10
B
2 | CELLS, TISSUES AND
HOMEOSTASIS

Figure 2.1 Schematic diagram showing the body systems involved in maintaining internal balance of the body

After studying this chapter, you will be able to:


The human body is composed of a variety of parts each working separately, but also working
together. At its healthiest, the human body can be compared to a perfectly functioning
complex piece of machinery.
This chapter begins with an explanation of the hierarchical organisation of the human body,
from the simplest (cell) to the most complex (organ system) level. It describes the basic
structure of the cell and the function of principal cell components, and then goes on to
introduce the basic tissue types of the human body. The chapter concludes with a discussion
on the concept of homeostasis and how the body works to maintain stable conditions for
itself.
2.1 | Structural Organization of the Human
Body
Before you begin to study the different structures and functions of the human body, it is
helpful to consider its basic architecture; that is, how its smallest parts are assembled into
larger structures.
Life processes of the human body are maintained at several levels of structural organisation.
These include the chemical, cellular, tissue, organ, organ system, and the organism level.
Higher levels of organisation are built from lower levels. Therefore, molecules combine to
form cells, cells combine to form tissues, tissues combine to form organs, organs combine to
form organ systems, and organ systems combine to form organisms.

Figure 2.2 Levels of structural organisation of the human body. The organisation of the body is often discussed
in terms of six distinct levels of increasing complexity, from the smallest chemical building blocks to a unique
human organism
It is convenient to consider the structures of the body in terms of fundamental levels of
organisation that increase in complexity: atoms, molecules, organelles, cells, tissues,
organs, organ systems.
The organisation of the body is often discussed in terms of six distinct levels of increasing
complexity, from the smallest chemical building blocks to the full human organism (Figure
2.2).
To study the chemical level of organisation, scientists consider the simplest building blocks
of matter: subatomic particles, atoms and molecules. 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. 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 in living things. Molecules are the
chemical building blocks of all body structures and are
termed biomolecules or macromolecules.
Cells contain four major biomolecules, including nucleic acids (DNA/RNA), proteins, lipids,
and carbohydrates. These each have distinct structural and biochemical properties that lend
to their individual functions within the cell. Collectively, they create the molecular players,
structures, and organelles that perform tasks within the cell. Table 2.1 outlines some of
these features for you just to provide an overview of their structure and function, however,
this will be explored in greater detail in later chapters.
A cell is the smallest independently functioning unit of a living organism. Even bacteria,
which are extremely small, independently-living organisms, have a cellular structure. Each
bacterium is a single cell. All living structures of human anatomy contain cells, and almost all
functions of human physiology are performed in cells or are initiated by cells.
A human cell typically consists of flexible membranes that enclose cytoplasm, a water-based
cellular fluid, together with a variety of tiny functioning units called organelles (Figure 2.3).

Figure 2.3 Simplified diagram of an animal cell showing cell membrane, cytoplasm, organelles (vacuole,
lysosome and mitochondria)

A tissue is a group of many similar cells (though sometimes composed of a few related
types) that work together to perform a specific function. An organ is an anatomically distinct
structure of the body composed of two or more tissue types. Each organ performs one or
more specific physiological functions. As we saw in Chapter 1, an organ system is a group of
organs that work together to perform major functions or meet physiological needs of the
body.
2.2 | An introduction to cell structure and
function
Ever wonder how paracetamol works to alleviate your headache? Do you know why the
caffeine in your coffee helps to kickstart your day? What causes mucosal build up in the
lungs of individuals with cystic fibrosis? What is the best way to treat a patient diagnosed
with diabetes? The answer to all of these questions begins with an understanding of cell
biology, in other words the structure and function of a cell.
Close your eyes and picture a brick wall. What is the basic building block of that wall? It is a
single brick, of course. Like a brick wall, your body is composed of basic building blocks, and
the building blocks of your body are cells. The cell is the smallest functional unit of a living
organism.
While some organisms are comprised of just a single cell (prokaryotes), humans
(eukaryotes) are much more complex and contain about 30 trillion cells that function
together to provide you with legs to walk to class, a digestive system to refuel after a long
night of studying, and brain power to learn subjects like anatomy and physiology.
Before we can tackle these big physiological questions about your body, it is important that
we take a closer look at the foundation of each organ system: the eukaryotic cell. Cells were
first recognised about 300 years ago by Robert Hooke, an English scientist who noted the
compartmentalisation of a cork (the outerbark of many trees) slice under the magnification
of a microscope. The sturdy cell walls of the plant specimen he observed were published
in Micrographia in 1665 (Figure 2.4). It was in this publication that Hooke coined the term
“cell” from the Latin word cellula (“small chamber”). Shortly thereafter, Antonie van
Leeuwenhoek, a Dutch microscopist, began visualising live microorganisms such as bacteria
which he called “animalcules”) along with specialised cell types in multicellular organisms,
including muscle fibres, red blood cells, and sperm. It was not until another 200 years
passed, however, that scientists began to understand the function of the cell. As
microscopes became more advanced, it became easier to study the behaviours within these
compartments, and by the mid-1800s, the pillars of the cell theory emerged. For an
entertaining and accurate birth of the cell theory, visit https://ed.ted.com/lessons/the-
wacky-history-of-cell-theory.
Over time, scientists have continued to build upon this foundation to establish a more
complete understanding of the functions that the cell carries out and the mechanisms by
which they take place.
Cell Theory

Figure 2.4 Robert Hooke published his discovery of plant cells in Micrographia in 1665. He coined the term
“cell” after the Latin root for small storage compartment. The cell is now understood to be the basic unit of all
living organisms. (Image courtesy of Alejandro Porto in the Public Domain)

Your body has many kinds of cells, each specialised for a specific purpose. Just as a home is
made from a variety of building materials, the human body is constructed from many cell
types. For example, epithelial cells protect the surface of the body and cover the organs and
body cavities within. Bone cells help to support and protect the body. Cells of the immune
system fight invading bacteria. Additionally, red blood cells carry oxygen throughout the
body. Each of these cell types plays a vital role during the growth, development, and day-to-
day maintenance of the body. In spite of their enormous variety, however, all cells share
certain fundamental characteristics.

How cells are studied


Microscopy
Cells vary in size. With few exceptions, individual cells are too small to be seen with the
naked eye, so scientists use microscopes to study them. A microscope is an instrument that
magnifies an object. Most images of cells are taken with a microscope and are called
micrographs.
Light Microscopes
To give you a sense of the size of a cell, a typical human red blood cell is about eight
millionths of a meter or eight micrometers (abbreviated as µm) in diameter; the head of a
pin is about two thousandths of a meter in diameter. That means that approximately 250
red blood cells could fit on the head of a pin. For another perspective on cell size,
visit HowBig? Interactive (http://openstaxcollege.org/l/cell_sizes2).
Light microscopes use two sets of lenses to magnify the image. Due to the manner in which
light travels through the lenses, this system of lenses produces an inverted image. Light
microscopes are advantageous for viewing living organisms, but since individual cells are
generally transparent, their components are not distinguishable unless they are coloured
with special stains. Staining, however, usually kills the cells.
Light microscopes (Figure 2.5a) commonly used in the undergraduate college laboratory
magnify up to approximately 400 times. Two parameters that are important in microscopy
are magnification and resolving power. Magnification is the degree of enlargement of an
object. Resolving power is the ability of a microscope to allow the eye to distinguish two
adjacent structures as separate; the higher the resolution, the closer those two objects can
be, and the better the clarity and detail of the image. When oil immersion lenses are used,
magnification is usually increased to 1,000 times for the study of smaller cells, like most
prokaryotic cells. Because light entering a specimen from below is focused onto the eye of
an observer, the specimen can be viewed using light microscopy. For this reason, for light to
pass through a specimen, the sample must be thin or translucent.

Figure 2.5 (a) Most student light microscopes used in laboratories can magnify cells up to approximately 400
times. (b) Dissecting microscopes have a lower magnification than light microscopes and are used to examine
larger objects, such as tissues.
A second type of microscope used in laboratories is the dissecting microscope (Figure 2.5b).
These microscopes have a lower magnification (20 to 80 times the object size) than light
microscopes and can provide a three-dimensional view of the specimen. Thick objects can
be examined with many components in focus at the same time. These microscopes are
designed to give a magnified and clear view of tissue structure as well as the anatomy of the
whole organism. Dissecting microscopes also have optics that correct the image so that it
appears as if being seen by the naked eye and not as an inverted image. The light
illuminating a sample under a dissecting microscope typically comes from above the sample,
but may also be directed from below.

Electron Microscopes
In contrast to light microscopes (Figure 2.6a), electron microscopes use a beam of electrons
instead of a beam of light. Cells and other structures are usually coated with a metal like
gold in preparation for viewing. There are 2 types of electron microscopes (EM): the
transmission EM and the scanning EM (Figure 2.6b).
Preparation of a specimen for viewing under an electron microscope will kill it; therefore,
live cells cannot be viewed using this type of microscopy. In addition, the electron beam
moves best in a vacuum, making it impossible to view living materials.

Figure 2.6 (a) Salmonella bacteria are viewed with a light microscope. (b) This scanning electron micrograph
shows Salmonella bacteria (in red) invading human cells. (credit a: modification of work by CDC, Armed Forces
Institute of Pathology, Charles N. Farmer; credit b: modification of work by Rocky Mountain Laboratories,
NIAID, NIH; scale-bar data from Matt Russell)

CAREERS IN ACTION
Cytotechnologist
Have you ever heard of a medical test called a Pap smear (Figure 2.7)? In this test, a doctor
takes a small sample of cells from the uterine cervix of a patient and sends it to a medical
lab where a cytotechnologist stains the cells and examines them for any changes that could
indicate cervical cancer or a microbial infection.
Cytotechnologists (cyto = cell) are professionals who study cells through microscopic
examinations and other laboratory tests. They are trained to determine which cellular
changes are within normal limits or are abnormal. Their focus is not limited to cervical cells;
they study cellular specimens that come from all organs. When they notice abnormalities,
they consult a pathologist, who is a medical doctor who can make a clinical diagnosis.
Cytotechnologists play vital roles in saving people’s lives. When abnormalities are
discovered early, a patient’s treatment can begin sooner, which usually increases the
chances of of the treatment being successful.

Figure 2.7 These uterine cervix cells, viewed through a light microscope, were obtained from a Pap smear.
Normal cells are on the left. The cells on the right are infected with human papillomavirus. (credit:
modification of work by Ed Uthman; scale-bar data from Matt Russell)

Parts of a cell
There are many types of cells, and all are grouped into one of two broad categories:
prokaryotic and eukaryotic. However, all cells share four common components:
1) a plasma membrane, or cell membrane, an outer covering that separates the cell’s
interior from its surrounding environment;
2) cytoplasm, consisting of a jelly-like region within the cell in which other cellular
components are found;
3) DNA, the genetic material of the cell; and
4) ribosomes, particles that synthesise proteins.
However, prokaryotes differ from eukaryotic cells in several ways. A prokaryotic cell is a
simple, single-celled (unicellular) organism that lacks a nucleus, or any other membrane-
bound organelle. We will shortly come to see that this is significantly different in
eukaryotes. Prokaryotic DNA is found in the central part of the cell: a darkened region called
the nucleoid (Figure 2.8).

Figure 2.8 This figure shows the generalised structure of a prokaryotic cell.

Unlike eukaryotes, bacteria have a cell wall (note this is different from the cell membrane)
which acts as an extra layer of protection, helps the cell maintain its shape, and prevents
dehydration. The capsule enables the cell to attach to surfaces in its environment. Some
prokaryotes have flagella, pili, or fimbriae. Flagella are used for locomotion, while most pili
are used to exchange genetic material during a type of reproduction called conjugation.

Components of Eukaryotic Cells


In nature, the relationship between form and function is apparent at all levels, including the
level of the cell, and this will become clear as we explore eukaryotic cells.
A eukaryotic cell (Figure 2.9) is a cell that has a membrane-bound nucleus and other
membrane-bound compartments or sacs, called organelles, which have specialised
functions. The word eukaryotic means “true kernel” or “true nucleus,” alluding to the
presence of the membrane-bound nucleus in these cells. The word “organelle” means “little
organ”; organelles have specialised cellular functions, just as the organs of your body have
specialised functions. At this point, it should be clear that eukaryotic cells have a more
complex structure than do prokaryotic cells.
The cell membrane (also called the plasma membrane) is the physical boundary between
the intracellular space (the inside of the cell) and the extracellular environment (the outside
of the cell). It acts almost like the "skin" of the cell, controlling the movement of substances
in and out of cells. The cell membrane is semi-permeable, allowing only select ions and
organic molecules to enter and/or leave the cell. Therefore, it is selectively permeable. The
cell membrane consists of two layers of phospholipids (a lipid bilayer) with embedded
proteins which have numerous functions (Figure 2.10). The structure and function of
the cell membrane will be discussed in greater detail in Chapter 5 The Cell Membrane.

Figure 2.9 Diagram showing the structure of the eukayotic cell


Figure 2.10 Diagram depicting the phospholipid bilayer with embedded proteins

The Cytoplasm
The cytoplasm comprises the contents of a cell between the plasma membrane and the
nuclear envelope (a structure to be discussed shortly). It is made up of organelles suspended
in the gel-like cytosol, the cytoskeleton, and various chemicals (Figure 2.9). Even though the
cytoplasm consists of 70 to 80 percent water, it has a semi-solid consistency, which comes
from the proteins within it. However, proteins are not the only organic molecules found in
the cytoplasm; glucose and other monosaccharides, polysaccharides, amino acids, nucleic
acids, fatty acids are also found there. In addition, ions of sodium, potassium, calcium, and
many other elements are also dissolved in the cytoplasm. Many chemical reactions,
essential for the functioning of the cell, take place in the cytoplasm.

The Cytoskeleton
If you were to remove all the organelles from a cell, would the plasma membrane and the
cytoplasm be the only components left? No. Within the cytoplasm, there would still be ions
and organic molecules, plus a network of protein fibres that helps to maintain the shape of
the cell, secures certain organelles in specific positions, allows cytoplasm and vesicles to
Figure 2.11 Microfilaments, intermediate filaments, and microtubules compose a cell’s cytoskeleton.

move within the cell, and enables unicellular organisms to move independently. Collectively,
this network of protein fibres is known as the cytoskeleton. There are three types of fibres
within the cytoskeleton: microfilaments, also known as actin filaments, intermediate
filaments, and microtubules (Figures 2.9 and 2.11):
The centrosome (Figure 2.9) is a region near the nucleus of animal cells that functions as a
microtubule-organising centre. It contains a pair of centrioles, two structures that lie
perpendicular to each other. Each centriole is a cylinder of nine triplets of
microtubules. The centrioles play a role in pulling the duplicated chromosomes to opposite
ends of the dividing cell.

Flagella and Cilia


Flagella (singular = flagellum) are long, hair-like structures that extend from the plasma
membrane, and are used to move an entire cell, (for example, sperm). When present, the
cell has just one flagellum or a few flagella. When cilia (singular = cilium) are present,
however, they are many in number and extend along the entire surface of the plasma
membrane. They are short, hair-like structures that are used to move substances along the
outer surface of the cell (for example, the cilia of cells lining the fallopian tubes that move
the ovum toward the uterus (Figure 2.17b), or cilia lining the cells of the respiratory tract
that move particulate matter that mucus has trapped, toward the throat.

The Endomembrane System


The endomembrane system (endo = within) is a group of membranes and organelles (Figure
2.15) in eukaryotic cells that work together to modify, package, and transport lipids and
proteins. It includes
Although not technically within the cell, the plasma membrane is included in the
endomembrane system because, as you will see, it interacts with the other
endomembranous organelles.

The Nucleus
Typically, the nucleus is the most prominent organelle in a cell (Figure 2.9).
The nucleus (plural = nuclei) houses the cell’s DNA in the form of chromatin and directs the
synthesis of ribosomes and proteins. Let us look at it in more detail (Figure 2.12).
The nuclear envelope is a double-membrane structure that constitutes the outermost
portion of the nucleus. Both the inner and outer membranes of the nuclear envelope are
phospholipid bilayers.
The nuclear envelope is punctuated with pores that control the passage of ions, molecules,
and RNA between the nucleoplasm and the cytoplasm. Like DNA, RNA is built from nucleic
acids. RNA is instrumental in converting the information stored in DNA into proteins, which
are in turn essential for the structure and function of cells. The process of how information
is carried from DNA to protein is discussed in greater detail in Chapter 6 The genetics of life.

Figure 2.12 The outermost boundary of the nucleus is the nuclear envelope. Notice that the nuclear envelope
consists of two phospholipid bilayers (membranes)—an outer membrane and an inner membrane—in contrast
to the plasma membrane (Figure), which consists of only one phospholipid bilayer. (credit: modification of
work by NIGMS, NIH)

To understand chromatin, it is helpful to first consider chromosomes. Chromosomes are


structures within the nucleus that are made up of DNA, the hereditary material, and
proteins. This combination of DNA and proteins is called chromatin. In eukaryotes,
chromosomes are linear structures. Every species has a specific number of chromosomes in
the nucleus of its body cells. For example, in humans, the chromosome number is 46,
whereas in fruit flies, the chromosome number is eight.
Chromosomes are only visible and distinguishable from one another when the cell is getting
ready to divide. When the cell is in the growth and maintenance phases of its life cycle, the
chromosomes resemble an unwound, jumbled bunch of threads.
We already know that the nucleus directs the synthesis of ribosomes, but how does it do
this? Some chromosomes have sections of DNA that encode ribosomal RNA. A darkly
staining area within the nucleus, called the nucleolus aggregates the ribosomal RNA with
associated proteins to assemble the ribosomal subunits that are then transported through
the nuclear pores into the cytoplasm; this process will be discussed in more detail again
in Chapter 6 The genetics of life.

The Endoplasmic Reticulum


The endoplasmic reticulum (ER) (Figure 2.9) is a series of interconnected membranous
tubules that collectively modify proteins and synthesise lipids. However, these two functions
are performed in separate areas of the endoplasmic reticulum: the rough endoplasmic
reticulum (RER) and the smooth endoplasmic reticulum (SER), respectively.
The hollow portion of the ER tubules is called the lumen. The membrane of the ER, which is
a phospholipid bilayer embedded with proteins, is continuous with the nuclear envelope.
The rough endoplasmic reticulum (RER) is so named because the ribosomes attached to its
cytoplasmic surface give it a studded appearance when viewed through an electron
microscope.
The ribosomes synthesise proteins while attached to the ER, resulting in transfer of their
newly synthesised proteins into the lumen of the RER where they undergo modifications
such as folding or addition of sugars. The RER also makes phospholipids for cell membranes.
If the phospholipids or modified proteins are not destined to stay in the RER, they will be
packaged within vesicles and transported from the RER by budding from the membrane
(Figure 2.14). Since the RER is engaged in modifying proteins that will be secreted from the
cell, it is abundant in cells that secrete proteins, such as the liver.
The smooth endoplasmic reticulum (SER) is continuous with the RER but has few or no
ribosomes on its cytoplasmic surface. The SER’s functions include synthesis of
carbohydrates, lipids (including phospholipids), and steroid hormones; detoxification of
medications and poisons; alcohol metabolism; and storage of calcium ions.
The Golgi Apparatus
We have already mentioned that vesicles can bud from the ER, but where do the vesicles
go? Before reaching their final destination, the lipids or proteins within the transport
vesicles need to be sorted, packaged, and tagged so that they wind up in the right place. The
sorting, tagging, packaging, and distribution of lipids and proteins take place in the Golgi
apparatus (also called the Golgi body), a series of flattened membranous sacs (Figures 2.9,
2.13 and 2.15).

Figure 2.13 The Golgi apparatus in this transmission electron micrograph of a white blood cell is visible as a
stack of semicircular flattened rings in the lower portion of this image. Several vesicles can be seen near the
Golgi apparatus. (credit: modification of work by Louisa Howard; scale-bar data from Matt Russell)

The Golgi apparatus has a receiving face near the endoplasmic reticulum and a releasing
face on the side away from the ER, toward the cell membrane. The transport vesicles that
form from the ER travel to the receiving face, fuse with it, and empty their contents into the
lumen of the Golgi apparatus. As the proteins and lipids travel through the Golgi, they
undergo further modifications. The most frequent modification is the addition of short
chains of sugar molecules. The newly modified proteins and lipids are then tagged with
small molecular groups to enable them to be routed to their proper destinations.
Finally, the modified and tagged proteins are packaged into vesicles that bud from the
opposite face of the Golgi (Figure 2.15). While some of these vesicles, transport vesicles,
deposit their contents into other parts of the cell where they will be used, others, secretory
vesicles, fuse with the plasma membrane and release their contents outside the cell.
The amount of Golgi in different cell types again illustrates that form follows function within
cells. Cells that engage in a great deal of secretory activity (such as cells of the salivary
glands that secrete digestive enzymes or cells of the immune system that secrete
antibodies) have an abundant number of Golgi.
Lysosomes
In animal cells, the lysosomes are the cell’s “garbage disposal.” Digestive enzymes within
the lysosomes aid the breakdown of proteins, polysaccharides, lipids, nucleic acids, and
even worn-out organelles. These enzymes are active at a much lower pH (more acidic) than
those located in the cytoplasm. Many reactions that take place in the cytoplasm could not
occur at a low pH, thus the advantage of compartmentalising the eukaryotic cell into
organelles is apparent.

Figure 2.14 A macrophage has phagocytised a potentially pathogenic bacterium into a vesicle, which then
fuses with a lysosome within the cell so that the pathogen can be destroyed. Other organelles are present in
the cell, but for simplicity, are not shown.

Lysosomes also use their enzymes to destroy disease-causing organisms that might enter
the cell. A good example of this occurs in a group of white blood cells called macrophages,
which are part of your body’s immune system. In a process known as phagocytosis (Figure
2.14), a section of the plasma membrane of the macrophage invaginates (folds in) and
engulfs a pathogen. The invaginated section, with the pathogen inside, then pinches itself
off from the plasma membrane and becomes a vesicle. The vesicle fuses with a lysosome.
The lysosome’s enzymes then destroy the pathogen.

Vesicles and Vacuoles


Vesicles and vacuoles are membrane-bound sacs that function in storage and transport.
Vacuoles are somewhat larger than vesicles, and the membrane of a vacuole does not fuse
with the membranes of other cellular components. Vesicles can fuse with other membranes
within the cell system.
Figure 2.15 The endomembrane system works to modify, package, and transport lipids and proteins. (credit:
modification of work by Magnus Manske)

Ribosomes
Ribosomes are the cellular structures responsible for protein synthesis. Electron microscopy
has shown that ribosomes consist of large and small subunits. Ribosomes are enzyme
complexes that are responsible for protein synthesis.
Because protein synthesis is essential for all cells, ribosomes are found in practically every
cell. They are particularly abundant in immature red blood cells for the synthesis of
hemoglobin, which functions in the transport of oxygen throughout the body.
Mitochondria
Mitochondria (singular = mitochondrion) are often called the “powerhouses” or “energy
factories” of a cell because they are responsible for making adenosine triphosphate (ATP),
the cell’s main energy-carrying molecule. The formation of ATP from the breakdown of
glucose is known as cellular respiration. Mitochondria are oval-shaped, double-membrane
organelles (Figure 2.16) that have their own ribosomes and DNA. Each membrane is a
phospholipid bilayer embedded with proteins. The inner layer has folds called cristae, which
increase the surface area of the inner membrane. The area surrounded by the folds is called
the mitochondrial matrix. The cristae and the matrix have different roles in cellular
respiration.
In keeping with our theme of form following function, it is important to point out that
muscle cells have a very high concentration of mitochondria because muscle cells need a lot
of energy to contract.

Figure 2.16 This transmission electron micrograph shows a mitochondrion as viewed with an electron
microscope. Notice the inner and outer membranes, the cristae, and the mitochondrial matrix. (credit:
modification of work by Matthew Britton; scale-bar data from Matt Russell)

Peroxisomes
Peroxisomes (Figure 2.9) are small, round organelles enclosed by single membranes. They
carry out oxidation reactions that break down fatty acids and amino acids. They also
detoxify many poisons that may enter the body. Alcohol is detoxified by peroxisomes in liver
cells. A byproduct of these oxidation reactions is hydrogen peroxide, H2O2, which is
contained within the peroxisomes to prevent the chemical from causing damage to cellular
components outside of the organelle. Hydrogen peroxide is safely broken down by
peroxisomal enzymes into water and oxygen.

2.3 | Tissues of the human body


A tissue is a group of connected cells that have a similar function within an organism. There
are four basic types of tissues in the bodies of all animals including the human body. These
make up all the organs, structures, and other contents of the body. Figure 2.17 shows an
example of each tissue type. The four basic types of tissues are muscle, epithelial, nervous,
and connective.

(a) skeletal muscle through the transmission electron microscope

(b) ciliated epithelial tissue of the oviduct through the light microscope
(c) nervous tissue through the light microscope

(d) fibrocartilage (found in skeletal joints) showing dense connective tissue through the light
microscope
Figure 2.17 Micrographs of examples of the 4 different types of tissue of the human body

Muscle tissue encompasses not only the muscles, such as those in our legs or fingers that
we actively control, but also the tissue that forms most of our internal organs. There are
three types of muscle tissue: skeletal, cardiac, and smooth. Skeletal muscle tissue forms
what we think of as our muscles; it is attached to our bones by our tendons and can be
relaxed or contracted voluntarily. Similar in structure to skeletal muscle, cardiac muscle is
found exclusively in the walls of the heart. The major difference, however, is that cardiac
muscle is involuntary and cannot be actively controlled. Similarly, smooth muscle, which
forms the muscle layers in internal organs such as the digestive tract and bladder, is an
involuntary tissue. Smooth muscle tissue controls slow involuntary movements such as
stomach wall contractions, and the contractions of arteries to regulate blood flow.
Epithelial tissue is made up of a layer or layers of tightly packed cells that line the surfaces
of the body. The largest example of epithelial tissue (also the largest organ in the human
body) is the skin. Mammalian skin consists of stratified epithelium, which has several layers
of cells. The outermost layers of cells, called squamous cells, are flat plate-like cells, while
the deeper layers are roughly cube shaped and called cuboidal cells. Epithelial tissue has
multiple functions, but it serves primarily to protect, absorb, and secrete. Our skin organ
covers our entire body and protects underlying tissues from bacteria, chemicals, and other
injury. Epithelial cells also line the oviduct as well as the small intestine where they absorb
nutrients, and similar cells in the glands secrete enzymes and hormones.
Nervous tissue is made up of the neurons and glia which are specialised cells that form the
nervous system, including the brain and spinal cord. Neurons are especially responsive to
stimuli, allowing nervous tissue to transmit stimuli from the brain to the body extremely
rapidly.
Connective tissue connects, supports, or separates other tissues and organs. Connective
tissue proper, a form of connective tissue, can be either loose or dense. Adipose tissue, or
fat, is loose connective tissue, while tendons and ligaments, composed of collagen, are
examples of dense connective tissue. Other forms of connective tissue include blood (fluid
connective tissue) and cartilage and bone (both forms of supporting connective tissue).

Over the past 10 years, mitochondrial defects have been implicated in a variety of degenerative
diseases, aging, and cancer. However, the pathophysiology of mitochondrial diseases remains
perplexing. One of the first indications that mitochondria may play a role in pathogenesis was
reported nearly 40 years ago. A patient with hyper-metabolism had skeletal muscle that contained
large numbers of abnormal mitochondria, a condition now known as mitochondrial myopathy.
Patients with mitochondrial myopathy suffer from muscle weakness and exercise intolerance, stroke
like episodes, dementia, blindness, deafness, heart failure, arrhythmias, and movement disorders.
Mitochondrial myopathies encompass an extraordinary assemblage of clinical problems, commonly
involving tissues that have high-energy requirements such as heart, muscle, and the renal and
endocrine systems.

2.4 | Homeostasis
To function properly, cells require appropriate conditions such as proper temperature, pH,
and concentrations of diverse chemicals. These conditions may, however, change from one
moment to the next. Organisms are able to maintain internal conditions within a narrow
range almost constantly, despite environmental changes, through a process
called homeostasis or “steady state”— the ability of an organism to maintain constant
internal conditions. For example, many organisms regulate their body temperature in a
process known as thermoregulation. Organisms that live in cold climates, such as the polar
bear (Figure 2.18), have body structures that help them withstand low temperatures and
conserve body heat. In hot climates, organisms have methods (such as perspiration in
humans or panting in dogs) that help them to shed excess body heat.

Figure 2.18 Polar bears and other mammals living in ice-covered regions maintain their body temperature by
generating heat and reducing heat loss through thick fur and a dense layer of fat under their skin. (credit:
"longhorndave"/Flickr)

Eleven distinct body systems exist in the human body. These systems can function alone,
however, they need to work together to make a living organism.
Each body system contributes to the homeostasis of other systems and of the entire
organism. No system of the body works in isolation, and the well-being of the person
depends upon the well-being of all the interacting body systems. A disruption within one
system generally has consequences for several other body systems.
Maintaining homeostasis requires that the body continuously monitor its internal
conditions. Each physiological parameter or condition (such as body temperature, blood
pressure, levels of certain nutrients) 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 centres 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 goes on throughout the body at all times, 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 (Figure 2.19a). A sensor, also
referred to as a receptor, is a component of a feedback system that monitors a physiological
value. This value is reported to the control centre. The control centre 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 centre 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.

Figure 2.19a Negative feedback loop works to oppose the change in the physiological parameter and bring it
back to normal (setpoint). Note the term control represents the control centre.
If a raise in body temperature (stimulus) is detected (receptor), a signal will cause the brain
to maintain homeostasis (response). Once the body temperature returns to normal,
negative feedback will cause the response to end. This sequence of stimulus-receptor-
signal-response is used throughout the body to maintain homeostasis.

Figure 2.19b Body temperature is regulated by negative feedback

In order to set the negative feedback 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, humans have a temperature regulation feedback system
that works by promoting either heat loss or heat gain (Figure 2.19b). When the brain’s
temperature regulation centre 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 centre.” This stimulation has three major effects:
In contrast, activation of the brain’s heat-gain centre 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; in addition, the brain also triggers the release of chemicals that ultimately
release heat, metabolism and heat production.

Positive Feedback
Some processes in the body are regulated by 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 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 (Figure
2.20).

Figure 2.20 Positive Feedback Loop. Normal childbirth is driven by a positive feedback loop. A positive
feedback loop results in a change in the body’s status, rather than a return to homeostasis.
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 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 centres 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.

KEY TERMS
amino acids basic units that make up proteins
atoms the smallest part of an element that exist either alone or in combination with other
biomolecules organic molecules in living organisms, especially macromolecules like
protein, carbohydrates, fatty acids, nucleic acids
cell biology the study of living organisms at the cellular level
cell theory scientific theory that describes the properties of a cell
cell membrane membrane surrounding all animal cells, composed of a phospholipid
bilayer interspersed with various molecules; also known as plasma membrane
chromatin a complex of DNA and proteins that forms chromosomes within the nucleus of
eukaryotic cells
control centre compares values to their normal range; deviations cause the activation of an
effector
cytoplasm internal material between the cell membrane and nucleus of a cell, mainly
consisting of a water-based fluid called cytosol, within which are all the other organelles
and cellular solute and suspended materials
cytoskeleton a network of protein fibres that helps to maintain the shape of the cell,
secures certain organelles in specific positions, allows cytoplasm and vesicles to move within
the cell, and enables unicellular organisms to move independently.
DNA is the material that carries all the information about how a living thing will look and
function
development changes an organism goes through during its life
differentiation process by which unspecialised cells become specialised in structure and
function
electron negatively charged sub-atomic particle
element substance that cannot be broken down into simpler substances by chemical means
endomembrane system the group of organelles and membranes in eukaryotic cells that
work together to modify, package, and transport lipids and proteins
endoplasmic reticulum (ER) a series of interconnected membranous structures within
eukaryotic cells that collectively modify proteins and synthesise lipids
effector organ that can cause a change in a value
enzyme substance (protein) produced by a living organism which acts as a catalyst to bring
about a specific biochemical reaction
epithelial cells arranged in one or more layers to form a membrane that covers or lines
body parts
eukaryotes multicellular organisms that are larger and more complex than prokaryotes
fatty acids are the building blocks of the fat in our bodies and in the food we eat
glia or glial cells most abundant cell type in the nervous system. Neurons would not work
properly without glia
golgi apparatus a eukaryotic organelle made up of a series of stacked membranes that
sorts, tags, and packages lipids and proteins for distribution
homeostasis any self-regulating process by which biological systems tend to maintain
stability while adjusting to conditions that are optimal for survival
ions an atom or molecule with a net electric charge due to the loss or gain of one or more
electrons
lipids made up of the same elements as carbohydrates: carbon, hydrogen, and oxygen.
However, lipids tend to contain many more hydrogen atoms than oxygen atoms
lysosome membrane-bound cellular organelle originating from the Golgi apparatus and
containing digestive enzymes
macromoles biomolecules like protein, fatty acids, carbohydrates and nucleic acids
macrophage a type of white blood cell that ingests foreign material
mitochondrion one of the cellular organelles bound by a double lipid bilayer that function
primarily in the production of cellular energy (ATP)
molecules form when two or more atoms form chemical bonds with each other
negative feedback homeostatic mechanism that tends to stabilise an upset in the body’s
physiological condition by preventing an excessive response to a stimulus
neuron specialised cell transmitting nerve impulses
neutron the particle in the atomic nucleus with a mass of 1 and charge of zero
normal range range of values around the set point that do not cause a reaction by the
control centre
nucleic acids the building blocks of DNA are nucleic acids,. Nucleic acids are complex
organic substances.
nuclear envelope the double-membrane structure that constitutes the outermost portion
of the nucleus
nucleolus the darkly staining body within the nucleus that is responsible for assembling
ribosomal subunits
nucleus cell’s central organelle; contains the cell’s DNA
organelles means "small organs" to denote membrane bound structures found floating in
the cytoplasm of a cell
organic molecules composed of carbon atoms attached to other atoms
oxytocin molecule made of amino acids; it is produced in the brain and released from the
pituitary gland
peroxisome a small, round organelle that contains hydrogen peroxide, oxidises fatty acids
and amino acids, and detoxifies many poisons
pH a measure of how acidic a solution is
phagocytosis a process that takes macromolecules that the cell needs from the extracellular
fluid; a variation of endocytosis
Phospholipid lipid molecule(s) attached to a phosphate molecule (one phosphorous
combined with four oxygen atoms)
pinocytosis a process that takes solutes that the cell needs from the extracellular fluid; a
variation of endocytosis
pituitary gland secretes hormones that controls other glands and many body funtions,
including growth
plasma membrane a phospholipid bilayer with embedded (integral) or attached
(peripheral) proteins that separates the internal contents of the cell from its surrounding
environment
polysaccharides a carbohydrate (e.g. starch) whose molecules consist of a number of sugar
molecules bonded together
positive feedback mechanism that intensifies a change in the body’s physiological
condition in response to a stimulus
prokaryote single celled organism lacking a true membrane-bound nucleus, for example,
bacteria
proton component of the atomic nucleus with a mass of 1 and charge of +1
rough endoplasmic reticulum (RER) the region of the endoplasmic reticulum that is
studded with ribosomes and engages in protein modification
ribosome a cellular organelle that carries out protein synthesis
RNA nucleic acid similar to DNA but containing ribose as the sugar molecule rather than
deoxyribose
selectively permeable or semipermeable the characteristic of a membrane that allows
some substances through but not others
set point ideal value for a physiological parameter; the level or small range within which a
physiological parameter such as blood pressure is stable and optimally healthful, that is,
within its parameters of homeostasis
sensor (also, receptor) reports a monitored physiological value to the control centre
smooth endoplasmic reticulum (SER) the region of the endoplasmic reticulum that has few
or no ribosomes on its cytoplasmic surface and synthesises carbohydrates, lipids, and
steroid hormones; detoxifies chemicals like pesticides, preservatives, medications,
and environmental pollutants, and stores calcium ion
stratified epithelium epithelial cells in layers
tissue a group of connected cells in an animal or plant that are similar to each other and
have the same purpose
vesicle a small, membrane-bound sac that functions in cellular storage and transport; its
membrane is capable of fusing with the plasma membrane and the membranes of the
endoplasmic reticulum and Golgi apparatus
vacuole a membrane-bound sac, somewhat larger than a vesicle, that functions in cellular
storage and transport
REVIEW QUESTIONS
Question 2.1
In eukaryotic cells, DNA is found in the
A nucleolus
B golgi apparatus
C rough endoplasmic reticulum
D nucleus

Question 2.2
Which of the following is not a component of the endomembrane system?
A Mitochondrion
B Golgi apparatus
C endoplasmic reticulum
D lysosome

Question 2.3
The rough endoplasmic reticulum has a dotted appearance when viewed through the transmission
electron microscope because it is covered with

A DNA

B proteins

C ribosomes

D chromosomes
Question 2.4
The body responds to overheating on a hot day by
A increasing the output of the sweat glands.
B causing constriction of blood vessels in the skin.
C making breathing slow and shallow.
D causing dilation (opposite of constriction) of blood vessels in the skin.

Question 2.5
AJ is stuck in a car during a bitterly cold blizzard. Her body responds to the cold by
A increasing blood flow to her skin
B becoming lethargic to conserve heat
C breaking down stored energy
D significantly increasing blood oxygen levels

Question 2.6
Which of the following is an example of a normal physiological process that uses a positive
feedback loop?
A Blood pressure regulation
B Child birth
C regulation of fluid balance
D temperature regulation

Question 2.7
If blood pressure goes up, the negative feedback loop kicks in to
A decrease blood pressure to normal
B further increase blood pressure
C decrease blood pressure significantly below normal
D induce sleep
ANSWERS
Question 2.1
D

Question 2.2
A

Question 2.3
C

Question 2.4
A, D

Question 2.5
C

Question 2.6
B

Question 2.7
A
3| THE BASIC ELEMENTS OF LIFE

Figure 3.1 Human DNA. Human DNA is described as a double helix that resembles a molecular spiral staircase.
DNA is made up of several elements which are represented here by different coloured plastic balls.

After studying this chapter, you will be able to:


• Describe the fundamental composition of matter
• Identify the three subatomic particles
• Explain the relationship between an atom’s number of electrons and its relative
stability
• Distinguish between ionic bonds, covalent bonds, and hydrogen bonds
The smallest, most fundamental material components of the human body are basic
chemical elements. The elements carbon, hydrogen, nitrogen, oxygen, sulfur, and
phosphorus are the key building blocks of the chemicals found in living things. They form
the carbohydrates, nucleic acids, proteins, lipids and water that are the fundamental
molecular components of all organisms. The properties of these molecules (called
biomolecules or macromolecules) are determined by how their constituent elements
interact with each other.
This chapter begins by examining elements by first exploring the structure of the basic unit
(atom) of an element. It goes on to discuss how the structure of the atom determines the
characteristics of an element. The chapter then explores the concepts behind chemical
bonding between elements as this bonding is necessary for the formation of the
biomolecules necessary for life.

3.1 | Elements and Atoms: The Building


Blocks of Matter
The substance of the universe—from a grain of sand to a star—is called matter. Scientists
define matter as anything that occupies space and has mass. An object’s mass and its weight
are related concepts, but not quite the same. An object’s mass is the amount of matter
contained in the object, and the object’s mass is the same whether that object is on Earth or
in the zero-gravity environment of outer space. An object’s weight, on the other hand, is its
mass as affected by the pull of gravity. Where gravity strongly pulls on an object’s mass its
weight is greater than it is where gravity is less strong. An object of a certain mass weighs
less on the moon, for example, than it does on Earth because the gravity of the moon is less
than that of Earth. In other words, weight is variable, and is influenced by gravity. A piece of
cheese that weighs a kilogram on Earth weighs only a few grams on the moon.

Elements and Compounds


All matter in the natural world is composed of one or more of the 92 fundamental
substances called elements. An element is a pure substance that is distinguished from all
other matter by the fact that it cannot be created or broken down by ordinary chemical
means. While your body can assemble many of the chemical compounds needed for life
from their constituent elements, it cannot make elements. They must come from the
environment. A familiar example of an element that you must take in is calcium (Ca++).
Calcium is essential to the human body; it is absorbed and used for a number of processes,
including strengthening bones. When you consume dairy products your digestive system
breaks down the food into components small enough to cross into the bloodstream. Among
these is calcium, which, because it is an element, cannot be broken down further. The
elemental calcium in cheese, therefore, is the same as the calcium that forms your bones.
Some other elements you might be familiar with are oxygen, sodium, and iron. The
elements in the human body are shown in Figure 3.2, beginning with the most abundant:
oxygen (O), carbon (C), hydrogen (H), and nitrogen (N). Each element’s name can be
replaced by a one- or two-letter symbol; you will become familiar with some of these during
this course. All the elements in your body are derived from the foods you eat and the air you
breathe.

Figure 3.2 Elements of the Human Body The main elements that compose the human body are shown from
most abundant to least abundant.

In nature, elements rarely occur alone. Instead, they combine to form compounds.
A compound is a substance composed of two or more elements joined by chemical bonds.
For example, the compound glucose is an important body fuel. It is always composed of the
same three elements: carbon, hydrogen, and oxygen. Moreover, the elements that make up
any given compound always occur in the same relative amounts. In glucose, there are
always six carbon and six oxygen units for every twelve hydrogen units. But what, exactly,
are these "units"?

Atoms and Subatomic Particles


An atom is the smallest quantity ("unit") of an element that retains the unique properties of
that element. In other words, an atom of hydrogen is a unit of hydrogen—the smallest
amount of hydrogen that can exist. As you might guess, atoms are almost unfathomably
small; to get an idea of how small, the full-stop at the end of this sentence is millions of
atoms wide.
Atomic Structure and Energy
Atoms are made up of even smaller subatomic particles: the main ones are:
• the proton
• the neutron
• the electron.
All atoms contain protons, electrons, and neutrons (Figure 3.3) shows a model of the helium
atom). The only exception is hydrogen (H), which is made of one proton and one electron.
A proton is a positively charged particle that resides in the nucleus (the core of the atom) of
an atom and has a mass of 1 and a charge of +1. An electron is a negatively charged particle
that travels in the space around the nucleus in an orbit. In other words, it resides outside of
the nucleus. It has a negligible mass and has a charge of –1.
Neutrons, like protons, reside in the nucleus of an atom. They have a mass of 1 and no
charge. The positive (protons) and negative (electrons) charges balance each other in a
neutral atom, which has a net zero charge.

Figure 3.3 Atoms are made up of protons and neutrons located within the nucleus, and electrons surrounding
the nucleus

As stated earlier, each element has its own unique properties. Each contains a different
number of protons and neutrons, giving it its own atomic number and mass number.
The atomic number of an element is equal to the number of protons that element contains.
The mass number, or atomic mass, is the number of protons plus the number of neutrons of
that element. Therefore, it is possible to determine the number of neutrons by subtracting
the atomic number from the mass number. For example, sodium (Na) has a mass number of
as 23 and 12 neutrons, making its atomic number 11.
Mass number = number protons (atomic number) + number of neutrons
These numbers provide information about the elements and how they will react when
combined. Different elements have different melting and boiling points, and are in different
states (liquid, solid, or gas) at room temperature. They also combine in different ways. Some
form specific types of bonds, whereas others do not. How they combine is based on the
number of electrons present. Because of these characteristics, the elements are arranged
into the periodic table of elements, a chart of the elements that includes the atomic
number and relative atomic mass of each element. The periodic table also provides key
information about the properties of elements (Figure 3.4)—often indicated by colour-
coding. The arrangement of the table also shows how the electrons in each element are
organised and provides important details about how atoms will react with each other to
form molecules.

Figure 3.4 Arranged in columns and rows based on the characteristics of the elements, the periodic table
provides key information about the elements and how they might interact with each other to form molecules.
Most periodic tables provide a key or legend to the information they contain.
Visit this website (http://openstaxcollege.org/l/ptable) to view the periodic table. In the
periodic table of the elements, elements in a single column have the same number of
electrons that can participate in a chemical reaction. These electrons are known as “valence
electrons.” For example, the elements in the first column all have a single valence electron,
an electron that can be “donated” in a chemical reaction with another atom.

Isotopes
Isotopes are different forms of the same element that have the same number of protons,
but a different number of neutrons (Figure 3.5). Some elements, such as carbon, potassium,
and uranium, have naturally occurring isotopes. Carbon-12, the most common isotope of
carbon, contains six protons and six neutrons. Therefore, it has a mass number of 12 (six
protons added to six neutrons) and an atomic number of 6 (which makes it carbon). Carbon-
14 contains six protons and eight neutrons. Therefore, it has a mass number of 14 (six
protons added to eight neutrons) and an atomic number of 6, meaning it is still the element
carbon. These two alternate forms of carbon are isotopes.

Question 3.1
How many neutrons does (K) potassium-39?

Question 3.2
How many neutrons does potassium-40 have?

Figure 3.5 Isotopes of Hydrogen. Protium, designated 1H, has one proton and no neutrons. It is by far the most
abundant isotope of hydrogen in nature. Deuterium, designated 2H, has one proton and one neutron.
Tritium, designated 3H, has two neutrons.

An isotope that contains more than the usual number of neutrons is referred to as a heavy
isotope. An example is 14C. Heavy isotopes tend to be unstable, and unstable isotopes are
radioactive. A radioactive isotope is an isotope whose nucleus readily decays, giving off
subatomic particles and electromagnetic energy. Different radioactive isotopes (also called
radioisotopes) differ in their half-life, the time it takes for half of any size sample of an
isotope to decay. For example, the half-life of tritium—a radioisotope of hydrogen—is about
12 years, indicating it takes 12 years for half of the tritium nuclei in a sample to decay.
Excessive exposure to radioactive isotopes can damage human cells and even cause cancer
and birth defects, but when exposure is controlled, some radioactive isotopes can be useful
in medicine. For more information, see the Career Connections.

Interventional Radiologist
The controlled use of radioisotopes has advanced medical diagnosis and treatment of
disease. Interventional radiologists are physicians who treat disease by using minimally
invasive techniques involving radiation. Many conditions that could once only be treated
with a lengthy and traumatic operation can now be treated non-surgically, reducing the
cost, pain, length of hospital stay, and recovery time for patients. For example, in the past,
the only options for a patient with one or more tumours in the liver were surgery and
chemotherapy (the administration of drugs to treat cancer). Some liver tumours, however,
are difficult to access surgically, and others could require the surgeon to remove too much
of the liver. Moreover, chemotherapy is highly toxic to the liver, and certain tumours do not
respond well to it anyway. In some such cases, an interventional radiologist can treat the
tumours by disrupting their blood supply, which they need if they are to continue to grow.
In this procedure, called radioembolisation, the radiologist accesses the liver with a fine
needle, threaded through one of the patient’s blood vessels. The radiologist then inserts
tiny radioactive “seeds” into the blood vessels that supply the tumours. In the days and
weeks following the procedure, the radiation emitted from the seeds destroys the vessels
and directly kills the tumour cells in the vicinity of the treatment.

The Behaviour of Electrons


In the human body, atoms do not exist as independent entities. Rather, they are constantly
reacting with other atoms to form and to break down more complex substances. To fully
understand anatomy and physiology you must grasp how atoms participate in such
reactions. The key is understanding the behaviour of electrons.
Although electrons do not follow rigid orbits a set distance away from the atom’s nucleus,
they do tend to stay within certain regions of space called electron shells. An electron
shell is a layer of electrons that encircle the nucleus.
The atoms of the elements found in the human body have from one to five electron shells,
and all electron shells hold eight electrons except the first shell, which can only hold two.
This configuration of electron shells is the same for all atoms.
The precise number of shells depends on the number of electrons in the atom. Hydrogen
and helium have just one and two electrons, respectively. If you take a look at the periodic
table of the elements (Figure 3.4), you will notice that hydrogen and helium are placed
alone on either sides of the top row; they are the only elements that have just one electron
shell (Figure 3.6 (a) and (b)). A second shell is necessary to hold the electrons in all elements
larger than hydrogen and helium.

(b)

(c) (d)

Figure 3.6 Electron Shells Electrons orbit the atomic nucleus at distinct levels of energy called electron
shells. (a) With one electron, hydrogen only half-fills its electron shell. Helium also has a single shell, but its
two electrons completely fill it. (b) The electrons of carbon completely fill its first electron shell, but only half-
fills its second. (c) Neon, an element that does not occur in the body, has 10 electrons, filling both of its
electron shells.

Lithium (Li), whose atomic number is 3, has three electrons. Two of these fill the first
electron shell, and the third spills over into a second shell. The second electron shell can
accommodate as many as eight electrons. Carbon, with its six electrons, entirely fills its first
shell, and half-fills its second (Figure 3.6 (c)). With ten electrons, neon (Ne) entirely fills its
two electron shells (Figure 3.6 (d)). Again, a look at the periodic table reveals that all of the
elements in the second row, from lithium to neon, have just two electron shells. Atoms with
more than ten electrons require more than two shells. These elements occupy the third and
subsequent rows of the periodic table. It is useful to remember that the first, second and
third electron shells can hold 2, 8 and 8 electrons respectively.
The factor that most strongly governs the tendency of an atom to participate in chemical
reactions is the number of electrons in its valence shell. A valence shell is an atom’s
outermost electron shell. If the valence shell is full, the atom is stable; meaning its electrons
are unlikely to be pulled away from the nucleus by the electrical charge of other atoms. If
the valence shell is not full, the atom is reactive; meaning it will tend to react with other
atoms in ways that make the valence shell full. Consider hydrogen, with its one electron
only half-filling its valence shell. This single electron is likely to be drawn into relationships
with the atoms of other elements, so that hydrogen’s single valence shell can be stabilised.
All atoms (except hydrogen and helium with their single electron shells) are most stable
when there are exactly eight electrons in their valence shell. For example, oxygen, with six
electrons in its valence shell, is likely to react with other atoms in a way that results in the
addition of two electrons to oxygen’s valence shell, bringing the number to eight.
In nature, atoms of one element tend to join with atoms of other elements in characteristic
ways. For example, carbon commonly fills its valence shell by linking up with four atoms of
hydrogen. In so doing, the two elements form the simplest of organic molecules, methane,
which also is one of the most abundant and stable carbon-containing compounds on Earth.
Another example is water; oxygen needs two electrons to fill its valence shell. It commonly
interacts with two atoms of hydrogen, forming H2O. Incidentally, the name “hydrogen”
reflects its contribution to water (hydro- = “water”; -gen = “maker”). Thus, hydrogen is the
“water maker.”

3.2 | Chemical Bonds


Atoms separated by a great distance cannot link; rather, they must come close enough for
the electrons in their valence shells to interact. But do atoms ever actually touch one
another? Most physicists would say no, because the negatively charged electrons in their
valence shells repel one another. No force within the human body—or anywhere in the
natural world—is strong enough to overcome this electrical repulsion. So when you read
about atoms linking together or colliding, bear in mind that the atoms are not merging in a
physical sense.
Instead, atoms link by forming a chemical bond. A bond is a weak or strong electrical
attraction that holds atoms in the same vicinity. The new grouping is typically more stable—
less likely to react again—than its component atoms were when they were separate. A more
or less stable grouping of two or more atoms held together by chemical bonds is called
a molecule. The bonded atoms may be of the same element, as in the case of H2, which is
called molecular hydrogen or hydrogen gas. When a molecule is made up of two or more
atoms of different elements, it is called a chemical compound. Thus, a unit of water, or H2O,
is a compound, as is a single molecule of the gas methane, or CH4.
Three types of chemical bonds are important in human physiology, because they hold
together substances that are used by the body for critical aspects of homeostasis, signaling,
and energy production, to name just a few important processes. These three types of
chemical bonds are ionic bonds, covalent bonds, and hydrogen bonds.

Ions and Ionic Bonds


Recall that an atom typically has the same number of positively charged protons and
negatively charged electrons. As long as this situation remains, the atom is electrically
neutral. But when an atom participates in a chemical reaction that results in the donation or
acceptance of one or more electrons, the atom will then become positively or negatively
charged. This happens frequently for most atoms in order to have a full valence shell, as
described previously, because a full valence shell makes the atom stable. This can happen
either by gaining electrons to fill a shell that is more than half-full, or by giving away
electrons to empty a shell than is less than half-full. Therefore, if an atom gains an electron,
the atom acquires a negative electrical charge, and if it loses an electron it acquires a
positive electrical charge—whether positive or negative—the atom is known as an ion.
Visit this website (http://openstaxcollege.org/l/electenergy)to learn about electrical energy
and the attraction/ repulsion of charges. What happens to the charged electroscope when a
conductor is moved between its plastic sheets, and why?
Sodium (Na), for instance, is an important element in all body cells. Its atomic number is 11.
It has just one electron in its valence shell. This characteristic makes sodium highly likely to
participate in chemical reactions in which it donates one electron. (It is easier for sodium to
donate one electron than to gain seven electrons.) The loss will cause the positive charge of
sodium's protons to be more influential than the negative charge of sodium electrons. In
other words, the resulting sodium ion will be slightly positive. A sodium ion is written Na+,
indicating that it has lost a single electron. A positively charged ion is known as a cation.
Now consider chlorine (Cl), another important element in the body. Its atomic number is 17,
and it has seven electrons in its valence shell. Thus, it is highly likely to bond with other
atoms in such a way that chlorine accepts one electron (it is easier for chlorine to gain one
electron than to donate seven electrons). When it does, its electrons will outnumber its
protons by one, and it will have an overall negative charge. The ionised form of chlorine is
called chloride, and is written as Cl-. A negatively charged ion is known as an anion.
Atoms that have more than one electron to donate or accept will end up with stronger
positive or negative charges. A cation that has donated two electrons has a net charge of +2.
Using magnesium (Mg) as an example, this can be written Mg++ or Mg2+. An anion that has
accepted two electrons has a net charge of –2. The ionic form of selenium (Se), for example,
is typically written Se2–.
The opposite charges of cations and anions exert a moderately strong mutual attraction that
keeps the atoms in close proximity forming an ionic bond. An ionic bond is an ongoing, close
association between ions of opposite charge. The table salt you sprinkle on your food owes
its existence to ionic bonding. As shown in Figure 3.7, sodium commonly donates
an electron to chlorine, becoming the cation Na+. When chlorine accepts the electron, it
becomes the chloride anion, Cl–. With their opposing charges, these two ions strongly
attract each other and become the compound Sodium Chloride (NaCl).

Figure 3.7 Ionic bonding (a) Sodium readily donates the solitary electron in its valence shell to chlorine, which
needs only one electron to have a full valence shell (b) The opposite electrical charges of the resulting sodium
cation and chloride anion result in the formation of a bond of attraction called an ionic bond.
You could say that the relationship between sodium and chlorine is a 'love story'! A love
story that is disrupted by water. Water is an essential component of life because it is able to
break the ionic bonds in salts to free the ions. In fact, in biological fluids, most individual
atoms exist as ions.
Therefore, when sodium chloride dissolves in water, the ionic bond between sodium and
chloride ions is disrupted; the properties of water which allow this disruption will be
revealed in the section on polar covalent bond later in the chapter.

Covalent Bonds
Unlike ionic bonds formed by the attraction between a cation’s positive charge and an
anion’s negative charge, molecules formed by a covalent bond share electrons in a mutually
stabilising relationship. Like next-door neighbours whose kids hang out first at one home
and then at the other, the atoms do not lose or gain electrons permanently. Instead, the
electrons move back and forth between the elements. Because of the close sharing of pairs
of electrons (one electron from each of two atoms), covalent bonds are stronger than ionic
bonds. There are two types of covalent bonds, non-polar and polar.

Figure 3.8 Covalent Bonding


Nonpolar Covalent Bonds
Figure 3.8 shows several common types of covalent bonds. Notice that the two covalently
bonded atoms typically share just one or two electron pairs, though larger shares are
possible. The important concept to take from this is that in covalent bonds, electrons in the
outermost valence shell are shared to fill the valence shells of both atoms, ultimately
stabilising both of the atoms involved. In a single covalent bond, a single electron is shared
between two atoms, while in a double covalent bond, two pairs of electrons are shared
between two atoms. There even are triple covalent bonds, where three atoms are shared.
You can see that the covalent bonds shown in Figure 3.8 are balanced, that is, the sharing of
the negative electrons is relatively equal, as is the electrical pull of the positive protons in
the nucleus of the atoms involved. This is why covalently bonded molecules that are
electrically balanced in this way are described as nonpolar; that is, no region of the molecule
is either more positive or more negative than any other.

Polar Covalent Bonds


Groups of legislators with completely opposite views on a particular issue are often
described as “polarised” by news writers. In chemistry, a polar molecule is a molecule that
contains regions that have opposite electrical charges. Polar molecules occur when atoms
share electrons unequally, in polar covalent bonds.
When two hydrogen atoms each share their single electron with oxygen, covalent bonds are
formed, resulting in a molecule of water, H2O. Water is the most familiar example of a polar
molecule (Figure 3.9). The molecule has three parts: one atom of oxygen, the nucleus of
which contains eight protons, and two hydrogen atoms, whose nuclei each contain only one
proton. Because every proton exerts an identical positive charge, a nucleus that contains
eight protons exerts a charge eight times greater than a nucleus that contains one proton.
This means that the negatively charged electrons present in the water molecule are more
strongly attracted to the oxygen nucleus than to the hydrogen nuclei. Each hydrogen atom’s
single negative electron therefore migrates toward the oxygen atom, making the oxygen
end of their bond slightly more negative than the hydrogen end of their bond.
What is true for the bonds is true for the water molecule as a whole; that is, the oxygen
region has a slightly negative charge and the regions of the hydrogen atoms have a slightly
positive charge. These charges are often referred to as “partial charges” because the
strength of the charge is less than one full electron, as would occur in an ionic bond. As
shown in Figure 3.9, regions of weak polarity are indicated with the Greek letter delta (∂)
and a plus (+) or minus (–) sign.
Even though a single water molecule is unimaginably tiny, it has mass, and the opposing
electrical charges on the molecule pull that mass in such a way that it creates a shape
somewhat like a triangular tent (see Figure 3.9b). This "dipole", with the positive charges at
one end formed by the hydrogen atoms at the “bottom” of the tent and the negative charge
at the opposite end (the oxygen atom at the “top” of the tent) makes the charged regions
highly likely to interact with charged regions of other polar molecules. For human
physiology, the resulting bond is one of the most important formed by water—the hydrogen
bond.

Figure 3.9 Polar Covalent Bonds in a Water Molecule. Note that the planetary model (a) is so called because
the electrons revolve around the nucleus as planets do around the sun
Hydrogen Bonds
A hydrogen bond is formed when a weakly positive hydrogen atom already bonded to one
electronegative atom (for example, the oxygen in the water molecule) is attracted to
another electronegative atom from another molecule. In other words, hydrogen bonds
always include hydrogen that is already part of a polar molecule.
The most common example of hydrogen bonding in the natural world occurs between
molecules of water. It happens before your eyes whenever two raindrops merge into a
larger bead, or a creek spills into a river. Hydrogen bonding occurs because the weakly
negative oxygen atom in one water molecule is attracted to the weakly positive hydrogen
atoms of two other water molecules (Figure 3.10).

Figure 3.10 Hydrogen Bonds between Water Molecules Notice that the bonds occur between the weakly
positive charge on the hydrogen atoms and the weakly negative charge on the oxygen atoms. Hydrogen bonds
are relatively weak, and therefore are indicated with a dotted (rather than a solid) line.

Water molecules also strongly attract other types of charged molecules as well as ions. This
explains why “table salt,” for example, actually is a molecule called a “salt” in chemistry,
which consists of equal numbers of positively-charged sodium (Na+) and negatively-charged
chloride (Cl–), dissolves so readily in water, in this case forming hydrogen bonds
between the water and the electrically-charged ions Na+ and Cl-. Water molecules also repel
molecules with nonpolar covalent bonds, like fats, lipids, and oils. You can demonstrate this
with a simple kitchen experiment: pour a teaspoon of vegetable oil, a compound formed by
nonpolar covalent bonds, into a glass of water. Instead of instantly dissolving in the water,
the oil forms a distinct bead because the polar water molecules repel the nonpolar oil. In
the next chapter, we will discuss the properties of water at greater length.
KEY TERMS
anion an atom which carries a negative charge
atomic number the number of protons in an element
cation an atom which carries a positive charge
chemical bond any rearrangement of electrons in two atoms generates a force, causing the
atoms to be bound together
compound is formed when two elements are chemically bonded together
covalent bond is a chemical link between two atoms where the electron pairs are shared
between them
electron sub-atomic particle that orbits around the nucleus of an atom. It carries a charge of -1
electron shell group of orbitals followed by electrons around the nucleus of an atom
element the simplest substances that cannot be broken down by chemical means
hydrogen bond electrostatic attraction between a hydrogen ion in one molecule and an
electronegative atom in another molecule
ion an atom that carries a positive or negative charge
ionic bond a chemical bond between two ions with opposite charges
mass number the sum of the protons and neutrons in an atom
molecule forms when 2 or more atoms form chemical bonds; it does not matter if the atoms
are the same or different from one another
nucleus positively charged centre of atoms. It is made up of protons and neutrons, except in
the case of the hydrogen atom which has only one proton and no neutron
neutron subatomic particle found in the nucleus an atom. It has the same mass as a proton
but has no charge
proton subatomic particle found in the nucleus of an atom. It has the same mass as a
neutron but carries a charge of +1
polar molecule has a partial positive charge in one part of the molecule and complementary
negative charge in another part
valence electron an electron of an atom, located in the outermost shell, that can be
transferred to or shared with another atom
valence shell the outermost shell of an atom containing the valence electrons
valency is the number of electrons an atom will accept, donate or share to forma a
chemical bond
REVIEW QUESTIONS
Question 3.3
Is the following statement TRUE or FALSE?
The mass of an object is always equal to its weight.
A TRUE
B FALSE

Question 3.4
The most abundant element in the human body is
A sodium (Na)
B Calcium (Ca)
C Hydrogen (H)
D Oxygen (O)

Question 3.5
The atoms of all elements, with the exception of hydrogen, are made up of three types of
sub-atomic particles, that is, protons, neutrons and electrons
A TRUE
B FALSE

Question 3.6
Which of the following statement(s) is(are) true
A The proton is positively charged and found in the nucleus of the atom
B The neutron is also found in the nucleus and has the same mass as the proton
C The atomic number of an element is equal to the number of neutrons in the nucleus of
each of its atoms
D The atomic number of an element is equal to the number of protons in the nucleus of
each one of its atoms
Question 3.7
Which of the following statements is true for the electron?
A It is found in the nucleus of the atom
B It has a mass equal to that of a proton
C It is positively charged and revolves around the nucleus
D In a neutral atom, the number of electrons equals the number of protons

Question 3.8
Which of the following statements is/are true about an isotope?
A Isotopes are different forms of the same elements
B The only difference between isotopes of the same element is that they have a different
number of electrons
C An isotope is said to be radioactive (a radioisotope) if its nucleus readily decays
D Radioisotopes can be used to treat cancer

Question 3.9
The first, second and third shells of an atom can hold 2, 8 and 8 electrons respectively.
A TRUE
B FALSE

Question 3.10
Which of the statements below is/are correct?
A A valence shell is the outermost electron shell of an atom
B The chemical reactivity of an element is governed by the number of electrons in its
valence shell
C The chemical reactivity of an element is governed by how readily its atoms can dissociate
into its subatomic particles
D Valency of an element can be described as the readiness with which its atoms can donate,
accept or share electrons to form a chemical bond with atoms of another element
Question 3.11
Ions are negatively or positively charged atoms.
A TRUE
B FALSE

Question 3.12
A covalent bond is formed between two elements when their atoms share the valence
electrons.
A TRUE
B FALSE
ANSWERS
Question 3.1
20
Potassium-39 or 39K: Atomic number is 19, therefore number of neutrons = 39 – 19 = 20

Question 3.2
21
Potassium-40 or 40K: Atomic number is 19, therefore number of neutrons = 40 – 19 = 21

Question 3.3
B

Question 3.4
D

Question 3.5
A

Question 3.6
A, B, D

Question 3.7
D

Question 3.8
A, C, D

Question 3.9
A

Question 3.10
A, B, D

Question 3.11
A

Question 3.12
A
4 | WATER BIOLOGY

Figure 4.1 Electron sharing by oxygen and hydrogen to form the water molecule

By the end of this chapter, you will be able to:


• Identify the properties of water that make it essential to life
• Explain the role of salts in body functioning
• Distinguish between acids and salts and explain their role in pH
• Discuss the role of buffers in helping the body maintain pH homeostasis
The concepts you learned about in Chapter 3 govern all forms of matter, and would work as
a foundation for geology as well as biology. We now narrow the focus to the chemistry of
human life; that is, the compounds important for the body’s structure and function. In
general, these compounds are either inorganic or organic.
This chapter examines the three groups of inorganic compounds essential to life:
• water,
• salts,
• acids and bases.

4.1 Properties of water


As much as 70 percent of an adult’s body weight is water. This water is contained both
within the cells and between the cells that make up tissues and organs (Figure 4.2). Its
several roles make water indispensable to human functioning.

Figure 4.2 Water is found inside cells (ICF), and outside cells (ECF). ECF is the term used for the fluid found
between cells (IF = Interstitial Fluid) and within plasma.
Water as a Lubricant and Cushion
Water is a major component of many of the body’s lubricating fluids. Just as oil lubricates
the hinge on a door, water in synovial fluid lubricates the actions of body joints, and water
in pleural fluid (the fluid between the layers of membranes, or pleura that surround the
lung) helps the lungs expand and recoil with breathing. Watery fluids help keep food flowing
through the digestive tract, and ensure that the movement of adjacent abdominal organs is
friction free.
Water also protects cells and organs from physical trauma, cushioning the brain within the
skull, for example, and protecting the delicate nerve tissue of the eyes. Water cushions a
developing fetus in the mother’s womb as well.

Water as a Heat Sink


A heat sink is a substance or object that absorbs and dissipates heat but does not
experience a corresponding increase in temperature. In the body, water absorbs the heat
generated by chemical reactions without greatly increasing in temperature. Moreover,
when the environmental temperature soars, the water stored in the body helps keep the
body cool. This cooling effect happens as warm blood from the body’s core flows to the
blood vessels just under the skin and is transferred to the environment. At the same time,
sweat glands release warm water in sweat. As the water evaporates into the air, it
carries away heat, and then the cooler blood from the periphery circulates back to the body
core.

Water as a Component of Liquid Mixtures


A mixture is a combination of two or more substances, each of which maintains its own
chemical identity. In other words, the constituent substances are not chemically bonded
into a new, larger chemical compound. The concept is easy to imagine if you think of
powdery substances such as flour and sugar; when you stir them together in a bowl,
they obviously do not bond to form a new compound. The room air you breathe is a
gaseous mixture, containing three discrete elements—nitrogen, oxygen, and argon—and
one compound, carbon dioxide. There are three types of liquid mixtures, all of which
contain water as a key component. These are solutions, colloids, and suspensions.
For cells in the body to survive, they must be kept moist in a water-based liquid called a
solution. In chemistry, a liquid solution consists of a solvent (in this case water) that
dissolves a substance called a solute.
Solution = Solvent + Solute
An important characteristic of solutions is that they are homogeneous; that is, the solute
molecules are distributed evenly throughout the solution. If you were to stir a teaspoon of
sugar into a glass of water, the sugar would dissolve into sugar molecules separated by
water molecules. The ratio of sugar to water in the left side of the glass would be the same
as the ratio of sugar to water in the right side of the glass. If you were to add more sugar,
the ratio of sugar to water would change, but the distribution—provided you had
stirred well—would still be even.
Water is considered the “universal solvent” and it is believed that life cannot exist without
water because of this. Water is certainly the most abundant solvent in the body; essentially
all of the body’s chemical reactions occur among compounds dissolved in water. Because
water molecules are polar, with regions of positive and negative electrical charge,
water readily dissolves ionic compounds and polar covalent compounds. Such compounds
are referred to as hydrophilic, or “water-loving.” As mentioned above, sugar dissolves well
in water. This is because sugar molecules contain regions of hydrogen-oxygen polar bonds,
making it hydrophilic. Nonpolar molecules, which do not readily dissolve in water, are
called hydrophobic, or “water-fearing.”

Concentrations of Solutes
Various mixtures of solutes and water are described in chemistry. The concentration of a
given solute is the number of particles of that solute in a given space. Therefore, oxygen
makes up about 21 percent of atmospheric air; this is the concentration of oxygen in air. On
the other hand the concentration of glucose in a solution could be expressed as milligrams
(mg) per litre (L).
Another method of measuring the concentration of a solute is by its molarity—which is
moles (M) of the molecules per litre (L). The mole of an element is its atomic weight, while
a mole of a compound is the sum of the atomic weights of its components, called the
molecular weight. An often-used example is calculating a mole of glucose, with the chemical
formula C6H12O6. Let us go through how the calculation is done, using the periodic table:
• the atomic weight of carbon (C) is 12.011 grams (g). therefore six carbons weigh
72.006g
• Doing the same calculations for hydrogen (H), twelve hydrogens weigh 12.096g
• and for oxygen (O), six oxygens weigh 95.994g
• giving the total molecular weight of 180.156g (the “gram molecular weight” of
glucose).
• When water is added to make one litre of solution, the resulting concentration is
said to be 1M because one molecule of glucose because one molecule of glucose
(180.156g is dissolved in it)
• Many substances in the bloodstream and other tissue of the body are measured in
thousandths of a mole, or millimoles (mM).

Colloids and suspensions


A colloid is a mixture that is somewhat like a heavy solution. The solute particles consist of
tiny clumps of molecules large enough to make the liquid mixture opaque (because the
particles are large enough to scatter light). Familiar examples of colloids are milk and cream.
In the thyroid glands, the thyroid hormone is stored as a thick protein mixture also called
a colloid.
A suspension is a liquid mixture in which a heavier substance is suspended temporarily in a
liquid, but over time, settles out. This separation of particles from a suspension is called
sedimentation. An example of sedimentation occurs in the blood test that establishes
sedimentation rate, or sed rate. The test measures how quickly red blood cells in a test tube
settle out of the watery portion of blood (known as plasma) over a set period of time. Rapid
sedimentation of blood cells does not normally happen in the healthy body, but aspects of
certain diseases can cause blood cells to clump together, and these heavy clumps of blood
cells settle to the bottom of the test tube more quickly than do normal blood cells.

The Role of Water in Chemical Reactions


Two types of chemical reactions involve the creation or the consumption of water:
dehydration synthesis and hydrolysis (Figure 4.3).
These reactions are reversible, and play an important role in the chemistry of organic
compounds (which will be discussed later in the textbook).

Figure 4.3 Dehydration Synthesis and Hydrolysis Monomers, the basic units for building larger molecules,
form polymers (two or more chemically-bonded monomers). (a) In dehydration synthesis, two monomers are
covalently bonded in a reaction in which one gives up a hydroxyl group and the other a hydrogen atom. A
molecule of water is released as a byproduct during dehydration reactions. (b) In hydrolysis, the covalent bond
between two monomers is split by the addition of a hydrogen atom to one and a hydroxyl group to the other,
which requires the contribution of one molecule of water.

4.2 Salts, Acids and Bases, Buffers


Recall that salts are formed when ions form ionic bonds. In these reactions, one atom gives
up one or more electrons, and thus becomes positively charged, whereas the other accepts
one or more electrons and becomes negatively charged. You can now define a salt as a
substance that, when dissolved in water, dissociates into ions other than H+ or OH–. This fact
is important in distinguishing salts from acids and bases, discussed next.

Figure 4.4 Dissociation of Sodium Chloride in Water Notice that the crystals of sodium chloride dissociate
not into molecules of NaCl, but into Na+ cations and Cl– anions, each completely surrounded by water
molecules. The symbol δ represents a partial charge, negative or positive.
A typical salt, NaCl, dissociates completely in water (Figure 4.4). The positive and negative
regions on the water molecule (the hydrogen and oxygen ends respectively) attract the
negative chloride and positive sodium ions, pulling them away from each other. In
solution, salts dissociate into ions. These ions are electrolytes; they are capable of
conducting an electrical current in solution. This property is critical to the function of ions in
transmitting nerve impulses and prompting muscle contraction.
Many other salts are important in the body. For example, bile salts produced by the liver
help break apart dietary fats, and calcium phosphate salts form the mineral portion of teeth
and bones.
Acids and bases, like salts, dissociate in water into electrolytes. Acids and bases can very
much change the properties of the solutions in which they are dissolved.

Acids
An acid is a substance that releases hydrogen ions (H+) in solution (Figure 4.5a). Because an
atom of hydrogen has just one proton and one electron, and the single electron is missing in
the H+, the hydrogen ion is simply just a proton. This solitary proton is highly likely
to participate in chemical reactions. Strong acids are compounds that release all of their
H+ in solution; that is, they ionise completely. Hydrochloric acid (HCl), which is released from
cells in the lining of the stomach, is a strong acid because it releases all of its H+ in the
stomach’s watery environment. This strong acid aids in digestion and kills ingested
microbes. Weak acids do not ionise completely; that is, some of their hydrogen ions remain
bonded within a compound in solution. An example of a weak acid is vinegar, or acetic acid;
it is called acetate after it gives up a proton.

Bases
A base (also called an alkali) is a substance that releases hydroxyl ions (OH–) in solution
(see Figure 4.5b), or one that accepts H+ already present in solution. The OH– combine with
H+ present to form a water molecule, thereby removing H+ and reducing the solution’s
acidity. Strong bases release most or all of their hydroxyl ions; weak bases release only
some hydroxyl ions or absorb only a few H+. Food mixed with hydrochloric acid from the
stomach would burn the small intestine, the next portion of the digestive tract after the
stomach, if it were not for the release of bicarbonate (HCO3–), a weak base that attracts H+.
Bicarbonate accepts some of the H+ protons, thereby reducing the acidity of the solution.

The Concept of pH
The pH of a solution is a measure of its acidity or its alkalinity (or how basic it is). You have
probably used litmus paper, paper that has been treated with a natural water-soluble dye
so it can be used as a pH indicator, to test how much acid or base exists in a solution. You
might have even used some to make sure the water in an outdoor swimming pool is
properly treated. In both cases, this pH test measures the amount of hydrogen ions that
exists in a given solution. High concentrations of hydrogen ions yield a low pH, whereas low
levels of hydrogen ions result in a high pH. The overall concentration of hydrogen ions is
inversely related to its pH and can be measured on the pH scale (Figure 4.6). Therefore, the
more hydrogen ions present, the lower the pH; conversely, the fewer hydrogen ions, the
higher the pH.

Figure 4.5 Acids and Bases (a) In aqueous solution, an acid dissociates into hydrogen ions (H+) and anions.
Nearly every molecule of a strong acid dissociates, producing a high concentration of H+. (b) In aqueous
solution, a base dissociates into hydroxyl ions (OH–) and cations. Nearly every molecule of a strong base
dissociates, producing a high concentration of OH–.

The pH scale ranges from 0 to 14. A change of one unit on the pH scale represents a change
in the concentration of hydrogen ions by a factor of 10, a change in two units represents a
change in the concentration of hydrogen ions by a factor of 100. Thus, small changes in pH
represent large changes in the concentrations of hydrogen ions. Pure water is neutral. It is
neither acidic nor basic, and has a pH of 7.0. Anything below 7.0 (ranging from 0.0 to 6.9) is
acidic, and anything above 7.0 (from 7.1 to 14.0) is alkaline. The blood in your veins is
slightly alkaline (pH = 7.4). The environment in your stomach is highly acidic (pH = 1 to 2).
Orange juice is mildly acidic (pH = approximately 3.5), whereas baking soda is basic/alkaline
(pH = 9.0).
Figure 4.6 The pH Scale
Buffers
Most cells in our bodies operate within a very narrow window of the pH scale, typically
ranging only from 7.2 to 7.6. If the pH of the body is outside of this range, the respiratory
system malfunctions, as do other organs in the body. Cells no longer function properly, and
proteins will break down. Deviation outside of the pH range can induce coma or even cause
death.
So how is it that we can ingest or inhale acidic or basic substances and not die? Buffers are
the key. Buffers readily absorb excess H+or OH–, keeping the pH of the body carefully
maintained in the aforementioned narrow range. Carbon dioxide is part of a prominent
buffer system in the human body; it keeps the pH within the proper range. This buffer
system involves carbonic acid (H2CO3) and bicarbonate (HCO3–) anion. If too much H+enters
the body, bicarbonate will combine with the H+ to create carbonic acid and limit the
decrease in pH. Likewise, if too much OH– is introduced into the system, carbonic acid will
rapidly dissociate into bicarbonate and H+ions. The H+ions can combine with the OH–ions,
limiting the increase in pH. While carbonic acid is an important product in this reaction, its
presence is fleeting because the carbonic acid is released from the body as carbon dioxide
gas each time we breathe. Without this buffer system, the pH in our bodies would fluctuate
too much and we would fail to survive. This buffer system will be discussed in more detail in
Chapter 9.
The pH of human blood normally ranges from 7.35 to 7.45, although it is typically identified
as pH 7.4. At this slightly basic pH, blood can reduce the acidity resulting from the carbon
dioxide (CO2) constantly being released into the bloodstream by the trillions of cells in the
body. Homeostatic mechanisms (along with exhaling CO2 while breathing) normally keep
the pH of blood within this narrow range. This is critical, because fluctuations—either too
acidic or too alkaline—can lead to life-threatening disorders.

Acids and Bases


Excessive acidity of the blood and other body fluids is known as acidosis. Common causes of
acidosis are situations and disorders that reduce the effectiveness of breathing, especially
the person’s ability to exhale fully, which causes a buildup of CO2 (and H+) in the
bloodstream. Acidosis can also be caused by metabolic problems that reduce the level or
function of buffers that act as bases, or that promote the production of acids. For instance,
with severe diarrhoea, too much bicarbonate can be lost from the body, allowing acids to
build up in body fluids. In people with poorly managed diabetes (ineffective regulation of
blood sugar), acids called ketones are produced as a form of body fuel. These can build up in
the blood, causing a serious condition called diabetic ketoacidosis. Kidney failure, liver
failure, heart failure, cancer, and other disorders also can prompt metabolic acidosis.
In contrast, alkalosis is a condition in which the blood and other body fluids are too alkaline
(basic). As with acidosis, respiratory disorders are a major cause; however, in respiratory
alkalosis, carbon dioxide levels fall too low. Lung disease, aspirin overdose, shock, and
ordinary anxiety can cause respiratory alkalosis, which reduces the normal concentration of
H+.
Metabolic alkalosis often results from prolonged, severe vomiting, which causes a loss of
hydrogen and chloride ions (as components of HCl). Medications also can prompt alkalosis.
These include diuretics that cause the body to lose potassium ions, as well as antacids when
taken in excessive amounts, for instance by someone with persistent heartburn or an ulcer.

KEY TERMS
acid a substance that releases hydrogen ions (H+) in solution
base a substance that releases hydroxyl ions (0H-) in solution
buffer a solution that can resist pH change upon the addition of an acidic
or basic components
colloid a mixture where very small particles of one substance are evenly distributed
throughout another substance
electrolytes ions that are capable of conducting an electrical current when in solution
hydrophilic water loving
hydrophobic water hating
pH a measure of acidity or alkalinity of water soluble substances
monomer a molecule that may react chemically with another molecule of the same type to
form a larger molecule called a polymer
salt a substance that, when dissolved in water, dissociates into ions other than H+ or OH–
solution a mixture of substances where one of the substances dissolves in the other
solute the substance that is dissolved in the solvent
solvent (liquid) the compound of a solution that is present in the greatest amount
suspension a mixture of liquid and solid particles where solid particles will eventually
separate and settle if left undisturbed
REVISION QUESTIONS
Question 4.1
In the human body, extracellular fluid is found inside cells whereas intracellular fluid is
found outside cells
A FALSE
B TRUE

Question 4.2
Which of the following statements is/are true about water?
A Water acts as a lubricating fluid in the body
B Water helps the body maintain its body temperature
C Water is a non-polar molecule
D Water is a polar molecule

Question 4.3
Each water molecule has regions which are partially positive and partially negative. This
means that water molecules are:
A non-polar
B polar

Question 4.4
Because of its polar nature, water dissolves polar and ionic compounds easily.
A TRUE
B FALSE
Question 4.5
A solution is formed when a solid substance (solute) dissolves in a liquid (solvent).
A FALSE
B TRUE

Question 4.6
What is the weight of one molecule of sodium chloride (NaCl)

Question 4.7
A salt is a substance that, when dissolved in water, dissociates into ions other than H+ or
OH–.
A TRUE
B FALSE

Question 4.8
Most cell in the body operate between the pH range of 7.2 and 7.6.
A FALSE
B TRUE

Question 4.9
Which one of the following statements is false"
A Buffers help to maintain a narrow range of pH in the body
B A strong acid dissociates in water resulting in a solution which is high in hydroxyl ions
C A strong acid dissociates in water resulting in a solution which is high in hydrogen ions
D A solution that has a pH of 14 is strongly alkaline or strongly basic
ANSWERS
Question 4.1
A

Question 4.2
A, B, D

Question 4.3
B

Question 4.4
A

Question 4.5
B

Question 4.6
From the periodic table it can be seen that the atomic weight of sodium is 22.99g and that
for chlorine is 35.45g. Since one molecule of NaCl is made up of one sodium atom and one
chloride atom, we simply add the two atomic weights to get a total of 58.44 g. Therefore the
molecular weight of sodium is 58.44g.

Question 4.7
A

Question 4.8
B

Question 4.9
B
5| THE CELL MEMBRANE

Figure 5.1 Electron micrograph showing lipid bilayer of adjacent cells. The lipid bilayer is less than 5nm thick
but the individual leaflets of the bilayer can be resolved. (photo Shigeki Watanabe and Erik Jorgensen)

By the end of this chapter, you will be able to:


• Describe the molecular components that make up the cell membrane
• Describe the major features and properties of the cell membrane
• Differentiate between materials that can and cannot diffuse through the lipid bilayer
• Compare and contrast different types of passive transport with active transport,
providing examples of each
Despite differences in structure and function, all living cells in multicellular organisms have a
surrounding cell membrane. As the outer layer of your skin separates your body from its
environment, the cell membrane (also known as the plasma membrane) separates the inner
contents of a cell from its exterior environment. This cell membrane provides a
protective barrier around the cell and regulates which materials can pass in or out.
5.1 Structure and Composition of the Cell
Membrane
The cell membrane is an extremely pliable structure composed primarily of back-to-back
phospholipids (a “bilayer”). Cholesterol is also present, which contributes to the fluidity of
the membrane, and there are various proteins embedded within the membrane that have a
variety of functions.

Figure 5.2 (a) The phospholipid contains a three-carbon glycerol backbone to which two lipid tails (these lipids
are fatty acids) and one phosphate group are attached. (b) The phospholipid bilayer forms by self-assembly
with two sheets of phospholipids aligning in an inverse arrangement. The polar head groups are arranged near
the extracellular and cytosolic fluids while the nonpolar tails form a barrier in between.

A single phospholipid molecule has a phosphate group on one end, called the “head,” and
two side-by-side chains of fatty acids that make up the lipid tails (Figure 5.2a). The
phosphate group is negatively charged, making the head polar and hydrophilic—or “water
loving.” A hydrophilic molecule (or region of a molecule) is one that is attracted to water.
The phosphate heads are thus attracted to the water molecules of both the extracellular
and intracellular environments. The lipid tails, on the other hand, are uncharged, or
nonpolar, and are hydrophobic—or “water fearing.” A hydrophobic molecule (or region of a
molecule) repels and is repelled by water. The tails are constantly in motion. If several
phospholipid molecules are placed in water, they will self assemble in a circle in a bilayer as
shown in Figure 5.2b below. Soap is made up similar molecules that assemble in the same
manner. Therefore, the hydrophilic portion can dissolve in water while the hydrophobic
portion can trap grease in micelles (structures like in Figure 5.2b) that then can be washed
away.
The important point to remember is that the phospholipid bilayer effectively establishes a
barrier between internal and external environments.
The cell membrane consists of two adjacent layers of phospholipids. The lipid tails of one
layer face the lipid tails of the other layer, meeting at the interface of the two layers. The
phospholipid heads face outward, one layer exposed to the interior of the cell and one layer
exposed to the exterior (Figure 5.3). Because the phosphate groups are polar and
hydrophilic, they are attracted to water in the intracellular fluid. Intracellular fluid (ICF) is
the fluid in the interior of the cell. The phosphate groups are also attracted to the
extracellular fluid. Extracellular fluid (ECF) is the fluid environment outside the enclosure of
the cell membrane. Because the lipid tails are hydrophobic, they meet in the inner region of
the membrane, excluding watery intracellular and extracellular fluid from this space. The
cell membrane has many proteins, as well as other lipids (such as cholesterol), that are
associated with the phospholipid bilayer. An important feature of the membrane is that it
remains fluid; the lipids and proteins in the cell membrane are not rigidly locked in place
Figure 5.5.

Figure 5.3 Phospholipid Bilayer The phospholipid bilayer consists of two adjacent sheets of phospholipids,
arranged tail to tail. The hydrophobic tails associate with one another, forming the interior of the membrane.
The polar heads contact the fluid inside and outside of the cell.
Membrane Proteins
The lipid bilayer forms the basis of the cell membrane, but it is peppered throughout with
various proteins. Two different types of proteins that are commonly associated with the cell
membrane are the integral proteins and peripheral protein (Figure 5.4). As its name
suggests, an integral protein is a protein that is embedded in the membrane.

Figure 5.4 Cell Membrane The cell membrane of the cell is a phospholipid bilayer containing many
different molecular components, including proteins and cholesterol, some with carbohydrate groups attached.

Some types of integral proteins are:


• Channel protein which selectively allows particular materials, such as certain ions, to
pass into or out of the cell.
• Cell recognition proteins (for example, glycoproteins), which serve to mark a cell’s
identity so that it can be recognised by other cells
• Receptor proteins that can selectively bind a specific molecule outside the cell, and
this binding induces a chemical reaction within the cell.
Peripheral proteins are typically found on the inner or outer surface of the lipid bilayer but
can also be attached to the internal or external surface of an integral protein. These
proteins typically perform a specific function for the cell. Some peripheral proteins on the
surface of intestinal cells, for example, act as digestive enzymes (enzymes are proteins!) to
break down nutrients to sizes that can pass through the cells and into the bloodstream.
Figure 5.5 The plasma membrane is fluid not rigid

5.2 Transport across the Cell Membrane


One of the great wonders of the cell membrane is its ability to regulate the concentration of
substances inside the cell. These substances include ions such as Ca++, Na+, K+, and Cl–;
nutrients including sugars, fatty acids, and amino acids; and waste products, particularly
carbon dioxide (CO2), which must leave the cell.
The membrane’s lipid bilayer structure provides the first level of control. The phospholipids
are tightly packed together, and the membrane has a hydrophobic interior. This structure
causes the membrane to be selectively permeable. A membrane that
has selective permeability allows only substances meeting certain criteria to pass through it
unaided. In the case of the cell membrane, only relatively small, nonpolar materials can
move through the lipid bilayer (remember, the lipid tails of the membrane are nonpolar).
Some examples of these are other lipids, oxygen and carbon dioxide gases, and alcohol.
However, water-soluble materials—like glucose, amino acids, and electrolytes—need some
assistance to cross the membrane because they are repelled by the hydrophobic tails of the
phospholipid bilayer. All substances that move through the membrane do so by one of two
general methods, which are categorised based on whether or not energy is required.
• Passive transport is the movement of substances across the membrane without the
expenditure of cellular energy.
• In contrast, active transport is the movement of substances across the membrane
using energy; this energy comes from the molecule adenosine triphosphate (ATP).

Passive Transport
In order to understand how substances move passively across a cell membrane, it is
necessary to understand concentration gradients and diffusion. A concentration gradient is
the difference in concentration of a substance across a space. Molecules (or ions) will
spread/diffuse from where they are more concentrated to where they are less concentrated
until they are equally distributed in that space. (When molecules move in this way, they are
said to move down their concentration gradient). Diffusion is the movement of particles
from an area of higher concentration to an area of lower concentration. A couple of
common examples will help to illustrate this concept. Imagine being inside a closed
bathroom. If a bottle of perfume were sprayed, the scent molecules would naturally diffuse
from the spot where they left the bottle to all corners of the bathroom, and this diffusion
would go on until no more concentration gradient remains. Another example is a
spoonful of sugar placed in a cup of tea. Eventually the sugar will diffuse throughout the tea
until no concentration gradient remains. In both cases, if the room is warmer or the tea
hotter, diffusion occurs even faster as the molecules are bumping into each other and
spreading out faster than at cooler temperatures. Having an internal body temperature
around 37.5oC F thus also aids in diffusion of particles within the body.
Visit this link (http://openstaxcollege.org/l/diffusion) to see diffusion and how it is
propelled by the kinetic energy of molecules in solution.
Whenever a substance exists in greater concentration on one side of a semipermeable
membrane, such as the cell membranes, any substance that can move down its
concentration gradient across the membrane will do so. Consider substances that can easily
diffuse through the lipid bilayer of the cell membrane, such as the gases oxygen (O2) and
CO2. O2 generally diffuses into cells because it is more concentrated outside of them, and
CO2 typically diffuses out of cells because it is more concentrated inside of them. Neither of
these examples requires any energy on the part of the cell, and therefore they use passive
transport to move across the membrane.
Before moving on, you need to review the gases that can diffuse across a cell membrane.
Because cells rapidly use up oxygen during metabolism, there is typically a lower
concentration of O2 inside the cell than outside. As a result, oxygen will diffuse from outside
the cell through the lipid bilayer of the membrane and into the cytoplasm within the cell. On
the other hand, because cells produce CO2 as a byproduct of metabolism,
CO2 concentrations rise within the cytoplasm; therefore, CO2 will move from the cell
through the lipid bilayer and outside the cell, where its concentration is lower. This
mechanism of molecules spreading from where they are more concentrated to where they
are less concentration is a form of passive transport called simple diffusion (Figure 5.6).
Figure 5.6 Simple Diffusion across the Cell (Plasma) Membrane The structure of the lipid bilayer allows
only small, non-polar substances such as oxygen and carbon dioxide to pass through the cell membrane, down
their concentration gradient, by simple diffusion.

Solutes dissolved in water on either side of the cell membrane will tend to diffuse down
their concentration gradients, but because most substances cannot pass freely through the
lipid bilayer of the cell membrane, their movement is restricted to protein channels and
specialised transport mechanisms in the membrane (Figure 5.7). Facilitated diffusion is the
diffusion process used for those substances that cannot cross the lipid bilayer due to their
size and/or polarity. A common example of facilitated diffusion is the movement of glucose
into the cell, where it is used to make ATP. Although glucose can be more concentrated
outside of a cell, it cannot cross the lipid bilayer via simple diffusion because it is both large
and polar. To resolve this, a specialised carrier protein called the glucose transporter will
transfer glucose molecules into the cell to facilitate its inward diffusion.
As another example, even though sodium ions (Na+) are highly concentrated outside of cells,
these electrolytes are polarised and cannot pass through the nonpolar lipid bilayer of the
membrane. Their diffusion is facilitated by membrane proteins that form sodium channels
(or “pores”), so that Na+ ions can move down their concentration gradient from outside the
cells to inside the cells. There are many other solutes that must undergo facilitated diffusion
to move into a cell, such as amino acids, or to move out of a cell, such as wastes. Because
facilitated diffusion is a passive process, it does not require energy expenditure by the cell.
Water also can move freely across the cell membrane of all cells, either through protein
channels or by slipping between the lipid tails of the membrane itself. Osmosis is the
diffusion of water through a semipermeable membrane (Figure 5.8).
The movement of water molecules is not itself regulated by cells, so it is important that cells
are exposed to an environment in which the concentration of solutes outside of the cells (in
the extracellular fluid) is equal to the concentration of solutes inside the cells (in the
cytoplasm). Two solutions that have the same concentration of solutes are said to
be isotonic (equal tension). When cells and their extracellular environments are isotonic,
the concentration of water molecules is the same outside and inside the cells, and the cells
maintain their normal shape (and function).

Figure 5.7 Facilitated Diffusion (a) Facilitated diffusion of substances crossing the cell (plasma) membrane
takes place with the help of proteins such as channel proteins and carrier proteins. Channel proteins are less
selective than carrier proteins, and usually mildly discriminate between their cargo based on size and charge.
(b) Carrier proteins are more selective, often only allowing one particular type of molecule to cross.
Figure 5.8 Osmosis is the diffusion of water through a semipermeable membrane down its
concentration gradient. If a membrane is permeable to water, though not to a solute, water will equalise its
own concentration by diffusing to the side of lower water concentration (and thus the side of higher solute
concentration). In the beaker on the left, the solution on the right side of the membrane is hypertonic to start
with, but becomes isotonic after osmosis.

Osmosis occurs when there is an imbalance of solutes outside of a cell versus inside the cell.
A solution that has a higher concentration of solutes than another solution is said to
be hypertonic, and water molecules tend to diffuse into a hypertonic solution. In contrast, a
solution that has a lower concentration of solutes than another solution is said to
be hypotonic, and water molecules tend to diffuse out of a hypotonic solutions it does in
the example shown in Figure 5.8.
Figure 5.9 illustrates what happens to red blood cells when they are placed in solutions of
different tonicity. Cells in a hypertonic solution will shrivel as water leaves the cell via
osmosis. . Cells in a hypotonic solution will take on too much water and swell, with the risk
of eventually bursting. Lastly, cells in an isotonic solution remain unchanged. Therefore, it
makes sense that the fluid (plasma) that red blood cells float around in is isotonic. A critical
aspect of homeostasis in living things is to create an internal environment in which all of the
body’s cells are in an isotonic solution. Various organ systems, particularly the kidneys, work
to maintain this homeostasis.
Another mechanism besides diffusion to passively transport materials between
compartments is filtration. Unlike diffusion of a substance from where it is more
concentrated to less concentrated, filtration uses a hydrostatic pressure gradient
that pushes the fluid—and the solutes within it—from a higher pressure area to a lower
pressure area. Filtration is an extremely important process in the body. For example, the
circulatory system uses filtration to move plasma and substances across the endothelial (the
inside of capillaries is lined with endothelial cells) lining of capillaries and into surrounding
tissues, supplying cells with the nutrients. Filtration pressure in the kidneys provides the
mechanism to remove wastes from the bloodstream.
Figure 5.9 Red blood cells in different concentration of solutions. In a hypertonic solution, water will move out
from the blood cells into the outside solution. In an isotonic solution, the net movement of water is zero. In
a hypotonic solution water will move into the cells.

Active Transport
For all of the transport methods described above, the cell expends no energy. Membrane
proteins that aid in the passive transport of substances do so without the use of ATP. In
contrast, during active transport, ATP is required to move a substance across a membrane,
often with the help of protein carriers, and usually against its concentration gradient.
In addition to moving small ions and molecules through the membrane, cells also need to
remove and take in larger molecules and particles. Some cells are even capable of engulfing
entire unicellular microorganisms. You might have correctly hypothesised that the uptake
and release of large particles by the cell requires energy. A large particle, however, cannot
pass through the membrane, even with energy supplied by the cell.
We have discussed simple concentration gradients—differential concentrations of a
substance across a space or a membrane—but in living systems, gradients are more
complex. Because cells contain proteins, most of which are negatively charged, and because
ions move into and out of cells, there is an electrical gradient, a difference of charge, across
the plasma membrane. The interior of living cells is electrically negative with respect to the
extracellular fluid in which they are bathed; at the same time, cells have higher
concentrations of potassium ions (K+) and lower concentrations of sodium ions (Na+) than
does the extracellular fluid. Thus, in a living cell, the concentration gradient and electrical
gradient of Na+ promotes diffusion of the ion into the cell, and the electrical gradient of
Na+ (a positive ion) tends to drive it inward to the negatively charged interior. The situation
is more complex, however, for other elements such as potassium. The electrical gradient of
K+ promotes diffusion of the ion into the cell, but the concentration gradient of K+ promotes
diffusion out of the cell (Figure 5.10). The combined gradient that affects an ion is called
its electrochemical gradient, and it is especially important to muscle and nerve cells.
Figure 5.10 Electrochemical gradients arise from the combined effects of concentration gradients and
electrical gradients. (credit: modification of work by “Synaptitude”/Wikimedia Commons)

Moving Against a Gradient


To move substances against a concentration or an electrochemical gradient, the cell must
use energy. This energy is harvested from ATP that is generated through cellular
metabolism. Active transport mechanisms, collectively called pumps or carrier proteins,
work against electrochemical gradients. With the exception of ions, small substances
constantly pass through plasma membranes. Active transport maintains concentrations of
ions and other substances needed by living cells in the face of these passive changes. Much
of a cell’s supply of metabolic energy may be spent maintaining these processes. Because
active transport mechanisms depend on cellular metabolism for energy, they are sensitive
to many metabolic poisons that interfere with the supply of ATP.
Two mechanisms exist for the transport of small-molecular weight material and
macromolecules. Primary active transport moves ions across a membrane and creates a
difference in charge across that membrane. The primary active transport system uses ATP to
move a substance, such as an ion, into the cell, and often at the same time, a second
substance is moved out of the cell. The sodium-potassium pump, an important pump in
animal cells, expends energy to move potassium ions into the cell and a different number of
sodium ions out of the cell (Figure 5.11). The action of this pump results in a concentration
and charge difference across the membrane.

Figure 5.11 The sodium-potassium pump facilitates movement of Na+and K+ ions across the membrane, each
going against their respective concentration gradients. The concentration of Na+is maintained at a higher level
in the extracellular environment while K+ is maintained at a higher concentration in the cytoplasm. One ATP
molecule is invested for each cycle of the pump, which moves 3 sodium ions and 2 potassium ions across the
membrane.

Endocytosis is a type of active transport that moves particles, such as large molecules, parts
of cells, and even whole cells, into a cell. There are different variations of endocytosis, but
all share a common characteristic: The plasma membrane of the cell invaginates, forming a
pocket around the target particle. The pocket pinches off, resulting in the particle being
contained in a newly created vacuole that is formed from the plasma membrane. Below, we
discuss three types of endocytosis (Figure 5.12).
Phagocytosis is the process by which large particles, such as cells, are taken in by a cell. For
example, when microorganisms invade the human body, a type of white blood cell called a
neutrophil removes the invader through this process, surrounding and engulfing the
microorganism (Figure 5.12a), which is then destroyed by the neutrophil . A variation of
endocytosis is called pinocytosis. This literally means “cell drinking” and was named at a
time when the assumption was that the cell was purposefully taking in extracellular fluid. In
reality, this process takes in solutes that the cell needs from the extracellular fluid (Figure
5.12b).
Figure 5.12 Three variations of endocytosis are shown. (a) In one form of endocytosis, phagocytosis, the cell
membrane surrounds the particle and pinches off to form an intracellular vacuole. (b) In another type of
endocytosis, pinocytosis, the cell membrane surrounds a small volume of fluid and pinches off, forming a
vesicle. (c) In receptor-mediated endocytosis, uptake of substances by the cell is targeted to a single type of
substance that binds at the receptor on the external cell membrane. (credit: modification of work by Mariana
Ruiz Villarreal)

A targeted variation of endocytosis employs binding proteins in the plasma membrane that
are specific for certain substances (Figure 5.12c). The particles bind to the receptors
(proteins on the external plasma membrane surface) and the plasma membrane
invaginates, bringing the substance and the receptors into the cell. If passage across the
membrane of the target of receptor-mediated endocytosis is ineffective, it will not be
removed from the tissue fluids or blood. Instead, it will stay in those fluids and increase in
concentration. Some human diseases are caused by a failure of receptor-mediated
endocytosis. For example, the form of cholesterol termed low-density lipoprotein or LDL
(also referred to as “bad” cholesterol) is removed from the blood by receptor-mediated
endocytosis. In the human genetic disease familial hypercholesterolemia, the LDL receptors
are defective or missing entirely. People with this condition have life-threatening levels of
cholesterol in their blood, because their cells cannot clear the chemical from their blood.
In contrast to these methods of moving material into a cell is the process of exocytosis.
Exocytosis is the opposite of the processes discussed above in that its purpose is to expel
material from the cell into the extracellular fluid. A particle enveloped in membrane (for
example, a vesicle) fuses with the interior of the plasma membrane. This fusion opens the
membranous envelope to the exterior of the cell, and the particle is expelled into the
extracellular space (Figure 5.13). Hormones and neurotransmitters (chemicals released by
neuronal cells) are released by exocytosis, for example.
Figure 5.13 In exocytosis, a vesicle migrates to the plasma membrane, binds, and releases its contents to the
outside of the cell. (credit: modification of work by Mariana Ruiz Villarreal)
KEY TERMS
chemical gradient refers to the concentration of an ion or a molecule
concentration gradient occurs when a substance is more concentrated in one area than
another
diffusion the passive movement of molecules or particles along a concentration gradient
electrical gradient a difference of charge across the cell membrane
electrochemical gradient combination of chemical gradient and electrical gradient
endocytosis type of active transport that moves particles or even whole cells into a cell
extracellular fluid all the fluid that is outside of cells
exocytosis active process (requiring ATP) by which that expels material from a cell
facilitated diffusion form of passive transport across a biological membrane in which a
transporter protein facilitates the movement of an otherwise membrane-impermeable
molecule or ion across the plasma membrane down its concentration gradient
hydrophilic water-loving
hydrophobic water-hating
hypertonic solution contains a higher concentration of solute compared to another solution
hypotonic solution contains a lower concentration of solute compared to another solution
intercellular fluid located between cells
intracellular fluid located within the cell
isotonic solution contains the same concentration of solute compared to another
metabolic energy energy derived from chemical processes in the cell
metabolism the sum of chemical reactions that take place within each cell of a living
organism
phagocytosis active (requiring ATP) process by which large particles, such as cells, are taken
in by a cell
pinocytosis active process that takes in extracellular fluid and solutes
selective permeability property of a membrane that allows only some substance pass
across it unaided
tonicity a description of the relative solute concentration in a solution as compared to
another solution
REVISION QUESTIONS
Question 5.1
Which plasma membrane component can be either found on its surface or embedded in the
membrane structure?
A protein
B cholesterol
C carbohydrate
D phospholipid

Question 5.2
The tails of the phospholipids of the plasma membrane are composed
of ________ ________ and are ________ ________ ?
A phosphate groups; hydrophobic
B fatty acid groups; hydrophilic
C phosphate groups; hydrophilic
D fatty acid groups; hydrophobic

Question 5.3
Water moves via osmosis ________ ________ .
A throughout the cytoplasm
B from an area with a high concentration of other solutes to a lower one
C from an area with a low concentration of solutes to an area with a higher one
D from an area with a low concentration of water to one of higher concentration
Question 5.4
The diffusion of substances within a solution tends to move those substances
____________________ their ____________________ gradient.
A up; electrical
B up; electrochemical
C down; pressure
D down; concentration

Question 5.5
Which of the factors listed below affect the rate of diffusion?
A temperature
B how much ATP is required
C concentration gradient

Question 5.6
Active transport must function continuously because ________ ________
A plasma membranes wear out
B cells must be in constant motion
C facilitated transport opposes active transport
D diffusion is constantly moving the solutes in the other direction

Question 5.7
Where does the cell get energy for active transport processes?

Question 5.8
Why does osmosis occur?
Question 5.9
Ion pumps and phagocytosis are both examples of ____________________.
A endocytosis
B passive transport
C active transport
D facilitated diffusion

Question 5.10
Pinocytosis, endocytosis and exocytosis
A are passive types of transport
B active types of transport
ANSWERS
Question 5.1
A

Question 5.2
D

Question 5.3
C

Question 5.4
D

Question 5.5
A, C

Question 5.6
D

Question 5.7
The cell harvests energy from ATP produced by its own metabolism to power active
transport processes, such as pumps.

Question 5.8
Water moves through a semipermeable membrane in osmosis because there is a
concentration gradient across the membrane of solute and solvent. The solute cannot
effectively move to balance the concentration on both sides of the membrane, so water
moves to achieve this balance.

Question 5.9
C

Question 5.10
B
6| THE GENETICS OF LIFE

Figure 6.1 The scientists responsible for the discovery of the double helical structure of DNA. Francis Crick,
James Watson, Maurice Wilkins and Rosalind Franklin. In 1962, Watson, Crick and Wilkins received the Nobel
prize in Physiology and Medicine, but Franklin was barely acknowledged

By the end of this chapter, you will be able to:


• Describe the structure of nucleotides and nucleic acids
• Describe the structure of the nucleus,
• Explain the process of DNA replication
• Explain the process by which a cell builds proteins using the DNA code
Before embarking on our journey into the genetics of life, it is worthwhile at this stage to
consider again the four biomolecules (macromolecules) which make up cells. As was
introduced in the chapter 2, cells contain four major biomolecules, including nucleic acids
(DNA/RNA), proteins, lipids, and carbohydrates. These each have distinct structural and
biochemical properties that lend to their individual functions within the cell. Table 6.2
outlines some of these features for you. Collectively, they create the molecular players,
structures, and organelles that perform tasks within the cell.
This chapter will covers the structure and function of nucleic acids whereas the next chapter
discusses the other three biomolecules (macromolecules) in more detail.
Table 6.2. The four biomolecules have distinct structures and biochemical properties. Nucleic acids, proteins
and carbohydrates are constructed from their monomer units into larger structures that are necessary for
their function.

6.1 Nucleic acids


The nucleic acid family includes deoxyribonucleic acid (DNA) and ribonucleic acid (RNA).
DNA is the genetic code that serves as a blueprint for the transfer of information within the
cell. During the lifespan of an individual cell, the information stored within DNA sequences
will determine which RNA coding is produced and ultimately which protein sequences are
generated by the cell (Figure 6.3). This is known as the "central dogma of genetics".
Figure 6.3 The cell uses the DNA code to build protein, RNA being the essential intermediary

As the story starts with DNA and RNA let us first briefly examine their structure. The
monomer unit (or repeating unit) that is used to build nucleic acids is called a nucleotide.
Each nucleotide has three distinct components: a nitrogenous base, five-carbon sugar, and a
phosphate group (Figure 6.4a).
• Nitrogenous bases are broken down into two families, including pyrimidines with a
one-ringed structure (thymine (T), cytosine (C), and uracil (U)) and purines with a
two-ringed structure (adenine (A) and guanine (G)) (Figure 6.4b). DNA is coded by
ATCG nucleotides whereas RNA uses AUCG, exchanging U in place of T.
• The five-carbon sugar can either be deoxyribose (for DNA only) or ribose (for RNA
only). (See next chapter for structure of different sugars).
• The phosphate group is a phosphorous atom attached to four oxygen atoms.

Figure 6.4 (a) Nucleotides are composed of a nitrogenous base, five-carbon sugar, and phosphate group. (b)
DNA and RNA utilise five different nitrogenous bases. DNA may contain adenine (A), guanine (G),
thymine (T) and cytosine (C) whereas RNA replaces thymine (T) with uracil (U).

Nucleotides assemble to form long strands of the nucleic acids DNA or RNA. The DNA strand
can be visualised as a chain consisting of a sugar phosphate backbone with bases sticking
out at regular intervals from it (Figure 6.5).
Nucleic acid strands can also form weaker associations to other nucleic acid strands through
hydrogen bonding between the nitrogenous bases. Due to the structure and biochemical
nature of each nucleotide, only complementary base pairing can occur. This includes A
bonding with T and C bonding with G.

Figure 6.5 DNA structure Each strand is composed of a sugar-phosphate backbone with bases sticking out at
regular intervals. The complementary DNA is also show in this figure.
https://y12hb.files.wordpress.com/2013/03/dna.png

In nature DNA only occurs in the form of two complementary DNA strands. Two DNA
strands arrange in anti-parallel configuration as shown in Figures 6.5 and 6.6. The two DNA
chains are complementary, that is A on one chain will hydrogen bond with the T on the
other, whereas the C on one chain will hydrogen bond with G on the other. Note that
scientists (Figure 6.1) discovered that the two complementary strands did not just exist in
straight strands but formed the helical structure known as the double helix (Figure 6.6).
Figure 6.6 DNA In the DNA double helix, two strands attach via hydrogen bonds between the bases of
the component nucleotides.
Note that the RNA strand can also be visualised as chain with a sugar-phosphate
backbone with bases sticking out at regular intervals, except that the base T is replaced
with the base U (Figure 6.7)

Figure 6.7 RNA structure RNA is a similar molecule to DNA, with some subtle but important
differences.https://y12hb.wordpress.com/2013/03/27/structure-of-rna/

6.2 The Nucleus and DNA Replication


The nucleus is the largest and most prominent of a cell’s organelles (Figure 6.8). The nucleus
is generally considered the control centre of the cell because it stores all of the genetic
instructions for manufacturing proteins. Interestingly, some cells in the body, such as
skeletal muscle cells, contain more than one nucleus (Figure 6.9); a skeletal muscle cell is
termed multinucleate. Other cells, such as mammalian red blood cells (RBCs), do not contain
nuclei at all. RBCs eject their nuclei as they mature, making space for the large numbers of
haemoglobin molecules that carry oxygen throughout the body. Without nuclei, the life
span of RBCs is short, and so the body must produce new ones constantly.
Figure 6.8 The Nucleus The nucleus is the control centre of the cell. The nucleus of living cells contains the
genetic material that determines the entire structure and function of that cell.

Figure 6.9 Multinucleate Muscle Cell Unlike cardiac muscle cells and smooth muscle cells, which have a
single nucleus, a skeletal muscle cell contains many nuclei, and is referred to as “multinucleated.” These
muscle cells are long and fibrous (often referred to as muscle fibres). During development, many smaller cells
fuse to form a mature muscle fibre. The nuclei of the fused cells are conserved in the mature cell, thus
imparting a multinucleate characteristic to mature muscle cells. LM × 104.3. (Micrograph provided by the
Regents of University of Michigan Medical School © 2012)
Inside the nucleus lies the blueprint that dictates everything a cell will do and all of the
products it will make. This information is stored within DNA. The nucleus sends
“commands” to the cell via molecular messengers that translate the information from DNA.
Each cell in your body (with the exception of germ cells, that is, egg and sperm) contains the
complete set of your DNA. When a cell divides, the DNA must be duplicated so that the each
new cell receives a full complement of DNA. The following section will explore the structure
of the nucleus and its contents, as well as the process of DNA replication.

Organization of the Nucleus and Its DNA


Like most other cellular organelles, the nucleus is surrounded by a membrane called
the nuclear envelope. This membranous covering consists of two adjacent lipid bilayers with
a thin fluid space in between them. Spanning these two bilayers are nuclear pores.
A nuclear pore is a tiny passageway for the passage of proteins, RNA, and solutes between
the nucleus and the cytoplasm.
Inside the nuclear envelope is a gel-like nucleoplasm with solutes that include the building
blocks of nucleic acids. There also can be a dark-staining mass often visible under a simple
light microscope, called a nucleolus . As we learnt in Chapter 2, the nucleolus is a region of
the nucleus that is responsible for manufacturing the RNA necessary for construction of
ribosomes. Once synthesised, newly made ribosomal subunits exit the cell’s nucleus through
the nuclear pores. These ribosomes attach to sections of the endoplasmic reticulum to give
it a studded appearance (refer to Chapter 2).
The genetic instructions that are used to build and maintain an organism are arranged in an
orderly manner in strands of DNA. Within the nucleus are threads of chromatin composed
of DNA and associated proteins (Figure 6.10). Along the chromatin threads, the DNA is
wrapped around a set of histone proteins. A nucleosome is a single, wrapped DNA-
histone complex. Multiple nucleosomes along the entire molecule of DNA appear like a
beaded necklace, in which the string is the DNA and the beads are the associated
histones.When a cell is in the process of division, the chromatin condenses
into chromosomes, so that the DNA can be safely transported to the “daughter cells.”
The chromosome is composed of DNA and proteins; it is the condensed form of chromatin.
It is estimated that humans have almost 22,000 genes distributed on 46 chromosomes.

DNA Replication
In order for an organism to grow, develop, and maintain its health, cells must reproduce
themselves by dividing to produce two new daughter cells, each with the full complement of
DNA as found in the original cell. Billions of new cells are produced in an adult human every
day. The division time of different cell types varies. Epithelial cells of the skin
and gastrointestinal lining, for instance, divide very frequently to replace those that are
constantly being rubbed off of the surface by friction.

Figure 6.10 DNA Macrostructure Strands of DNA are wrapped around supporting histones. These proteins
are increasingly bundled and condensed into chromatin, which is packed tightly into chromosomes when the
cell is ready to divide.

The particular sequence of bases along the DNA molecule determines the genetic code.
Therefore, if the two complementary strands of DNA were pulled apart, you could infer the
order of the bases in one strand from the bases in the other, complementary strand. For
example, if one strand has a region with the sequence AGTGCCT, then the sequence of the
complementary strand would be TCACGGA.
DNA replication is the copying of DNA that occurs before cell division can take place. After a
great deal of debate and experimentation, the general method of DNA replication was
deduced in 1958 by two scientists in California, Matthew Meselson and Franklin Stahl. This
method is illustrated in Figure 6.11 and described below. Note that several enzymes take
part in DNA replication and that the suffix "ase" is used to denote an enzyme
Figure 6.11 DNA Replication DNA replication faithfully duplicates the entire genome of the cell. During
DNA replication, a number of different enzymes work together to pull apart the two strands so each strand can
be used as a template to synthesise new complementary strands. The two new daughter DNA molecules each
contain one preexisting strand and one newly synthesised strand. Thus, DNA replication is said to be
“semiconservative.”

Stage 1: Initiation. The two complementary strands are separated, much like unzipping a
zipper. Special enzymes, including helicase, untwist and separate the two strands of DNA.
Stage 2: Elongation. Each strand becomes a template along which a new complementary
strand is built. DNA polymerase brings in the correct bases to complement the template
strand, synthesising a new strand base by base. A DNA polymerase is an enzyme that adds
free nucleotides to the end of a chain of DNA, making a new double strand. This growing
strand continues to be built until it has fully complemented the template strand.
Stage 3: Termination. Once the two original strands are bound to their own, finished,
complementary strands, DNA replication is stopped and the two new identical DNA
molecules are complete.
Each new DNA molecule contains one strand from the original molecule and one newly
synthesised strand. The term for this mode of replication is “semiconservative,” because
half of the original DNA molecule is conserved in each new DNA molecule. This process
continues until the cell’s entire genome, the entire complement of an organism’s DNA,
is replicated. As you might imagine, it is very important that DNA replication take place
precisely so that new cells in the body contain the exact same genetic material as their
parent cells. Mistakes made during DNA replication, such as the accidental addition of an
inappropriate nucleotide, have the potential to render a gene dysfunctional or useless.
Fortunately, there are mechanisms in place to minimise such mistakes. A DNA proofreading
process enlists the help of special enzymes that scan the newly synthesised molecule for
mistakes and corrects them. Once the process of DNA replication is complete, the cell is
ready to divide.

6.3 | Protein Synthesis


It was mentioned earlier that DNA provides a “blueprint” for the cell structure and
physiology. This refers to the fact that DNA contains the information necessary for the cell
to build one very important type of molecule: the protein. Most structural components of
the cell are made up, at least in part, by proteins and virtually all the functions that a cell
carries out are completed with the help of proteins. One of the most important classes of
proteins is enzymes, which help speed up necessary biochemical reactions that take place
inside the cell. Some of these critical biochemical reactions include building larger molecules
from smaller components (such as occurs during DNA replication) and breaking down larger
molecules into smaller components (such as when harvesting chemical energy from
nutrient molecules). Whatever the cellular process may be, it is almost sure to involve
proteins. Protein synthesis begins with genes. A gene is a functional segment of DNA that
provides the genetic information necessary to build a protein. Each particular gene provides
the code necessary to construct a particular protein. Gene expression, which transforms the
information coded in a gene to a final gene product, ultimately dictates the structure and
function of a cell by determining which proteins are made.
The interpretation of genes works in the following way. Recall that proteins are polymers, or
chains, of many amino acid building blocks. The sequence of bases in a gene (that is, its
sequence of A, T, C, G nucleotides) translates to an amino acid sequence. A triplet is a
section of three DNA bases in a row that codes for a specific amino acid. Similar to the way
in which the three-letter code d-o-g signals the image of a dog, the three-letter DNA base
code signals the use of a particular amino acid. For example, the DNA triplet CAC (cytosine,
adenine, and cytosine) specifies the amino acid valine. Therefore, a gene, which is
composed of multiple triplets in a unique sequence, provides the code to build an entire
protein, with multiple amino acids in the proper sequence (Figure 6.12). The mechanism by
which cells turn the DNA code into a protein product is a two-step process, with an RNA
molecule as the intermediate.
Figure 6.12 The Genetic Code DNA holds all of the genetic information necessary to build a cell’s proteins.
The nucleotide sequence of a gene is ultimately translated into an amino acid sequence of the gene’s
corresponding protein.

Step 1: From DNA to RNA: Transcription


DNA is housed within the nucleus, and protein synthesis takes place in the cytoplasm, thus
there must be some sort of intermediate messenger that leaves the nucleus and manages
protein synthesis. This intermediate messenger is messenger RNA (mRNA), a single-
stranded nucleic acid that carries a copy of the genetic code for a single gene out of the
nucleus and into the cytoplasm where it is used to produce proteins.
There are several different types of RNA, each having different functions in the cell. The
structure of RNA is similar to DNA with a few small exceptions. For one thing, unlike DNA,
most types of RNA, including mRNA, are single-stranded and contain no complementary
strand. Second, the ribose sugar in RNA contains an additional oxygen atom compared with
DNA. Finally, instead of the base thymine, RNA contains the base uracil. This means that
adenine will always pair up with uracil during the protein synthesis process.
Gene expression begins with the process called transcription, which is the synthesis of a
strand of mRNA that is complementary to the gene of interest. This process is called
transcription because the mRNA is like a transcript, or copy, of the gene’s DNA code.
Transcription begins in a fashion somewhat like DNA replication, in that a region of DNA
unwinds and the two strands separate, however, only that small portion of the DNA will be
split apart. The triplets within the gene on this section of the DNA molecule are used as the
template to transcribe the complementary strand of RNA (Figure 6.13). A codon is a three-
base sequence of mRNA, so-called because they directly encode amino acids. Like DNA
replication, there are three stages to transcription: initiation, elongation, and termination.
Figure 6.13 Transcription: from DNA to mRNA In the first of the two stages of making protein from DNA, a gene
on the DNA molecule is transcribed into a complementary mRNA molecule.

Stage 1: Initiation. A region at the beginning of the gene called a promoter—a particular
sequence of nucleotides—triggers the start of transcription.
Stage 2: Elongation. Transcription starts when RNA polymerase unwinds the DNA segment.
One strand, referred to as the coding strand, becomes the template with the genes to be
coded. The polymerase then aligns the correct nucleic acid (A, C, G, or U) with its
complementary base on the coding strand of DNA. RNA polymerase is an enzyme that adds
new nucleotides to a growing strand of RNA. This process builds a strand of mRNA.
Stage 3: Termination. When the polymerase has reached the end of the gene, one of three
specific triplets (UAA, UAG, or UGA) codes a “stop” signal, which triggers the enzymes to
terminate transcription and release the mRNA transcript.
After being modified, the mRNA molecule leaves the nucleus and proceeds to protein
synthesis.

Step 2: From RNA to Protein: Translation


Like translating a book from one language into another, the codons on a strand of mRNA
must be translated into the amino acid alphabet of proteins. Translation is the process of
synthesising a chain of amino acids called a polypeptide, which is what the larger protein
molecules are made of. Translation requires two major aids: first, a “translator,” the
molecule that will conduct the translation, and second, a specific place where the mRNA
strand is translated into a new protein, like the translator’s “desk.” Both of these
requirements are fulfilled by other types of RNA. The "desk" on which translation takes
place is the ribosome. The ribosome provides a place for translation, bringing together and
aligning the mRNA molecule with the molecular “translators” that must decipher its code.
The other major requirement for protein synthesis is the translator molecules that physically
“read” the mRNA codons. Transfer RNA (tRNA) is a type of RNA that ferries the appropriate
corresponding amino acids to the ribosome, and attaches each new amino acid to the last,
building the polypeptide chain one-by-one. Thus tRNA transfers specific amino acids from
the cytoplasm to a growing polypeptide. The tRNA molecules must be able to recognise the
codons on mRNA and match them with the correct amino acid. The tRNA is modified for this
function. On one end of its structure is a binding site for a specific amino acid. On the other
end is a base sequence that matches the codon specifying its particular amino acid. This
sequence of three bases on the tRNA molecule is called an anticodon. For example, a tRNA
responsible for shuttling the amino acid glycine contains a binding site for glycine on one
end. On the other end it contains an anticodon that complements the glycine codon (GGA is
a codon for glycine, and so the tRNAs anticodon would read CCU). Equipped with its
particular cargo and matching anticodon, a tRNA molecule can read its recognised mRNA
codon and bring the corresponding amino acid to the growing chain (Figure 6.14).
Much like the processes of DNA replication and transcription, translation consists of three
main stages: initiation, elongation, and termination. Initiation takes place with the binding
of a ribosome to an mRNA transcript. The elongation stage involves the recognition of a
tRNA anticodon with the next mRNA codon in the sequence. Once the anticodon and codon
sequences are bound (remember, they are complementary base pairs), the tRNA presents
its amino acid cargo and the growing polypeptide strand is attached to this next amino acid.
This attachment takes place with the assistance of various enzymes and requires energy.
The tRNA molecule then releases the mRNA strand, the mRNA strand shifts one codon
over in the ribosome, and the next appropriate tRNA arrives with its matching anticodon.
This process continues until the final codon on the mRNA is reached which provides a “stop”
message that signals termination of translation and triggers the release of the complete,
newly synthesised protein. Thus, a gene within the DNA molecule is transcribed into mRNA,
which is then translated into a protein product (Figure 6.15).
Commonly, an mRNA transcription will be translated simultaneously by several adjacent
ribosomes. This increases the efficiency of protein synthesis. A single ribosome might
translate an mRNA molecule in approximately one minute; so multiple ribosomes aboard a
single transcript could produce multiple times the number of the same protein in the
same minute.
Figure 6.14 Translation from RNA to Protein During translation, the mRNA transcript is “read” by a
functional complex consisting of the ribosome and tRNA molecules. tRNAs bring the appropriate amino acids
in sequence to the growing polypeptide chain by matching their anti-codons with codons on the mRNA strand.
Figure 6.15 From DNA to Protein: Transcription through Translation Transcription within the cell
nucleus produces an mRNA molecule, which is modified and then sent into the cytoplasm for translation. The
transcript is decoded into a protein with the help of a ribosome and tRNA molecules.
KEY TERMS
anticodon consecutive sequence of three nucleotides on a tRNA molecule that is
complementary to a specific codon on an mRNA molecule
chromatin substance consisting of DNA and associated proteins
codon consecutive sequence of three nucleotides on an mRNA molecule that corresponds
to a specific amino acid
DNA polymerase enzyme that functions in adding new nucleotides to a growing strand of
DNA during DNA replication
DNA replication process of duplicating a molecule of DNA
gene functional length of DNA that provides the genetic information necessary to build a
protein
gene expression active interpretation of the information coded in a gene to produce a
functional gene product
genome entire complement of an organism’s DNA; found within virtually every cell
helicase enzyme which untwists and separates the two strands of DNA in the first stage of
replication
histone family of proteins that associate with DNA in the nucleus to form chromatin
messenger RNA (mRNA) nucleotide molecule that serves as an intermediate in the genetic
code between DNA and protein
nucleic acids consist of either one or two long chains of repeating units called nucleotides,
which consist of a nitrogen base (a purine or pyrimidine) attached to a sugar phosphate. The
two main nucleic acids are DNA and RNA.
nuclear envelope membrane that surrounds the nucleus; consisting of a double lipid-bilayer
nuclear pore one of the small, protein-lined openings found scattered throughout the
nuclear envelope
nucleolus small region of the nucleus that functions in ribosome synthesis
nucleosome unit of chromatin consisting of a DNA strand wrapped around histone proteins
nucleus cell’s central organelle; contains the cell’s DNA
polypeptide chain of amino acids linked by peptide bonds
promoter region of DNA that signals transcription to begin at that site within the gene
ribosomal RNA (rRNA) RNA that makes up the subunits of a ribosome
ribosome cellular organelle that functions in protein synthesis
RNA polymerase enzyme that unwinds DNA and then adds new nucleotides to a growing
strand of RNA for the transcription phase of protein synthesis
transcription process of producing an mRNA molecule that is complementary to a particular
gene of DNA
transcription factor one of the proteins that regulate the transcription of genes
transfer RNA (tRNA) molecules of RNA that serve to bring amino acids to a growing
polypeptide strand and properly place them into the sequence
translation process of producing a protein from the nucleotide sequence code of an mRNA
transcript
triplet consecutive sequence of three nucleotides on a DNA molecule that, when
transcribed into an mRNA codon, corresponds to a particular amino acid
REVISION QUESTIONS
Question 6.1
Which of the following sequences on a DNA molecule would be complementary to
GCTTATAT?
A TAGGCGCG
B ATCCGCGC
C CGAATATA
D TGCCTCTC

Question 6.2
Place the following structures in order from least to most complex organization: chromatin,
nucleosome, DNA, chromosome
A DNA, nucleosome, chromatin, chromosome
B nucleosome, DNA, chromosome, chromatin
C DNA, chromatin, nucleosome, chromosome
D nucleosome, chromatin, DNA, chromosome

Question 6.3
Which of the following is part of the elongation step of DNA synthesis?
A pulling apart the two DNA strands
B attaching complementary nucleotides to the template strand
C untwisting the DNA helix
D none of the above
Question 6.4
Which of the following is not a difference between DNA and RNA?
A DNA contains thymine whereas RNA contains uracil
B DNA contains deoxyribose and RNA contains ribose
C DNA contains alternating sugar-phosphate molecules whereas RNA does not contain
sugars
D RNA is single stranded and DNA is double stranded

Question 6.5
Transcription and translation take place in the ____________________ and
____________________, respectively.
A nucleus; cytoplasm
B nucleolus; nucleus
C nucleolus; cytoplasm
D cytoplasm; nucleus

Question 6.6
How many “letters” of an RNA molecule, in sequence, does it take to provide the code for a
single amino acid?
A1
B2
C3
D4
Question 6.7
DNA codes for
A carbohydrates
B lipids
C proteins
D none of the above

Question 6.8
Transcription is the process by which an mRNA transcript is made of a gene's DNA code
A TRUE
B FALSE

Question 6.9
Which of the following statements is/are false?
A Each codon of mRNA corresponds to an amino acid
B mRNA is found only in the cytoplasm
C tRNA is found only in the cytoplasm
D Translation of mRNA to a protein is carried out by ribosomes in the cytoplasm

Question 6.10
Which of the following statements is/are true?
A Amino acids are the building blocks of proteins
B During protein synthesis, peptide bonds are formed between adjacent amino acids
C DNA is coded by ATCG nucleotides whereas RNA uses AUCG, exchanging U in place of T
D The sugar-phosphate backbone is present in both RNA and DNA
Question 6.11
Place the following in the order that they occur: translation, DNA replication, transcription,
protein synthesis
A translation, DNA replication, transcription, protein synthesis
B DNA replication, transcription, translation, protein synthesis
C DNA replication, translation, transcription, protein synthesis
D transcription, DNA replication, translation, protein synthesis
ANSWERS
Question 6.1
C

Question 6.2
A

Question 6.3
B

Question 6.4
C

Question 6.5
A

Question 6.6
C

Question 6.7
C

Question 6.8
A

Question 6.9
B

Question 6.10
A, B, C, D

Question 6.11
B
Chapter 7 | The essential
compounds of life

Figure 7.1 Foods such as bread, fruit, and cheese are rich sources of biological macromolecules. (credit:
modification of work by Bengt Nyman)

After studying this chapter you will be able to:


• Identify the building blocks of carbohydrates, lipids and proteins and describe how
they join together
• Describe the structure of carbohydrates and identify the primary functions in the
body
• Describe the structure of lipids and list the four types important in human
functioning
• Describe the structure of proteins and list their functions in the body
Introduction
The large molecules necessary for life are called biological macromolecules (biomolecules).
There are four major classes of biological biomolecules (carbohydrates, lipids, proteins, and
nucleic acids), and each is an important component of the cell, performing a wide array of
functions. Combined, these molecules make up the majority of a cell’s mass. Biomolecules
are organic, meaning that they contain carbon. In addition, they may contain hydrogen,
oxygen, nitrogen, phosphorus, sulfur, and additional minor elements. Table 7.1 lists the four
classes of biomolecules along with their major functions. Nucleic acids have been discussed
in Chapter 6 so in this chapter, we will explore the structure of proteins, carbohydrate and
lipids and discuss how these molecules are formed and the functions they fulfil.

Table 7.2. The four main classes of biomolecules

Organic compounds
All biomolecules contain the elements carbon, hydrogen and oxygen. Nitrogen, sulfur and
phosphorus can also be present. Elements have been discussed in detail in Chapter 3 but
we will briefly revise the properties of carbon before discussing the classes of biomolecules.
Carbon
It is often said that life is “carbon-based.” This means that carbon atoms, bonded to other
carbon atoms or other elements, form the fundamental components of many, if not most,
of the molecules found uniquely in living things. Other elements play important roles in
biological molecules, but carbon certainly qualifies as the “foundation” element for
molecules in living things. It is the bonding properties of carbon atoms that are responsible
for its important role.

Carbon Bonding
Carbon contains four electrons in its outer shell. Therefore, it can form four covalent bonds
with other atoms or molecules. The simplest organic carbon molecule is methane (CH4), in
which four hydrogen atoms bind to a carbon atom (Figure 7.3).

Figure 7.3 Carbon can form four covalent bonds to create an organic molecule. The simplest carbon molecule
is methane (CH4), depicted here.

However, structures that are more complex are made using carbon. Any of the hydrogen
atoms can be replaced with another carbon atom covalently bonded to the first carbon
atom. In this way, long and branching chains of carbon compounds can be made (Figure
7.4a). The carbon atoms may bond with atoms of other elements, such as nitrogen, oxygen,
and phosphorus (Figure 7.4b). The molecules may also form rings, which themselves can
link with other rings (Figure 7.4c). This diversity of molecular forms accounts for the
diversity of functions of the biological macromolecules and is based to a large degree on the
ability of carbon to form multiple bonds with itself and other atoms.
Figure 7.4 These examples show three molecules (found in living organisms) that contain carbon atoms
bonded in various ways to other carbon atoms and the atoms of other elements. (a) This molecule of stearic
acid has a long chain of carbon atoms. (b) Glycine, a component of proteins, contains carbon, nitrogen, oxygen,
and hydrogen atoms. (c) Glucose, a sugar, has a ring of carbon atoms and one oxygen atom.

Carbohydrates
Carbohydrates are macromolecules with which most consumers are somewhat familiar. To
lose weight, some individuals adhere to “low-carb” diets. Athletes, in contrast, often “carb-
load” before important competitions to ensure that they have sufficient energy to compete
at a high level. Carbohydrates are, in fact, an essential part of our diet; grains, fruits, and
vegetables are all natural sources of carbohydrates. Carbohydrates provide energy to the
body, particularly through glucose, a simple sugar. Carbohydrates also have other important
functions in humans, animals, and plants.
Carbohydrates can be represented by the formula (CH2O)n, where n is the number of carbon
atoms in the molecule. In other words, the ratio of carbon to hydrogen to oxygen is 1:2:1 in
carbohydrate molecules. Carbohydrates are classified into three
subtypes: monosaccharides, disaccharides, and polysaccharides.
Monosaccharides (mono- = “one”; sacchar- = “sweet”) are simple sugars, the most common
of which is glucose. In monosaccharides, the number of carbon atoms usually ranges from
three to six. Most monosaccharide names end with the suffix -ose. Depending on the
number of carbon atoms in the sugar, they may be known as trioses (three carbon atoms),
pentoses (five carbon atoms), and hexoses (six carbon atoms).
Monosaccharides may exist as a linear chain or as ring-shaped molecules; in aqueous
solutions, they are usually found in the ring form.
The chemical formula for glucose is C6H12O6. In most living species, glucose is an important
source of energy. During cellular respiration, energy is released from glucose, and that
energy is used to help make adenosine triphosphate (ATP). Plants synthesise glucose using
carbon dioxide and water by the process of photosynthesis, and the glucose, in turn, is used
for the energy requirements of the plant. The excess synthesised glucose is often stored as
starch that is broken down by other organisms that feed on plants.
Galactose (part of lactose, or milk sugar) and fructose (found in fruit) are other common
monosaccharides. Although glucose, galactose, and fructose all have the same chemical
formula (C6H12O6), they differ structurally and chemically (and are known as isomers)
because of differing arrangements of atoms in the carbon chain (Figure 7.5).
Disaccharides (di- = “two”) form when two monosaccharides undergo a dehydration
reaction (a reaction in which the removal of a water molecule occurs). During this process,
the hydroxyl group (–OH) of one monosaccharide combines with a hydrogen atom of
another monosaccharide, releasing a molecule of water (H2O) and forming a covalent bond
between atoms in the two sugar molecules.
Common disaccharides include lactose, maltose, and sucrose. Lactose is a disaccharide
consisting of the monomers glucose and galactose. It is found naturally in milk. Maltose, or
malt sugar, is a disaccharide formed from a dehydration reaction between two glucose
molecules. The most common disaccharide is sucrose, or table sugar, which is composed of
the monomers glucose and fructose.
A long chain of monosaccharides linked by covalent bonds is known as
a polysaccharide (poly- = “many”). The chain may be branched or unbranched, and it may
contain different types of monosaccharides. Polysaccharides may be very large molecules.
Starch, glycogen, cellulose, and chitin are examples of polysaccharides.
Figure 7.5 These examples show three molecules (found in living organisms) that contain carbon atoms
bonded in various ways to other carbon atoms and the atoms of other elements. (a) This molecule of stearic
acid has a long chain of carbon atoms. (b) Glycine, a component of proteins, contains carbon, nitrogen, oxygen,
and hydrogen atoms. (c) Glucose, a sugar, has a ring of carbon atoms and one oxygen atom.

Starch is the stored form of sugars in plants and is made up of amylose and amylopectin
(both polymers of glucose). Plants are able to synthesise glucose, and the excess glucose is
stored as starch in different plant parts, including roots and seeds. The starch that is
consumed by animals is broken down into smaller molecules, such as glucose. The cells can
then absorb the glucose.
Glycogen is the storage form of glucose in humans and other vertebrates, and is made up of
monomers of glucose. Glycogen is the animal equivalent of starch and is a highly branched
molecule usually stored in liver and muscle cells. Whenever glucose levels decrease,
glycogen is broken down to release glucose.
Cellulose is one of the most abundant natural biopolymers. The cell walls of plants are
mostly made of cellulose, which provides structural support to the cell. Wood and paper are
mostly cellulosic in nature. Cellulose is made up of glucose monomers that are linked by
bonds between particular carbon atoms in the glucose molecule.
Every other glucose monomer in cellulose is flipped over and packed tightly as extended
long chains. This gives cellulose its rigidity and high tensile strength—which is so important
to plant cells. Cellulose passing through our digestive system is called dietary fibre. While
the glucose-glucose bonds in cellulose cannot be broken down by human digestive enzymes,
herbivores such as cows, buffalos, and horses are able to digest grass that is rich in cellulose
and use it as a food source. In these animals, certain species of bacteria reside in the rumen
(part of the digestive system of herbivores) and secrete the enzyme cellulase. The appendix
also contains bacteria that break down cellulose, giving it an important role in the digestive
systems of ruminants. Cellulases can break down cellulose into glucose monomers that can
be used as an energy source by the animal.
Carbohydrates serve other functions in different animals. Arthropods, such as insects,
spiders, and crabs, have an outer skeleton, called the exoskeleton, which protects their
internal body parts. This exoskeleton is made of the biological macromolecule chitin, which
is a nitrogenous carbohydrate. It is made of repeating units of a modified sugar containing
nitrogen.
Thus, through differences in molecular structure, carbohydrates are able to serve the very
different functions of energy storage (starch and glycogen) and structural support and
protection (cellulose and chitin) (Figure 7.6).

Figure 7.6 Although their structures and functions differ, all polysaccharide carbohydrates are made up of
monosaccharides and have the chemical formula (CH2O)n.
CAREERS IN ACTION

Registered Dietitian
Obesity is a worldwide health concern, and many diseases, such as diabetes and heart
disease, are becoming more prevalent because of obesity. This is one of the reasons why
registered dietitians are increasingly sought after for advice. Registered dietitians help plan
food and nutrition programs for individuals in various settings. They often work with
patients in health-care facilities, designing nutrition plans to prevent and treat diseases. For
example, dietitians may teach a patient with diabetes how to manage blood-sugar levels by
eating the correct types and amounts of carbohydrates. Dietitians may also work in nursing
homes, schools, and private practices.
To become a registered dietitian, one needs to earn at least a bachelor’s degree in dietetics,
nutrition, food technology, or a related field. In addition, registered dietitians must
complete a supervised internship program and pass a national exam. Those who pursue
careers in dietetics take courses in nutrition, chemistry, biochemistry, biology, microbiology,
and human physiology. Dietitians must become experts in the chemistry and functions of
food (proteins, carbohydrates, and fats).

Lipids
Lipids include a diverse group of compounds that are united by a common
feature. Lipids are hydrophobic (“water-fearing”), or insoluble in water, because they are
nonpolar molecules. This is because they are hydrocarbons that include only nonpolar
carbon-carbon or carbon-hydrogen bonds. Lipids perform many different functions in a cell.
Cells store energy for long-term use in the form of lipids called fats. Lipids also provide
insulation from the environment for plants and animals (Figure 7.7). For example, they help
keep aquatic birds and mammals dry because of their water-repelling nature. Lipids are also
the building blocks of many hormones and are an important constituent of the plasma
membrane. Lipids include fats, oils, waxes, phospholipids, and steroids.
A fat molecule, such as a triglyceride, consists of two main components—glycerol and fatty
acids. Glycerol is an organic compound with three carbon atoms, five hydrogen atoms, and
three hydroxyl (–OH) groups. Fatty acids have a long chain of hydrocarbons to which an
acidic carboxyl group is attached, hence the name “fatty acid.” The number of carbons in
the fatty acid may range from 4 to 36; most common are those containing 12–18 carbons. In
a fat molecule, a fatty acid is attached to each of the three oxygen atoms in the –OH groups
of the glycerol molecule with a covalent bond (Figure 7.8).
Figure 7.7 Hydrophobic lipids in the fur of aquatic mammals, such as this river otter, protect them from the
elements. (credit: Ken Bosma)

Figure 7.8 Lipids include fats, such as triglycerides, which are made up of fatty acids and glycerol,
phospholipids, and steroids.
During this covalent bond formation, three water molecules are released. The three fatty
acids in the fat may be similar or dissimilar. These fats are also called triglycerides because
they have three fatty acids. Some fatty acids have common names that specify their origin.
For example, palmitic acid, a saturated fatty acid, is derived from the palm tree. Arachidic
acid is derived from Arachis hypogaea, the scientific name for peanuts.
Fatty acids may be saturated or unsaturated. In a fatty acid chain, if there are only single
bonds between neighbouring carbons in the hydrocarbon chain, the fatty acid is
saturated. Saturated fatty acids are saturated with hydrogen; in other words, the number of
hydrogen atoms attached to the carbon skeleton is maximised.
When the hydrocarbon chain contains a double bond, the fatty acid is an unsaturated fatty
acid.
Most unsaturated fats are liquid at room temperature and are called oils. If there is one
double bond in the molecule, then it is known as a monounsaturated fat (e.g., olive oil), and
if there is more than one double bond, then it is known as a polyunsaturated fat (e.g.,
canola oil).
Saturated fats tend to get packed tightly and are solid at room temperature. Animal fats
with stearic acid and palmitic acid contained in meat, and the fat with butyric acid contained
in butter, are examples of saturated fats. Mammals store fats in specialised cells called
adipocytes, where globules of fat occupy most of the cell. In plants, fat or oil is stored in
seeds and is used as a source of energy during embryonic development.
Unsaturated fats or oils are usually of plant origin and contain unsaturated fatty acids. The
double bond causes a bend or a “kink” that prevents the fatty acids from packing tightly,
keeping them liquid at room temperature. Olive oil, corn oil, canola oil, and cod liver oil are
examples of unsaturated fats. Unsaturated fats help to improve blood cholesterol levels,
whereas saturated fats contribute to plaque formation in the arteries, which increases the
risk of a heart attack.
In the food industry, oils are artificially hydrogenated to make them semi-solid, leading to
less spoilage and increased shelf life. Simply speaking, hydrogen gas is bubbled through oils
to solidify them. During this hydrogenation process, double bonds of the cis-conformation in
the hydrocarbon chain may be converted to double bonds in the trans-conformation. This
forms a trans-fat from a cis-fat. The orientation of the double bonds affects the chemical
properties of the fat (Figure 7.9).
Figure 7.9 During the hydrogenation process, the orientation around the double bonds is changed, making a
trans-fat from a cis-fat. This changes the chemical properties of the molecule.

Margarine, some types of peanut butter, and shortening are examples of artificially
hydrogenated trans-fats. Recent studies have shown that an increase in trans-fats in the
human diet may lead to an increase in levels of low-density lipoprotein (LDL), or “bad”
cholesterol, which, in turn, may lead to plaque deposition in the arteries, resulting in heart
disease. Many fast food restaurants have recently eliminated the use of trans-fats, and U.S.
food labels are now required to list their trans-fat content.
Essential fatty acids are fatty acids that are required but not synthesised by the human
body. Consequently, they must be supplemented through the diet. Omega-3 fatty acids fall
into this category and are one of only two known essential fatty acids for humans (the other
being omega-6 fatty acids). They are a type of polyunsaturated fat and are called omega-3
fatty acids because the third carbon from the end of the fatty acid participates in a double
bond.
Salmon, trout, and tuna are good sources of omega-3 fatty acids. Omega-3 fatty acids are
important in brain function and normal growth and development. They may also prevent
heart disease and reduce the risk of cancer.
Like carbohydrates, fats have received a lot of bad publicity. It is true that eating an excess
of fried foods and other “fatty” foods leads to weight gain. However, fats do have important
functions. Fats serve as long-term energy storage. They also provide insulation for the body.
Therefore, “healthy” unsaturated fats in moderate amounts should be consumed on a
regular basis.
Phospholipids are the major constituent of the plasma membrane. Like fats, they are
composed of fatty acid chains attached to a glycerol or similar backbone. Instead of three
fatty acids attached, however, there are two fatty acids and the third carbon of the glycerol
backbone is bound to a phosphate group. The phosphate group is modified by the addition
of an alcohol.
A phospholipid has both hydrophobic and hydrophilic regions. The fatty acid chains are
hydrophobic and exclude themselves from water, whereas the phosphate is hydrophilic and
interacts with water.
Cells are surrounded by a membrane, which has a bilayer of phospholipids. The fatty acids
of phospholipids face inside, away from water, whereas the phosphate group can face
either the outside environment or the inside of the cell, which are both aqueous.

Steroids and Waxes


Unlike the phospholipids and fats discussed earlier, steroids have a ring structure. Although
they do not resemble other lipids, they are grouped with them because they are also
hydrophobic. All steroids have four, linked carbon rings and several of them, like
cholesterol, have a short tail.
Cholesterol is a steroid. Cholesterol is mainly synthesised in the liver and is the precursor of
many steroid hormones, such as testosterone and estradiol. It is also the precursor of
vitamins E and K. Cholesterol is the precursor of bile salts, which help in the breakdown of
fats and their subsequent absorption by cells. Although cholesterol is often spoken of in
negative terms, it is necessary for the proper functioning of the body. It is a key component
of the plasma membranes of animal cells.
Waxes are made up of a hydrocarbon chain with an alcohol (–OH) group and a fatty acid.
Examples of animal waxes include beeswax and lanolin. Plants also have waxes, such as the
coating on their leaves, that helps prevent them from drying out.
CONCEPT IN ACTION

For an additional perspective on lipids, explore “Biomolecules: The Lipids” through


this interactive animation (http://openstaxcollege.org/l/lipids) .

Proteins
Proteins are one of the most abundant organic molecules in living systems and have the
most diverse range of functions of all macromolecules. Proteins may be structural,
regulatory, contractile, or protective; they may serve in transport, storage, or membranes;
or they may be toxins or enzymes. Each cell in a living system may contain thousands of
different proteins, each with a unique function. Their structures, like their functions, vary
greatly. They are all, however, polymers of amino acids, arranged in a linear sequence.
The functions of proteins are very diverse because there are 20 different chemically distinct
amino acids that form long chains, and the amino acids can be in any order. For example,
proteins can function as enzymes or hormones. Enzymes, which are produced by living cells,
are catalysts in biochemical reactions (like digestion) and are usually proteins. Each enzyme
is specific for the substrate (a reactant that binds to an enzyme) upon which it acts. Enzymes
can function to break molecular bonds, to rearrange bonds, or to form new bonds. An
example of an enzyme is salivary amylase, which breaks down amylose, a component of
starch.
Hormones are chemical signalling molecules, usually proteins or steroids, secreted by an
endocrine gland or group of endocrine cells that act to control or regulate specific
physiological processes, including growth, development, metabolism, and reproduction. For
example, insulin is a protein hormone that maintains blood glucose levels.
Proteins have different shapes and molecular weights; some proteins are globular in shape
whereas others are fibrous in nature. For example, haemoglobin is a globular protein, but
collagen, found in our skin, is a fibrous protein. Protein shape is critical to its function.
Changes in temperature, pH, and exposure to chemicals may lead to permanent changes in
the shape of the protein, leading to a loss of function or denaturation (to be discussed in
more detail later). All proteins are made up of different arrangements of the same 20 kinds
of amino acids.
Amino acids are the monomers that make up proteins. Each amino acid has the same
fundamental structure, which consists of a central carbon atom bonded to an amino group
(–NH2), a carboxyl group (–COOH), and a hydrogen atom. Every amino acid also has another
variable atom or group of atoms bonded to the central carbon atom known as the R group.
The R group is the only difference in structure between the 20 amino acids; otherwise, the
amino acids are identical (Figure 7.10).
The chemical nature of the R group determines the chemical nature of the amino acid
within its protein (that is, whether it is acidic, basic, polar, or nonpolar).
The sequence and number of amino acids ultimately determine a protein’s shape, size, and
function. Each amino acid is attached to another amino acid by a covalent bond, known as a
peptide bond, which is formed by a dehydration reaction. The carboxyl group of one amino
acid and the amino group of a second amino acid combine, releasing a water molecule. The
resulting bond is the peptide bond.
The products formed by such a linkage are called polypeptides. While the terms polypeptide
and protein are sometimes used interchangeably, a polypeptide is technically a polymer of
amino acids, whereas the term protein is used for a polypeptide or polypeptides that have
combined together, have a distinct shape, and have a unique function.
Figure 7.10: Amino acids are made up of a central carbon bonded to an amino group (–NH2), a carboxyl group
(–COOH), and a hydrogen atom. The central carbon’s fourth bond varies among the different amino acids, as
seen in these examples of alanine, valine, lysine, and aspartic acid.
EVOLUTION IN ACTION
The Evolutionary Significance of Cytochrome c
Cytochrome c is an important component of the molecular machinery that harvests energy
from glucose. Because this protein’s role in producing cellular energy is crucial, it has
changed very little over millions of years. Protein sequencing has shown that there is a
considerable amount of sequence similarity among cytochrome c molecules of different
species; evolutionary relationships can be assessed by measuring the similarities or
differences among various species’ protein sequences.
For example, scientists have determined that human cytochrome c contains 104 amino
acids. For each cytochrome c molecule that has been sequenced to date from different
organisms, 37 of these amino acids appear in the same position in each cytochrome c. This
indicates that all of these organisms are descended from a common ancestor. On comparing
the human and chimpanzee protein sequences, no sequence difference was found. When
human and rhesus monkey sequences were compared, a single difference was found in one
amino acid. In contrast, human-to-yeast comparisons show a difference in 44 amino acids,
suggesting that humans and chimpanzees have a more recent common ancestor than
humans and the rhesus monkey, or humans and yeast.

Protein Structure
As discussed earlier, the shape of a protein is critical to its function. To understand how the
protein gets its final shape or conformation, we need to understand the four levels of
protein structure: primary, secondary, tertiary, and quaternary (Figure 7.11).
The unique sequence and number of amino acids in a polypeptide chain is its primary
structure. The unique sequence for every protein is ultimately determined by the gene that
encodes the protein. Any change in the gene sequence may lead to a different amino acid
being added to the polypeptide chain, causing a change in protein structure and function. In
sickle cell anemia, the haemoglobin β chain has a single amino acid substitution, causing a
change in both the structure and function of the protein. What is most remarkable to
consider is that a haemoglobin molecule is made up of two alpha chains and two beta
chains that each consist of about 150 amino acids. The molecule, therefore, has about 600
amino acids. The structural difference between a normal haemoglobin molecule and a sickle
cell molecule—that dramatically decreases life expectancy in the affected individuals—is a
single amino acid of the 600.
Because of this change of one amino acid in the chain, the normally biconcave, or disc-
shaped, red blood cells assume a crescent or “sickle” shape, which clogs arteries. This can
lead to a myriad of serious health problems, such as breathlessness, dizziness, headaches,
and abdominal pain for those who have this disease.
Folding patterns resulting from interactions between the non-R group portions of amino
acids give rise to the secondary structure of the protein. The most common are the alpha
(α)-helix and beta (β)-pleated sheet structures. Both structures are held in shape by
hydrogen bonds. In the alpha helix, the bonds form between every fourth amino acid and
cause a twist in the amino acid chain.
In the β-pleated sheet, the “pleats” are formed by hydrogen bonding between atoms on the
backbone of the polypeptide chain. The R groups are attached to the carbons, and extend
above and below the folds of the pleat. The pleated segments align parallel to each other,
and hydrogen bonds form between the same pairs of atoms on each of the aligned amino
acids. The α-helix and β-pleated sheet structures are found in many globular and fibrous
proteins.
The unique three-dimensional structure of a polypeptide is known as its tertiary structure.
This structure is caused by chemical interactions between various amino acids and regions
of the polypeptide. Primarily, the interactions among R groups create the complex three-
dimensional tertiary structure of a protein. There may be ionic bonds formed between R
groups on different amino acids, or hydrogen bonding beyond that involved in the
secondary structure. When protein folding takes place, the hydrophobic R groups of
nonpolar amino acids lay in the interior of the protein, whereas the hydrophilic R groups lay
on the outside. The former types of interactions are also known as hydrophobic
interactions.
In nature, some proteins are formed from several polypeptides, also known as subunits, and
the interaction of these subunits forms the quaternary structure. Weak interactions
between the subunits help to stabilise the overall structure. For example, hemoglobin is a
combination of four polypeptide subunits.
Each protein has its own unique sequence and shape held together by chemical
interactions. If the protein is subject to changes in temperature, pH, or exposure to
chemicals, the protein structure may change, losing its shape in what is known as
denaturation as discussed earlier. Denaturation is often reversible because the primary
structure is preserved if the denaturing agent is removed, allowing the protein to resume its
function. Sometimes denaturation is irreversible, leading to a loss of function. One example
of protein denaturation can be seen when an egg is fried or boiled. The albumin protein in
the liquid egg white is denatured when placed in a hot pan, changing from a clear substance
to an opaque white substance. Not all proteins are denatured at high temperatures; for
instance, bacteria that survive in hot springs have proteins that are adapted to function at
those temperatures.
Figure 7.11 The four levels of protein structure can be observed in these illustrations. (credit: modification of
work by National Human Genome Research Institute)
CONCEPT IN ACTION

For an additional perspective on proteins, explore “Biomolecules: The Proteins” through


this interactive animation (http://openstaxcollege.org/l/proteins) .
KEY TERMS
trans-fat a form of unsaturated fat with the hydrogen atoms neighbouring the double bond
across from each other rather than on the same side of the double bond
amino acid a monomer of a protein
atomic number the number of protons in an atom
carbohydrate a biological macromolecule in which the ratio of carbon to hydrogen to
oxygen is 1:2:1; carbohydrates serve as energy sources and structural support in cells
cellulose a polysaccharide that makes up the cell walls of plants and provides structural
support to the cell
chemical bond an interaction between two or more of the same or different elements that
results in the formation of molecules
chitin a type of carbohydrate that forms the outer skeleton of arthropods, such as insects
and crustaceans, and the cell walls of fungi
covalent bond a type of strong bond between two or more of the same or different
elements; forms when electrons are shared between elements
denaturation the loss of shape in a protein as a result of changes in temperature, pH, or
exposure to chemicals
disaccharide two sugar monomers that are linked together by a peptide bond
element one of 118 unique substances that cannot be broken down into smaller substances
and retain the characteristic of that substance; each element has a specified number of
protons and unique properties
enzyme a catalyst in a biochemical reaction that is usually a complex or conjugated protein
fat a lipid molecule composed of three fatty acids and a glycerol (triglyceride) that typically
exists in a solid form at room temperature
glycogen a storage carbohydrate in animals
hormone a chemical signalling molecule, usually a protein or steroid, secreted by an
endocrine gland or group of endocrine cells; acts to control or regulate specific physiological
processes
hydrophilic describes a substance that dissolves in water; water-loving
hydrophobic describes a substance that does not dissolve in water; water-fearing
lipids a class of macromolecules that are nonpolar and insoluble in water
macromolecule a large molecule, often formed by polymerisation of smaller monomers
monosaccharide a single unit or monomer of carbohydrates
oil an unsaturated fat that is a liquid at room temperature
phospholipid a major constituent of the membranes of cells; composed of two fatty acids
and a phosphate group attached to the glycerol backbone
polypeptide a long chain of amino acids linked by peptide bonds
polysaccharide a long chain of monosaccharides; may be branched or unbranched
protein a biological macromolecule composed of one or more chains of amino acids
saturated fatty acid a long-chain hydrocarbon with single covalent bonds in the carbon
chain; the number of hydrogen atoms attached to the carbon skeleton is maximised
starch a storage carbohydrate in plants
steroid a type of lipid composed of four fused hydrocarbon rings
triglyceride a fat molecule; consists of three fatty acids linked to a glycerol molecule
unsaturated fatty acid a long-chain hydrocarbon that has one or more than one double
bonds in the hydrocarbon chain

CHAPTER SUMMARY
Biological Molecules
Living things are carbon-based because carbon plays such a prominent role in the chemistry
of living things. The four covalent bonding positions of the carbon atom can give rise to a
wide diversity of compounds with many functions, accounting for the importance of carbon
in living things. Carbohydrates are a group of macromolecules that are a vital energy source
for the cell, provide structural support to many organisms, and can be found on the surface
of the cell as receptors or for cell recognition. Carbohydrates are classified as
monosaccharides, disaccharides, and polysaccharides, depending on the number of
monomers in the molecule.
Lipids are a class of macromolecules that are nonpolar and hydrophobic in nature. Major
types include fats and oils, waxes, phospholipids, and steroids. Fats and oils are a stored
form of energy and can include triglycerides. Fats and oils are usually made up of fatty acids
and glycerol.
Proteins are a class of macromolecules that can perform a diverse range of functions for the
cell. They help in metabolism by providing structural support and by acting as enzymes,
carriers or as hormones. The building blocks of proteins are amino acids. Proteins are
organized at four levels: primary, secondary, tertiary, and quaternary. Protein shape and
function are intricately linked; any change in shape caused by changes in temperature, pH,
or chemical exposure may lead to protein denaturation and a loss of function.
Footnotes
1. Humphrey, W., Dalke, A. and Schulten, K., "VMD—Visual Molecular Dynamics", J. Molec.
Graphics, 1996, vol. 14, pp. 33-38. http://www.ks.uiuc.edu/Research/vmd/

REVIEW QUESTIONS
Question 7.1
An example of a monosaccharide is ________ ________
A fructose
B glucose
C galactose
D all of the above

Question 7.2
Cellulose and starch are examples of ________ ________
A monosaccharides
B disaccharides
C lipids
D polysaccharides

Question 7.3
In the body, carbohydrates are stored as ________ ________
A glycogen
B starch
C cholesterol
D polypeptides
Question 7.4
Phospholipids are important components of ________ ________
A the plasma membrane of cells
B the ring structure of steroids
C the waxy covering on leaves
D the double bond in hydrocarbon chains

Question 7.5
The monomers that make up proteins are called ________ ________
A nucleotides
B disaccharides
C amino acids
D chaperones

Question 7.6
A triglyceride consists of ________
A glycerol plus three fatty acids
B a sugar-phosphate backbone to which two amino groups are attached
C two to several hexoses
D amino acids that have been thoroughly saturated with hydrogen

Question 7.7
A molecule has an amine group and an organic acid group. It does not, however, have any
peptide bonds. It is ________
A a monosaccharide
B an amino acid
C a protein
D a fat
Question 7.8
Which of the following is not a function of proteins?
A structural components of cells
B glucose storage
C cell movement
D all are functions of proteins

Question 7.9
Which of the following is a function of lipids?
A structural component of cell membranes
B long term energy storage
C production of hormones
D all are functions of lipids

Question 7.10
If a polypeptide contains 10 peptide bonds how many amino acids does it contain?
A9
B 10
C 11
D 12

Question 7.11
Explain at least three functions that lipids serve in plants and/or animals.

Question 7.12
Explain what happens if even one amino acid is substituted for another in a polypeptide
chain. Provide a specific example.
Question 7.13
Differentiate clearly between primary, secondary and tertiary protein structure.

Question 7.14
Look at the structures below and identify which class of biomolecule they belong to.
a)

__________________________________________________
b)

____________________________________________________
c)

_____________________________________________________

Question 7.15
Why is cholesterol necessary in the body?
ANSWERS
Question 7.1
D

Question 7.2
D

Question 7.3
A

Question 7.4
A

Question 7.5
C

Question 7.6
A

Question 7.7
B

Question 7.8
B

Question 7.9
D

Question 7.10
C

Question 7.11
Fat serves as a valuable way for animals to store energy. It can also provide insulation.
Phospholipids and steroids are important components of cell membranes

Question 7.12
A change in gene sequence can lead to a different amino acid being added to a polypeptide
chain instead of the normal one. This causes a change in protein structure and function. For
example, in sickle cell anemia, the hemoglobin β chain has a single amino acid substitution.
Because of this change, the disc-shaped red blood cells assume a crescent shape, which can
result in serious health problems.
Question 7.13
Primary structure is the chain of amino acids. Secondary structure is the forming of the
alpha helix and the beta-pleated sheet because of hydrogen bonding of the peptide
backbone. Tertiary structure is the folding of the polypeptide into its 3D shape due to
interactions between the side chains of the amino acids. Quaternary structure is the
association of more than one polypeptide to form the complete protein.

Question 7.14
a) Lipid (triglyceride)
b) Protein (amino acid)
c) Carbohydrate (monosaccharide)

Question 7.15
Cholesterol is a necessary molecule in the construction of the cell membrane. It is also
needed to make steroid hormones such as testosterone and estradiol, vitamins E and K and
bile salts needed for digestion of fats.
Chapter 8 | The Essential
Reactions of Life

Figure 8.0 Metabolism. Metabolism is the sum of all energy-requiring and energy-consuming
processes of the body. Many factors contribute to overall metabolism, including lean muscle mass,
the amount and quality of food consumed, and the physical demands placed on the human body.
(credit: "tableatny"/flickr.com)

After studying this chapter you will be able to:


• By the end of this section, you will be able to:
• Explain what a chemical reaction is
• Explain the role of enzymes in a cell
• Explain the role of ATP in the cell
• Describe how ATP is generated from glucose metabolism in glycolysis
• Explain how the citric acid cycle and oxidative phosphorylation produce further ATP
from the products of glycolysis and how carbon dioxide and water are formed in the
process
• Explain how fats are metabolised to produce ATP in the mitochondria
• Explain which pathways of ATP production are available in the presence and absence
of oxygen

8.1 Chemical Reactions


One characteristic of a living organism is metabolism, which is the sum total of all of the
chemical reactions that go on to maintain that organism’s health and life. The bonding
processes you have learned thus far are anabolic chemical reactions; that is, they form
larger molecules from smaller molecules or atoms. But recall that metabolism can proceed
in another direction: in catabolic chemical reactions, bonds between components of larger
molecules break, releasing smaller molecules or atoms. Both types of reaction involve
exchanges not only of matter, but of energy.

8.1.1 The Role of Energy in Chemical


Reactions
Chemical reactions require a sufficient amount of energy to cause the matter to collide with
enough precision and force that old chemical bonds can be broken and new ones formed. In
general, kinetic energy is the form of energy powering any type of matter in motion.
Imagine you are building a brick wall. The energy it takes to lift and place one brick
atop another is kinetic energy—the energy matter possesses because of its motion. Once
the wall is in place, it stores potential energy. Potential energy is the energy of position, or
the energy matter possesses because of the positioning or structure of its components. If
the brick wall collapses, the stored potential energy is released as kinetic energy as the
bricks fall.
In the human body, potential energy is stored in the bonds between atoms and
molecules. Chemical energy is the form of potential energy in which energy is stored in
chemical bonds. When those bonds are formed, chemical energy is invested, and when they
break, chemical energy is released. Notice that chemical energy, like all energy, is neither
created nor destroyed; rather, it is converted from one form to another. When you eat an
energy bar before heading out the door for a hike, the honey, nuts, and other foods the bar
contains are broken down and rearranged by your body into molecules that your muscle
cells convert to kinetic energy.
8.1.2 Forms of Energy Important in Human
Functioning
You have already learned that chemical energy is absorbed, stored, and released by
chemical bonds. In addition to chemical energy, mechanical, radiant, and electrical energy
are important in human functioning. In the following sections we will discuss how chemical
energy is transferred from nutrients to the main energy source in the cell: adenine
triphosphate (ATP)/

8.1.3 Characteristics of Chemical Reactions


All chemical reactions begin with a reactant, the general term for the one or more
substances that enter into the reaction. Sodium and chloride ions, for example, are the
reactants in the production of table salt. The one or more substances produced by a
chemical reaction are called the product.
In chemical reactions, the components of the reactants—the elements involved and the
number of atoms of each—are all present in the product(s). Similarly, there is nothing
present in the products that are not present in the reactants. This is because chemical
reactions are governed by the law of conservation of mass, which states that matter cannot
be created or destroyed in a chemical reaction.
Just as you can express mathematical calculations in equations such as 2 + 7 = 9, you can use
chemical equations to show how reactants become products. As in math, chemical
equations proceed from left to right, but instead of an equal sign, they employ an arrow or
arrows indicating the direction in which the chemical reaction proceeds. For example, the
chemical reaction in which one atom of nitrogen and three atoms of hydrogen produce
ammonia would be written as N + 3H → NH3 . Correspondingly, the breakdown of ammonia
into its components would be written as NH3 → N + 3H.
Notice that, in the first example, a nitrogen (N) atom and three hydrogen (H) atoms bond to
form a compound. This anabolic reaction requires energy, which is then stored within the
compound’s bonds. Such reactions are referred to as synthesis reactions. A synthesis
reaction is a chemical reaction that results in the synthesis (joining) of components that
were formerly separate (Figure 8.1a). Again, nitrogen and hydrogen are reactants in a
synthesis reaction that yields ammonia as the product. The general equation for a synthesis
reaction is A + B → AB.
Figure 8.1 The Three Fundamental Chemical Reactions. The atoms and molecules involved in the
three fundamental chemical reactions can be imagined as words.

In the second example, ammonia is catabolised into its smaller components, and the
potential energy that had been stored in its bonds is released. Such reactions are referred to
as decomposition reactions. A decomposition reaction is a chemical reaction that breaks
down or “de-composes” something larger into its constituent parts (see Figure 8.1b). The
general equation for a decomposition reaction is: AB → A + B .
An exchange reaction is a chemical reaction in which both synthesis and decomposition
occur, chemical bonds are both formed and broken, and chemical energy is absorbed,
stored, and released (see Figure 8.1c). The simplest form of an exchange reaction might be:
A + BC → AB + C. Notice that, to produce these products, B and C had to break apart in
a decomposition reaction, whereas A and B had to bond in a synthesis reaction. A more
complex exchange reaction might be: AB + CD → AC + BD. Another example might be: AB +
CD → AD + BC.
In theory, any chemical reaction can proceed in either direction under the right conditions.
Reactants may synthesise into a product that is later decomposed. Reversibility is also a
quality of exchange reactions. For instance, A + BC → AB + C could then reverse to AB + C →
A + BC. This reversibility of a chemical reaction is indicated with a double arrow: A + BC ⇄
AB + C. Still, in the human body, many chemical reactions do proceed in a predictable
direction, either one way or the other. You can think of this more predictable path as the
path of least resistance because, typically, the alternate direction requires more energy.

8.1.4 Factors Influencing the Rate of


Chemical Reactions
If you pour vinegar into baking soda, the reaction is instantaneous; the concoction will
bubble and fizz. But many chemical reactions take time. A variety of factors influence the
rate of chemical reactions. This section, however, will consider only the most important in
human functioning.

Properties of the Reactants


If chemical reactions are to occur quickly, the atoms in the reactants have to have easy
access to one another. Thus, the greater the surface area of the reactants, the more readily
they will interact. When you pop a cube of cheese into your mouth, you chew it before you
swallow it. Among other things, chewing increases the surface area of the food so that
digestive chemicals can more easily get at it. As a general rule, gases tend to react faster
than liquids or solids, again because it takes energy to separate particles of a substance, and
gases by definition already have space between their particles. Similarly, the larger the
molecule, the greater the number of total bonds, so reactions involving smaller molecules,
with fewer total bonds, would be expected to proceed faster.
In addition, recall that some elements are more reactive than others. Reactions that involve
highly reactive elements like hydrogen proceed more quickly than reactions that involve less
reactive elements. Reactions involving stable elements like helium are not likely to happen
at all.

Temperature
Nearly all chemical reactions occur at a faster rate at higher temperatures. Recall that
kinetic energy is the energy of matter in motion. The kinetic energy of subatomic particles
increases in response to increases in thermal energy. The higher the temperature, the faster
the particles move, and the more likely they are to come in contact and react.

Concentration and Pressure


If just a few people are dancing at a club, they are unlikely to step on each other’s toes. But
as more and more people get up to dance—especially if the music is fast—collisions are
likely to occur. It is the same with chemical reactions: the more particles present within a
given space, the more likely those particles are to bump into one another. This means
that chemists can speed up chemical reactions not only by increasing the concentration of
particles—the number of particles in the space—but also by decreasing the volume of the
space, which would correspondingly increase the pressure. If there were 100 dancers in that
club, and the manager abruptly moved the party to a room half the size, the concentration
of the dancers would double in the new space, and the likelihood of collisions would
increase accordingly.
Enzymes and Other Catalysts
For two chemicals in nature to react with each other they first have to come into contact,
and this occurs through random collisions. Because heat helps increase the kinetic energy of
atoms, ions, and molecules, it promotes their collision. But in the body, extremely high
heat—such as a very high fever—can damage body cells and be life-threatening. On the
other hand, normal body temperature is not high enough to promote the chemical reactions
that sustain life. That is where catalysts come in.
In chemistry, a catalyst is a substance that increases the rate of a chemical reaction without
itself undergoing any change. You can think of a catalyst as a chemical change agent. They
help increase the rate and force at which atoms, ions, and molecules collide, thereby
increasing the probability that their valence shell electrons will interact.
The most important catalysts in the human body are enzymes. An enzyme is a catalyst
composed of protein or ribonucleic acid (RNA), both of which will be discussed later in this
chapter. Like all catalysts, enzymes work by lowering the level of energy that needs to be
invested in a chemical reaction. A chemical reaction’s activation energy is the “threshold”
level of energy needed to break the bonds in the reactants. Once those bonds are broken,
new arrangements can form. Without an enzyme to act as a catalyst, a much larger
investment of energy is needed to ignite a chemical reaction (Figure 8.2).
Enzymes are critical to the body’s healthy functioning. They assist, for example, with the
breakdown of food and its conversion to energy. In fact, most of the chemical reactions in
the body are facilitated by enzymes.

Figure 8.2 Enzymes Enzymes decrease the activation energy required for a given chemical reaction to occur.
(a) Without an enzyme, the energy input needed for a reaction to begin is high. (b) With the help of an
enzyme, less energy is needed for a reaction to begin.
Extra info from A&P in context - integrate with above
Enzymes and Catalysis
Enzymes are essential catalysts that speed up the rate of the biological reactions happening
non-stop in your living cells. Without them, studies have estimated that certain biological
processes, such as the hydrolysis of phosphate dianions during the phosphatidylinositol
signalling pathway discussed further in the chapter on Endocrinology, would take a trillion
years to occur, whereas a phosphatase enzyme can accomplish this same task in one
hundredth of a second! I think we all are a little too impatient to wait for reactions to
happen on their own, so let’s discuss how enzymes perform this important duty within our
cells.
Enzymes are typically protein molecules that have the ability
to hydrolyse or condense biomolecules. Each enzyme is specialised to its own particular
task; for instance, sucrase is only capable of breaking down sucrose sugars and DNA ligase is
responsible for bonding DNA nucleotides together. Enzymes interact with their substrate at
an active site due to a very specific lock-and-key fit (see Figure 8.3). The enzyme undergoes
a slight conformational change, called induced fit, in order to interact strongly with the
substrate and carry out the intended reaction.

Figure 8.3. Enzymes bind to their specific substrate(s) at the active site. To associate closely with the
substrate(s), the enzyme undergoes a slight shape change called “induced fit” which will facilitate the proper
biological reaction and generation of the product(s).

Environmental conditions such as pH and temperature can affect the rate at which enzymes
function due to modifying enzyme-substrate interaction through shape change or molecular
movement. Importantly, enzymes are not consumed in the reaction and can be reused
many times over. Watch this video to see how enzymes exert their effect on substrates.
https://www.youtube.com/watch?v=pVoytz_3H_s

8.2 | How cells obtain energy


Virtually every task performed by living organisms requires energy. Energy is needed to
perform heavy labor and exercise, but humans also use energy while thinking, and even
during sleep. In fact, the living cells of every organism constantly use energy. Nutrients and
other molecules are imported into the cell, metabolised (broken down) and possibly
synthesised into new molecules, modified if needed, transported around the cell, and
possibly distributed to the entire organism. For example, the large proteins that make up
muscles are built from smaller molecules imported from dietary amino acids. Complex
carbohydrates are broken down into simple sugars that the cell uses for energy. Just as
energy is required to both build and demolish a building, energy is required for the synthesis
and breakdown of molecules as well as the transport of molecules into and out of cells. In
addition, processes such as ingesting and breaking down pathogenic bacteria and viruses,
exporting wastes and toxins, and movement of the cell require energy. From where, and in
what form, does this energy come? How do living cells obtain energy, and how do they use
it? At the end of this chapter you will be able to name the metabolic pathways cells use to
harness chemical energy from the breakdown of carbohydrates and fat and describe how
these pathways allow cells to produce ATP to power cellular processes.
8.2.1 Metabolic Pathways
Consider the metabolism of sugar. This is a classic example of one of the many cellular
processes that use and produce energy. Living things consume sugars as a major energy
source, because sugar molecules have a great deal of energy stored within their
bonds. Just as the dollar is used as currency to buy goods, cells use molecules of Adenosine
Triphosphate (ATP) as energy currency to perform immediate work. In contrast, energy-
storage molecules such as glucose are consumed only to be broken down to use their
energy. The process used for breaking down carbohydrate to liberate the energy to
produce ATP can be summarised as:

The process for breaking down sugar molecules to produce ATP actually involves three
metabolic A metabolic pathway is a series of chemical reactions that takes a starting
molecule and modifies it, step-by-step, through a series of metabolic intermediates,
eventually yielding a final product.
It is important to know that the chemical reactions of metabolic pathways do not take place
on their own. Each reaction step is facilitated, or catalysed, by a protein called an enzyme.
Enzymes are important for catalysing all types of biological reactions—those that require
energy as well as those that release energy.
There are three key catabolic pathways involved in converting the energy in glucose to
ATP. These pathways are glycolysis, the citric acid (TCA) cycle and oxidative
phosphorylation.

8.2.2 What is Adenosine Triphosphate (ATP)?


Even exergonic, energy-releasing reactions require a small amount of activation energy to
proceed. However, consider endergonic reactions, which require much more energy input
because their products have more free energy than their reactants. Within the cell, where
does energy to power such reactions come from? The answer lies with an energy-supplying
molecule called adenosine triphosphate, or ATP. ATP is a small, relatively simple molecule,
but within its bonds contains the potential for a quick burst of energy that can be harnessed
to perform cellular work. This molecule can be thought of as the primary energy currency of
cells in the same way that money is the currency that people exchange for things they need.
ATP is used to power the majority of energy-requiring cellular reactions.

ATP in Living Systems


A living cell cannot store significant amounts of free energy. Excess free energy would result
in an increase of heat in the cell, which would denature enzymes and other proteins, and
thus destroy the cell. Rather, a cell must be able to store energy safely and release it for use
only as needed. Living cells accomplish this using ATP, which can be used to fill any energy
need of the cell. How? It functions as a rechargeable battery.
When ATP is broken down, usually by the removal of its terminal phosphate group, energy is
released. This energy is used to do work by the cell, usually by the binding of the released
phosphate to another molecule, thus activating it. For example, in the mechanical work of
muscle contraction, ATP supplies energy to move the contractile muscle proteins.

ATP Structure and Function


At the heart of ATP is a molecule of adenosine monophosphate (AMP), which is composed
of an adenine molecule bonded to both a ribose molecule and a single phosphate
group (Figure 8.4). Ribose is a five-carbon sugar found in RNA and AMP is one of the
nucleotides in RNA. The addition of a second phosphate group to this core molecule results
in adenosine diphosphate (ADP); the addition of a third phosphate group forms adenosine
triphosphate (ATP).

Figure 8.4 The structure of ATP shows the basic components of a two-ring adenine, five-carbon ribose, and
three phosphate groups.

The addition of a phosphate group to a molecule requires a high amount of energy and
results in a high-energy bond. Phosphate groups are negatively charged and thus repel one
another when they are arranged in series, as they are in ADP and ATP. This repulsion makes
the ADP and ATP molecules inherently unstable. The release of one or two phosphate
groups from ATP, a process called hydrolysis, releases energy.

8.2.3 Energy metabolism part 1: glycolysis


You have read that nearly all of the energy used by living things comes to them in the bonds
of the sugar, glucose. Glycolysis is the first step in the breakdown of glucose to extract
energy for cell metabolism. Many living organisms carry out glycolysis as part of their
metabolism. Glycolysis takes place in the cytoplasm of most prokaryotic and all eukaryotic
cells.
Glycolysis begins with the six-carbon, ring-shaped structure of a single glucose molecule and
ends with two molecules of a three-carbon sugar called pyruvate. Glycolysis consists of two
distinct phases. In the first part of the glycolysis pathway, energy is used to make
adjustments so that the six-carbon sugar molecule can be split evenly into two three-carbon
pyruvate molecules. In the second part of glycolysis, ATP and nicotinamide-adenine
dinucleotide (NADH) are produced (Figure 8.5).
If the cell cannot catabolise the pyruvate molecules further, it will harvest only two ATP
molecules from one molecule of glucose. For example, mature mammalian red blood cells
are only capable of glycolysis, which is their sole source of ATP. If glycolysis is interrupted,
these cells would eventually die.

Figure 8.5 In glycolysis, a glucose molecule is converted into two pyruvate molecules.
8.2.4 Energy metabolism part 2: the Citric
Acid Cycle
In eukaryotic cells, the pyruvate molecules produced at the end of glycolysis are transported
into mitochondria, which are sites of cellular respiration. If oxygen is available, aerobic
respiration will go forward. In mitochondria, pyruvate will be transformed into a two-carbon
acetyl group (by removing a molecule of carbon dioxide) that will be picked up by a carrier
compound called coenzyme A (CoA), which is made from vitamin B5. The resulting
compound is called acetyl CoA. (Figure 8.6). Acetyl CoA can be used in a variety of ways by
the cell, but its major function is to deliver the acetyl group derived from pyruvate to the
next pathway in glucose catabolism.

Figure 8.6 Pyruvate is converted into acetyl-CoA before entering the citric acid cycle.

Like the conversion of pyruvate to acetyl CoA, the citric acid cycle in eukaryotic cells takes
place in the matrix of the mitochondria. Unlike glycolysis, the citric acid cycle is a closed
loop: The last part of the pathway regenerates the compound used in the first step. The
eight steps of the cycle are a series of chemical reactions that produces two carbon dioxide
molecules, one ATP molecule (or an equivalent), and reduced forms (NADH and FADH2) of
NAD+ and FAD+, important coenzymes in the cell. Part of this is considered an aerobic
pathway (oxygen-requiring) because the NADH and FADH2 produced must transfer their
electrons to the next pathway in the system, which will use oxygen. If oxygen is not present,
this transfer does not occur.
Two carbon atoms come into the citric acid cycle from each acetyl group. Two carbon
dioxide molecules are released on each turn of the cycle; however, these do not contain the
same carbon atoms contributed by the acetyl group on that turn of the pathway. The two
acetyl-carbon atoms will eventually be released on later turns of the cycle; in this way, all six
carbon atoms from the original glucose molecule will be eventually released as carbon
dioxide. It takes two turns of the cycle to process the equivalent of one glucose molecule.
Each turn of the cycle forms three high-energy NADH molecules and one high-energy
FADH2 molecule. These high-energy carriers will connect with the last portion of aerobic
respiration to produce ATP molecules. One ATP (or an equivalent) is also made in each cycle.
Several of the intermediate compounds in the citric acid cycle can be used in synthesising
non-essential amino acids; therefore, the cycle is both anabolic and catabolic.

8.2.5 Energy metabolism part 3: oxidative


phosphorylation
You have just read about two pathways in glucose catabolism—glycolysis and the citric acid
cycle—that generate ATP. Most of the ATP generated during the aerobic catabolism of
glucose, however, is not generated directly from these pathways. Rather, it derives from a
process that begins with passing electrons through a series of chemical reactions to a final
electron acceptor, oxygen. These reactions take place in specialised protein complexes
located in the inner membrane of the mitochondria of eukaryotic organisms and on the
inner part of the cell membrane of prokaryotic organisms. The energy of the electrons is
harvested and used to generate a electrochemical gradient across the inner mitochondrial
membrane. The potential energy of this gradient is used to generate ATP. The entirety of
this process is called oxidative phosphorylation.
The electron transport chain (Figure 8.7a) is the last component of aerobic respiration and is
the only part of metabolism that uses atmospheric oxygen. In animals, oxygen enters the
body through the respiratory system for this purpose. Electron transport is a series of
chemical reactions that resembles a bucket brigade in that electrons are passed rapidly from
one component to the next, to the endpoint of the chain where oxygen is the final electron
acceptor and water is produced. There are four complexes composed of proteins, labeled I
through IV in Figure 8.7c, and the aggregation of these four complexes, together with
associated mobile, accessory electron carriers, is called the electron transport
chain. The electron transport chain is present in multiple copies in the inner mitochondrial
membrane of eukaryotes and in the plasma membrane of prokaryotes. In each transfer of
an electron through the electron transport chain, the electron loses energy, but with some
transfers, the energy is stored as potential energy by using it to pump hydrogen ions across
the inner mitochondrial membrane into the intermembrane space, creating an
electrochemical gradient.
Cyanide inhibits cytochrome c oxidase, a component of the electron transport chain. If
cyanide poisoning occurs, would you expect the pH of the intermembrane space to increase
or decrease? What affect would cyanide have on ATP synthesis?
Electrons from NADH and FADH2 are passed to protein complexes in the electron transport
chain. As they are passed from one complex to another (there are a total of four), the
electrons lose energy, and some of that energy is used to pump hydrogen ions from the
mitochondrial matrix into the intermembrane space. In the fourth protein complex, the
electrons are accepted by oxygen, the terminal acceptor. The oxygen with its extra electrons
then combines with two hydrogen ions, further enhancing the electrochemical gradient, to
form water. If there were no oxygen present in the mitochondrion, the electrons could not
be removed from the system, and the entire electron transport chain would back up and
stop. The mitochondria would be unable to generate new ATP in this way, and the cell
would ultimately die from lack of energy. This is the reason we must breathe to draw in new
oxygen.

Figure 8.7 (a) The electron transport chain is a set of molecules that supports a series of oxidation-reduction
reactions. (b) ATP synthase is a complex, molecular machine that uses an H+ gradient to regenerate ATP from
ADP. (c) Chemiosmosis relies on the potential energy provided by the H+ gradient across the membrane.
In the electron transport chain, the free energy from the series of reactions just described is
used to pump hydrogen ions across the membrane. The uneven distribution of H+ ions
across the membrane establishes an electrochemical gradient, owing to the H+ ions’ positive
charge and their higher concentration on one side of the membrane.
Hydrogen ions diffuse through the inner membrane through an integral membrane protein
called ATP synthase (Figure 8.7b). This complex protein acts as a tiny generator, turned by
the force of the hydrogen ions diffusing through it, down their electrochemical gradient
from the intermembrane space, where there are many mutually repelling hydrogen ions to
the matrix, where there are few. The turning of the parts of this molecular machine
regenerate ATP from ADP. This flow of hydrogen ions across the membrane through ATP
synthase is called chemiosmosis.
Chemiosmosis (Figure 8.7c) is used to generate 90 percent of the ATP made during aerobic
glucose catabolism. The result of the reactions is the production of ATP from the energy of
the electrons removed from hydrogen atoms. These atoms were originally part of a glucose
molecule. At the end of the electron transport system, the electrons are used to reduce an
oxygen molecule to oxygen ions. The extra electrons on the oxygen ions attract hydrogen
ions (protons) from the surrounding medium, and water is formed. The electron transport
chain and the production of ATP through chemiosmosis are collectively called oxidative
phosphorylation.

ATP Yield
The number of ATP molecules generated from the catabolism of glucose varies. For
example, the number of hydrogen ions that the electron transport chain complexes can
pump through the membrane varies between species. Another source of variance stems
from the shuttle of electrons across the mitochondrial membrane. The NADH generated
from glycolysis cannot easily enter mitochondria. Thus, electrons are picked up on the inside
of the mitochondria by either NAD+ or FAD+. Fewer ATP molecules are generated when
FAD+ acts as a carrier. NAD+ is used as the electron transporter in the liver and FAD+in the
brain, so ATP yield depends on the tissue being considered.
Another factor that affects the yield of ATP molecules generated from glucose is that
intermediate compounds in these pathways are used for other purposes. Glucose
catabolism connects with the pathways that build or break down all other biochemical
compounds in cells, and the result is somewhat messier than the ideal situations described
thus far. For example, sugars other than glucose are fed into the glycolytic pathway for
energy extraction. Other molecules that would otherwise be used to harvest energy in
glycolysis or the citric acid cycle may be removed to form nucleic acids, amino acids, lipids,
or other compounds. Overall, in living systems, these pathways of glucose catabolism
extract about 34 percent of the energy contained in glucose.
CAREERS IN ACTION

Mitochondrial Disease Physician


What happens when the critical reactions of cellular respiration do not proceed correctly?
Mitochondrial diseases are genetic disorders of metabolism. Mitochondrial disorders can
arise from mutations in nuclear or mitochondrial DNA, and they result in the production of
less energy than is normal in body cells. Symptoms of mitochondrial diseases can include
muscle weakness, lack of coordination, stroke-like episodes, and loss of vision and hearing.
Most affected people are diagnosed in childhood, although there are some adult-onset
diseases. Identifying and treating mitochondrial disorders is a specialised medical field. The
educational preparation for this profession requires a college education, followed by
medical school with a specialisation in medical genetics. Medical geneticists can be board
certified by the American Board of Medical Genetics and go on to become associated with
professional organisations devoted to the study of mitochondrial disease, such as the
Mitochondrial Medicine Society and the Society for Inherited Metabolic Disease.

8.2.6 Fermentation
By the end of this section, you will be able to:
In aerobic respiration, the final electron acceptor is an oxygen molecule, O2. If aerobic
respiration occurs, then ATP will be produced using the energy of the high-energy electrons
carried by NADH or FADH2 to the electron transport chain. If aerobic respiration does not
occur, NADH must be reoxidised to NAD+ for reuse as an electron carrier for glycolysis to
continue. How is this done? Some living systems use an organic molecule as the final
electron acceptor. Processes that use an organic molecule to regenerate NAD+ from NADH
are collectively referred to as fermentation. In contrast, some living systems use an
inorganic molecule as a final electron acceptor; both methods are a type of anaerobic
cellular respiration. Anaerobic respiration enables organisms to convert energy for their use
in the absence of oxygen.

Lactic acid (or lactate) Fermentation


The fermentation method used by animals and some bacteria like those in yogurt is lactate
fermentation (Figure 8.8). This occurs routinely in mammalian red blood cells and in skeletal
muscle that has insufficient oxygen supply to allow aerobic respiration to continue (that is,
in muscles used to the point of fatigue). In muscles, lactic acid produced by fermentation
must be removed by the blood circulation and brought to the liver for further metabolism.
The chemical reaction of lactate fermentation is the following:
Pyruvate + NADH ↔lactate + H++ NAD+
The enzyme that catalyses this reaction is lactate dehydrogenase. The reaction can proceed
in either direction, but the left-to-right reaction is inhibited by acidic conditions. This lactate
(and acid) build-up causes muscle stiffness and fatigue. Once the lactic acid has been
removed from the muscle and is circulated to the liver, it can be converted back to pyruvate
and further catabolised for energy.

Figure 8.8 Lactic acid (or lactate) fermentation is common in muscles that have become exhausted by use.

Tremetol, a metabolic poison found in white snake root plant, prevents the metabolism of
lactate. When cows eat this plant, Tremetol is concentrated in the milk. Humans who
consume the milk become ill. Symptoms of this disease, which include vomiting, abdominal
pain, and tremors, become worse after exercise. Why do you think this is the case?
Alcohol Fermentation
Another familiar fermentation process is alcohol fermentation (Figure 8.9a), which produces
ethanol, an alcohol. The alcohol fermentation reaction is the following:

Figure 8.9a The reaction resulting in alcohol fermentation is shown.

In the first reaction, a carboxyl group is removed from pyruvic acid, releasing carbon dioxide
as a gas. The loss of carbon dioxide reduces the molecule by one carbon atom, making

Figure 8.9b Fermentation of grape juice to make wine produces CO2 as a byproduct. Fermentation tanks have
valves so that pressure inside the tanks can be released.

acetaldehyde. The second reaction removes an electron from NADH, forming NAD+ and
producing ethanol from the acetaldehyde, which accepts the electron. The fermentation of
pyruvic acid by yeast produces the ethanol found in alcoholic beverages (Figure 8.9a). If the
carbon dioxide produced by the reaction is not vented from the fermentation chamber, for
example in beer and sparkling wines (Figure 8.9b), it remains dissolved in the medium until
the pressure is released. Ethanol above 12 percent is toxic to yeast, so natural levels of
alcohol in wine occur at a maximum of 12 percent.

8.2.7 Connections to Other Metabolic


Pathways
You have learned about the catabolism of glucose, which provides energy to living cells. But
living things consume more than just glucose for food. How does a turkey sandwich, which
contains protein, provide energy to your cells? This happens because all of the catabolic
pathways for carbohydrates, proteins, and lipids eventually connect into glycolysis and the
citric acid cycle pathways (Figure 8.10). Metabolic pathways should be thought of as
porous—that is, substances enter from other pathways, and other substances leave for
other pathways. These pathways are not closed systems. Many of the products in a
particular pathway are reactants in other pathways.

Connections of Other Sugars to Glucose Metabolism


Glycogen, a polymer of glucose, is a short-term energy storage molecule in animals. When
there is adequate ATP present, excess glucose is converted into glycogen for storage.
Glycogen is made and stored in the liver and muscle. Glycogen will be taken out of storage if
blood sugar levels drop. The presence of glycogen in muscle cells as a source of glucose
allows ATP to be produced for a longer time during exercise.
Sucrose is a disaccharide made from glucose and fructose bonded together. Sucrose is
broken down in the small intestine, and the glucose and fructose are absorbed separately.
Fructose is one of the three dietary monosaccharides, along with glucose and galactose
(which is part of milk sugar, the disaccharide lactose), that are absorbed directly into the
bloodstream during digestion. The catabolism of both fructose and galactose produces the
same number of ATP molecules as glucose.

Connections of Proteins to Glucose Metabolism


Proteins are broken down by a variety of enzymes in cells. Most of the time, amino acids are
recycled into new proteins. If there are excess amino acids, however, or if the body is in a
state of famine, some amino acids will be shunted into glycolysis and the citric acid cycle.
Each amino acid must have its amino group removed prior to entry into these pathways.
The amino group is converted into ammonia. In mammals, the liver synthesises urea from
two ammonia molecules and a carbon dioxide molecule. Thus, urea is the principal waste
product in mammals from the nitrogen originating in amino acids, and it leaves the body in
urine.
Connections of Lipids to Glucose Metabolism
The lipids that are connected to the glucose pathways are cholesterol and triglycerides.
Cholesterol is a lipid that contributes to cell membrane flexibility and is a precursor of
steroid hormones. The synthesis of cholesterol starts with acetyl CoA and proceeds in only
one direction. The process cannot be reversed, and ATP is not produced.
Triglycerides are a form of long-term energy storage in animals. Triglycerides store about
twice as much energy as carbohydrates. Triglycerides are made of glycerol and three fatty
acids. Animals can make most of the fatty acids they need. Triglycerides can be both made
and broken down through parts of the glucose catabolism pathways. Glycerol is metabolised
in glycolysis. Fatty acids are converted to acetyl CoA and enter the citric acid cycle.

Figure 8.10 Glycogen from the liver and muscles, together with fats, can feed into the catabolic pathways for
carbohydrates.
KEY TERMS
ATP synthase a membrane-embedded protein complex that regenerates ATP from ADP
with energy from protons diffusing through it
ATP (also, adenosine triphosphate) the cell’s energy currency
acetyl CoA the combination of an acetyl group derived from pyruvic acid and coenzyme A
which is made from pantothenic acid (a B-group vitamin)
activation energy the amount of initial energy necessary for reactions to occur
active site a specific region on the enzyme where the substrate binds
anabolic describes the pathway that requires a net energy input to synthesise complex
molecules from simpler ones
anaerobic cellular respiration the use of an electron acceptor other than oxygen to
complete metabolism using electron transport-based chemiosmosis
bioenergetics the concept of energy flow through living systems
chemiosmosis the movement of hydrogen ions down their electrochemical gradient across
a membrane through ATP synthase to generate ATP
citric acid cycle a series of enzyme-catalysed chemical reactions of central importance in all
living cells that harvests the energy in carbon-carbon bonds of sugar molecules to generate
ATP; the citric acid cycle is an aerobic metabolic pathway because it requires oxygen in later
reactions to proceed
electron transport chain a series of four large, multi-protein complexes embedded in the
inner mitochondrial membrane that accepts electrons from donor compounds and harvests
energy from a series of chemical reactions to generate a hydrogen ion gradient across the
membrane
enzyme a molecule that catalyses a biochemical reaction
fermentation the steps that follow the partial oxidation of glucose via glycolysis to
regenerate NAD+; occurs in the absence of oxygen and uses an organic compound as the
final electron acceptor
glycolysis the process of breaking glucose into two three-carbon molecules with the
production of ATP and NADH
metabolism all the chemical reactions that take place inside cells, including those that use
energy and those that release energy
oxidative phosphorylation the production of ATP by the transfer of electrons down the
electron transport chain to create a proton gradient that is used by ATP synthase to add
phosphate groups to ADP molecules
potential energy the type of energy that refers to the potential to do work
substrate a molecule on which the enzyme acts
thermodynamics the science of the relationships between heat, energy, and work

Summary
8.1 Energy and Metabolism
Cells perform the functions of life through various chemical reactions. A cell’s metabolism
refers to the combination of chemical reactions that take place within it. Catabolic reactions
break down complex chemicals into simpler ones and are associated with energy release.
Anabolic processes build complex molecules out of simpler ones and require energy.
In studying energy, the term system refers to the matter and environment involved in
energy transfers. Entropy is a measure of the disorder of a system. The physical laws that
describe the transfer of energy are the laws of thermodynamics. The first law states that the
total amount of energy in the universe is constant. The second law of thermodynamics
states that every energy transfer involves some loss of energy in an unusable form, such as
heat energy. Energy comes in different forms: kinetic, potential, and free. The change in free
energy of a reaction can be negative (releases energy, exergonic) or positive (consumes
energy, endergonic). All reactions require an initial input of energy to proceed, called the
activation energy.
Enzymes are chemical catalysts that speed up chemical reactions by lowering their
activation energy. Enzymes have an active site with a unique chemical environment that fits
particular chemical reactants for that enzyme, called substrates. Enzymes and substrates are
thought to bind according to an induced-fit model. Enzyme action is regulated to conserve
resources and respond optimally to the environment.

8.2 Glycolysis
ATP functions as the energy currency for cells. It allows cells to store energy briefly and
transport it within itself to support endergonic chemical reactions. The structure of ATP is
that of an RNA nucleotide with three phosphate groups attached. As ATP is used for energy,
a phosphate group is detached, and ADP is produced. Energy derived from glucose
catabolism is used to recharge ADP into ATP.
Glycolysis is the first pathway used in the breakdown of glucose to extract energy. Because
it is used by nearly all organisms on earth, it must have evolved early in the history of life.
Glycolysis consists of two parts: The first part prepares the six-carbon ring of glucose for
separation into two three-carbon sugars. Energy from ATP is invested into the molecule
during this step to energise the separation. The second half of glycolysis extracts ATP and
high-energy electrons from hydrogen atoms and attaches them to NAD+. Two ATP molecules
are invested in the first half and four ATP molecules are formed during the second half. This
produces a net gain of two ATP molecules per molecule of glucose for the cell.

8.3 Citric Acid Cycle and Oxidative Phosphorylation


The citric acid cycle is a series of chemical reactions that removes high-energy electrons and
uses them in the electron transport chain to generate ATP. One molecule of ATP (or an
equivalent) is produced per each turn of the cycle.
The electron transport chain is the portion of aerobic respiration that uses free oxygen as
the final electron acceptor for electrons removed from the intermediate compounds in
glucose catabolism. The electrons are passed through a series of chemical reactions, with a
small amount of free energy used at three points to transport hydrogen ions across the
membrane. This contributes to the gradient used in chemiosmosis. As the electrons are
passed from NADH or FADH2 down the electron transport chain, they lose energy. The
products of the electron transport chain are water and ATP. A number of intermediate
compounds can be diverted into the anabolism of other biochemical molecules, such as
nucleic acids, non-essential amino acids, sugars, and lipids. These same molecules, except
nucleic acids, can serve as energy sources for the glucose pathway.

8.4 Fermentation
If NADH cannot be metabolised through aerobic respiration, another electron acceptor is
used. Most organisms will use some form of fermentation to accomplish the regeneration of
NAD+, ensuring the continuation of glycolysis. The regeneration of NAD+ in fermentation is
not accompanied by ATP production; therefore, the potential for NADH to produce ATP
using an electron transport chain is not utilised.

8.5 Connections to Other Metabolic Pathways


The breakdown and synthesis of carbohydrates, proteins, and lipids connect with the
pathways of glucose catabolism. The carbohydrates that can also feed into glucose
catabolism include galactose, fructose, and glycogen. These connect with glycolysis. The
amino acids from proteins connect with glucose catabolism through pyruvate, acetyl CoA,
and components of the citric acid cycle. Cholesterol synthesis starts with acetyl CoA, and the
components of triglycerides are picked up by acetyl CoA and enter the citric acid cycle.
REVIEW QUESTIONS
Question 8.1
Which of the following is not true about enzymes?
A They are consumed by the reactions they catalyze.
B They are usually made of amino acids.
C They lower the activation energy of chemical reactions.
D Each one is specific to the particular substrate(s) to which it binds.

Question 8.2
Energy is stored long-term in the bonds of ________ ________ and used short-term to
perform work from a(n) ________ ________ molecule.
A ATP : glucose
B an anabolic molecule : catabolic molecule
C glucose : ATP
D a catabolic molecule : anabolic molecule

Question 8.3
The energy currency used by cells is ________ ________
A ATP
B ADP
C AMP
D adenosine
Question 8.4
The glucose that enters the glycolysis pathway is split into two molecules
of ________ ________
A ATP
B phosphate
C NADH
D pyruvate

Question 8.5
What do the electrons added to NAD+ do?
A They become part of a fermentation pathway.
B They go to another pathway for ATP production.
C They energize the entry of the acetyl group into the citric acid cycle.
D They are converted into NADP.

Question 8.6
Which of the following fermentation methods can occur in animal skeletal muscles?
A lactic acid fermentation
B alcohol fermentation
C mixed acid fermentation
D propionic fermentation

Question 8.7
Beta oxidation is ________ ________ .
A the breakdown of sugars
B the assembly of sugars
C the breakdown of fatty acids
D the removal of amino groups from amino acids
ANSWERS
Question 8.1
A

Question 8.2
C

Question 8.3
A

Question 8.4
D

Question 8.5
B

Question 8.6
A

Question 8.7
C
9| WHY WE BEAT, BREATHE
AND EAT

Figure 9.1 Human Heart This artist’s conception of the human heart suggests a powerful engine—not
inappropriate for a muscular pump that keeps the body continually supplied with blood. (credit: Patrick J.
Lynch)

Chapter Objectives
After studying this chapter, you will be able to:
1. Describe the role of the heart and blood vessels in circulation.
2. Explain what blood pressure is and how it can be varied.
3. Describe the role of the airways, lungs and thoracic cavity in breathing.
4. Explain how pressure gradients, which drive gas movements in to and out of the
bloodstream, are produced.
5. Describe how and where each of the nutrients are broken down to their basic units
6. Explain how the basic units get from the gut in to the bloodstream.

9.1 Heart Anatomy - Why we Beat


In this section, you will explore the remarkable pump that propels the blood into the
vessels. There is no single better word to describe the function of the heart other than
“pump,” since its contraction develops the pressure that ejects blood into the major vessels:
the aorta and pulmonary trunk. From these vessels, the blood is distributed to the
remainder of the body. Although the connotation of the term “pump” suggests a
mechanical device made of steel and plastic, the anatomical structure is a living,
sophisticated muscle. As you read this section try to keep these twin concepts in mind:
pump and muscle.
The vital importance of the heart is obvious. If one assumes an average rate of contraction
of 75 contractions per minute, a human heart would contract approximately 108,000 times
in one day, more than 39 million times in one year, and nearly 3 billion times during a 75-
year lifespan. Each of the major pumping chambers of the heart ejects approximately 70 mL
blood per contraction in a resting adult. This would be equal to 5.25 liters of fluid per
minute and approximately 14,000 liters per day. Over one year, that would equal
10,000,000 liters or 2.6 million gallons of blood sent through roughly 60,000 miles
of vessels. In order to understand how that happens, it is necessary to understand the
anatomy and physiology of the heart.

Location of the Heart


The human heart is located within the thoracic cavity, medially between the lungs in the
space known as the mediastinum. Figure 9.2 shows the position of the heart within the
thoracic cavity. Within the mediastinum, the heart is separated from the other mediastinal
structures by a tough membrane known as the pericardium, or pericardial sac, and sits in its
own space called the pericardial cavity. It is important to remember the position and
Figure 9.2 Position of the Heart in the Thorax The heart is located within the thoracic cavity, medially
between the lungs in the mediastinum. It is about the size of a fist, is broad at the top, and tapers toward the
base.

orientation of the heart when placing a stethoscope on the chest of a patient and listening
for heart sounds, and also when looking at images taken from a midsagittal perspective. The
slight deviation of the apex to the left is reflected in a depression in the medial surface of
the inferior lobe of the left lung, called the cardiac notch.

Shape and Size of the Heart


The shape of the heart is similar to a pinecone, rather broad at the superior surface and
tapering to the apex (see Figure 9.2). A typical heart is approximately the size of your fist: 12
cm (5 in) in length, 8 cm (3.5 in) wide, and 6 cm (2.5 in) in thickness. Given the size
difference between most members of the sexes, the weight of a female heart is
approximately 250–300 grams (9 to 11 ounces), and the weight of a male heart is
approximately 300–350 grams (11 to 12 ounces). The heart of a well-trained athlete,
especially one specializing in aerobic sports, can be considerably larger than this. Cardiac
muscle responds to exercise in a manner similar to that of skeletal muscle. That is, exercise
results in the addition of protein myofilaments that increase the size of the individual cells
without increasing their numbers, a concept called hypertrophy. Hearts of athletes can
pump blood more effectively at lower rates than those of non athletes. Enlarged hearts are
not always a result of exercise; they can result from pathologies, such as hypertrophic
cardiomyopathy. The cause of an abnormally enlarged heart muscle is unknown, but the
condition is often undiagnosed and can cause sudden death in apparently otherwise healthy
young people.

Chambers and Circulation through the Heart


The human heart consists of four chambers: The left side and the right side each have
one atrium and one ventricle (Figure 9.3). Each of the upper chambers, the right atrium
(plural = atria) and the left atrium, acts as a receiving chamber and contracts to push blood
into the lower chambers, the right ventricle and the left ventricle. The ventricles serve as
the primary pumping chambers of the heart, propelling blood to the lungs or to the rest of
the body.
There are two distinct but linked circuits in the human circulation called the pulmonary and
systemic circuits (Figure 9.3). Although both circuits transport blood and everything it
carries, we can initially view the circuits from the point of view of gases.
The pulmonary circuit transports blood to and from the lungs, where it picks up oxygen and
delivers carbon dioxide for exhalation. The systemic circuit transports oxygenated blood to
virtually all of the tissues of the body and returns relatively deoxygenated blood and carbon
dioxide to the heart to be sent back to the pulmonary circulation.
Figure 9.3 Dual System of the Human Blood Circulation Blood flows from the right atrium to the right
ventricle, where it is pumped into the pulmonary circuit. The blood in the pulmonary artery branches is low in
oxygen but relatively high in carbon dioxide. Gas exchange occurs in the pulmonary capillaries (oxygen into the
blood, carbon dioxide out), and blood high in oxygen and low in carbon dioxide is returned to the left atrium.
From here, blood enters the left ventricle, which pumps it into the systemic circuit. Following exchange in the
systemic capillaries (oxygen and nutrients out of the capillaries and carbon dioxide and wastes in), blood
returns to the right atrium and the cycle is repeated.
9.2 BLOOD VESSELS AND CIRCULATION

Figure 9.4 Blood Vessels While most blood vessels are located deep from the surface and are not visible,
the superficial veins of the upper limb provide an indication of the extent, prominence, and importance of
these structures to the body. (credit: Colin Davis)

In this section, you will learn about the vascular part of the cardiovascular system, that is,
the vessels that transport blood throughout the body and provide the physical site where
gases, nutrients, and other substances are exchanged with body cells. When vessel
functioning is reduced, blood-borne substances do not circulate effectively throughout the
body. As a result, tissue injury occurs, metabolism is impaired, and the functions of every
bodily system are threatened.
Blood is carried through the body via blood vessels (Figure 9.4). An artery is a blood vessel
that carries blood away from the heart, where it branches into ever-smaller vessels.
Eventually, the smallest arteries, vessels called arterioles, further branch into tiny capillaries,
where nutrients and wastes are exchanged, and then combine with other vessels that exit
capillaries to form venules, small blood vessels that carry blood to a vein, a larger blood
vessel that returns blood to the heart. Arteries and veins transport blood in two distinct
circuits: the systemic circuit and the pulmonary circuit (Figure 9.5). The flow of blood
through the blood vessels causes pressure to be exerted on the walls of the blood vessels -
this is what we measure as 'blood pressure'.
Figure 9.5 Cardiovascular Circulation The pulmonary circuit moves blood from the right side of the heart to
the lungs and back to the heart. The systemic circuit moves blood from the left side of the heart to the head
and body and returns it to the right side of the heart to repeat the cycle. The arrows indicate the direction of
blood flow, and the colors show the relative levels of oxygen concentration.

The flow of blood through the blood vessels causes pressure to be exerted on the walls of
the blood vessels - this is what we measure as 'blood pressure'. Blood pressure varies
depending on activity and the tissues demand for blood supply. A number of homeostatic
mechanisms alter the diameter of the blood vessels or change heart rate to alter blood
pressure to meet demand and return to normal levels.
So, the heart pumps blood through two circuits: the pulmonary and systemic circuits. One of
the most important tasks this flow of blood through the body achieves is the distribution of
nutrients, including gases and those we eat, through the body, as well as the collection of
waste gases and other substances for expulsion from the body. Let’s now look at the reason
we breathe - it's all to do with getting those gases in to and out of the blood.
9.3 THE RESPIRATORY SYSTEM - Why we
Breathe

Figure 9.6 Mountain Climbers The thin air at high elevations can strain the human respiratory system.
(credit: “bortescristian”/flickr.com)

Hold your breath. Really! See how long you can hold your breath as you continue
reading…How long can you do it? Chances are you are feeling uncomfortable already. A
typical human cannot survive without breathing for more than 3 minutes, and even if you
wanted to hold your breath longer, your autonomic nervous system would take control. This
is because every cell in the body needs to run the oxidative stages of cellular respiration, the
process by which energy is produced in the form of adenosine triphosphate (ATP). You may
be surprised to learn that although oxygen is a critical need for cells, it is actually the
accumulation of carbon dioxide that primarily drives your need to breathe. Carbon dioxide
is exhaled and oxygen is inhaled through the respiratory system, which includes muscles to
move air into and out of the lungs, passageways through which air moves, and microscopic
gas exchange surfaces covered by capillaries. The circulatory system transports gases from
the lungs to tissues throughout the body and vice versa. A variety of diseases can affect
the respiratory system, such as asthma, emphysema, chronic obstruction pulmonary
disorder (COPD), and lung cancer. All of these conditions affect the gas exchange process
and result in labored breathing and other difficulties.

Organs and Structures of the Respiratory System

Figure 9.7 Major Respiratory Structures The major respiratory structures span the nasal cavity to the
diaphragm.

Functionally, the respiratory system can be divided into a conducting zone and a respiratory
zone. The conducting zone of the respiratory system includes the organs and structures not
directly involved in gas exchange. The gas exchange occurs in the respiratory zone.
Conducting Zone
The major functions of the conducting zone are to provide a route for incoming and
outgoing air, remove debris and pathogens from the incoming air, and warm and humidify
the incoming air. The conducting zone is made of all structures from the nose down to the
terminal bronchioles (Figure 9.7).

Respiratory Zone
In contrast to the conducting zone, the respiratory zone includes structures that are directly
involved in gas exchange. The respiratory zone begins where the terminal bronchioles join
a respiratory bronchiole, the smallest type of bronchiole (Figure 9.8), which then leads to
an alveolar duct, opening into a cluster of alveoli.

Figure 9.8 Respiratory Zone Bronchioles lead to alveolar sacs in the respiratory zone, where gas
exchange occurs.
Alveoli
An alveolar duct is a tube composed of smooth muscle and connective tissue, which opens
into a cluster of alveoli. An alveolus is one of the many small, grape-like sacs that are
attached to the alveolar ducts (Figure 9.9a).

Figure 9.9 Structures of the Respiratory Zone (a) The alveolus is responsible for gas exchange. (b) A micrograph
shows the alveolar structures within lung tissue. LM × 178. (Micrograph provided by the Regents of University
of Michigan Medical School © 2012)

The alveolar epithelium (Figure 9.9b) is extremely thin and borders the endothelial
membrane of capillaries. Taken together, the alveoli and capillary membranes form
a respiratory membrane that is approximately 0.5 mm thick. The respiratory membrane
allows gases to cross by simple diffusion, allowing oxygen to be picked up by the blood for
transport and CO2 to be released into the air of the alveoli.

9.4 The Lungs


A major organ of the respiratory system, each lung houses structures of both the conducting
and respiratory zones. The main function of the lungs is to perform the exchange of oxygen
and carbon dioxide with air from the atmosphere. To this end, the lungs exchange
respiratory gases across a very large epithelial surface area—about 70 square meters—that
is highly permeable to gases.
Gross Anatomy of the Lungs
The lungs are pyramid-shaped, paired organs that are connected to the trachea by the right
and left bronchi (Figure 9.10); on the inferior surface, the lungs are bordered by the
diaphragm. The diaphragm is the flat, dome-shaped muscle located at the base of the lungs
and thoracic cavity.

Figure 9.10 Gross Anatomy of the Lungs

Pleura of the Lungs


Each lung is enclosed within a cavity that is surrounded by the pleura. The pleura (plural =
pleurae) is a serous membrane that surrounds the lung. The pleurae consist of two layers.
The visceral pleura is the layer that is superficial to the lungs, and extends into and lines the
lung fissures (Figure 9.11). In contrast, the parietal pleura is the outer layer that connects to
the thoracic wall. The pleural cavity is the space between the visceral and parietal layers.
The pleurae produce pleural fluid which acts to lubricate their surfaces. This
lubrication reduces friction between the two layers to prevent trauma during breathing, and
creates surface tension that helps maintain the position of the lungs against the thoracic
wall. This adhesive characteristic of the pleural fluid causes the lungs to enlarge when the
thoracic wall expands during ventilation, allowing the lungs to fill with air.
Figure 9.11 Parietal and Visceral Pleurae of the Lungs

The Process of Breathing


Pulmonary ventilation is the act of breathing, which can be described as the movement of
air into and out of the lungs. The major mechanisms that drive pulmonary ventilation are
atmospheric pressure (Patm); the air pressure within the alveoli, called alveolar pressure
(Palv); and the pressure within the pleural cavity, called intrapleural pressure (Pip).

Mechanisms of Breathing
The ability to breathe—to have air enter the lungs during inspiration and air leave the lungs
during expiration—is dependent on the air pressure of the atmosphere and the air pressure
within the lungs.

Pressure Relationships
Boyle’s law describes the relationship between volume and pressure in a gas at a constant
temperature. Boyle discovered that the pressure of a gas is inversely proportional to its
volume: If volume increases, pressure decreases (Figure 9.12). Likewise, if volume
decreases, pressure increases. Boyle’s law is expressed by the following formula:
P1V1 = P2V2
In this formula, P1 represents the initial pressure and V1 represents the initial volume,
whereas the final pressure and volume are represented by P2 and V2, respectively. If the
two- and one-liter containers were connected by a tube and the volume of one of the
containers were changed, then the gases would move from higher pressure (lower volume)
to lower pressure (higher volume).

Figure 9.12 Boyle's Law In a gas, pressure increases as volume decreases.

Pulmonary ventilation is dependent on atmospheric and intra-alveolar


pressures. Atmospheric pressure is the amount of force that is exerted by gases in the air
surrounding any given surface, such as the body. Atmospheric pressure can be expressed in
terms of millimeters of mercury (mm Hg). 760 mm Hg is the atmospheric pressure at sea
level. Intra-alveolar pressure is the pressure of the air within the alveoli (Figure 9.13).
Because the alveoli are connected to the atmosphere via the tubing of the airways (similar
to the two- and one-liter containers in the example above), the interpulmonary pressure of
the alveoli always equalizes with the atmospheric pressure.

Figure 9.13 Intrapulmonary and Intrapleural Pressure Relationships Alveolar pressure changes during
the different phases of the cycle. It equalizes at 760 mm Hg but does not remain at 760 mm Hg.

Physical Factors Affecting Ventilation


In addition to the differences in pressures, breathing is also dependent upon the contraction
and relaxation of muscle fibers of both the diaphragm and thorax. The lungs themselves are
passive during breathing, meaning they are not involved in creating the movement that
helps inspiration and expiration. This is because of the adhesive nature of the pleural
fluid, which allows the lungs to be pulled outward when the thoracic wall moves during
inspiration. The recoil of the thoracic wall during expiration causes compression of the
lungs. Contraction and relaxation of the diaphragm and intercostals muscles (found
between the ribs) cause most of the pressure changes that result in inspiration and
expiration. These muscle movements and subsequent pressure changes cause air to either
rush in or be forced out of the lungs.
The difference in pressures drives pulmonary ventilation because air flows down a pressure
gradient, that is, air flows from an area of higher pressure to an area of lower pressure. Air
flows into the lungs largely due to a difference in pressure; atmospheric pressure is greater
than intra-alveolar pressure, and intra-alveolar pressure is greater than intrapleural
pressure. Air flows out of the lungs during expiration based on the same principle; pressure
within the lungs becomes greater than the atmospheric pressure.
Pulmonary ventilation comprises two major steps: inspiration and expiration. Inspiration is
the process that causes air to enter the lungs, and expiration is the process that causes air
to leave the lungs (Figure 9.14). A respiratory cycle is one sequence of inspiration and
expiration. In general, two muscle groups are used during normal inspiration: the
diaphragm and the external intercostal muscles. Additional muscles can be used if a bigger
breath is required. When the diaphragm contracts, it moves inferiorly toward the abdominal
cavity, creating a larger thoracic cavity and more space for the lungs. Contraction of the
external intercostal muscles moves the ribs upward and outward, causing the rib cage to
expand, which increases the volume of the thoracic cavity. Due to the adhesive force of the
pleural fluid, the expansion of the thoracic cavity forces the lungs to stretch and expand as
well. This increase in volume leads to a decrease in intra-alveolar pressure, creating a
pressure lower than atmospheric pressure. As a result, a pressure gradient is created that
drives air into the lungs.

Figure 9.14 Inspiration and Expiration Inspiration and expiration occur due to the expansion and contraction
of the thoracic cavity, respectively.
The process of normal expiration is passive, meaning that energy is not required to push air
out of the lungs. Instead, the elasticity of the lung tissue causes the lung to recoil, as the
diaphragm and intercostal muscles relax following inspiration. In turn, the thoracic cavity
and lungs decrease in volume, causing an increase in interpulmonary pressure. The
interpulmonary pressure rises above atmospheric pressure, creating a pressure gradient
that causes air to leave the lungs.
The lungs are so cool! They achieve the important job, of gas exchange in to and out of the
blood in such a simple and efficient process. Let’s now turn to why we eat - and this is all
about how we get nutrients from foods into the blood!

9.5 DIGESTIVE SYSTEM - Why we Eat


All living organisms need nutrients to survive. While plants can obtain nutrients from their
roots and the energy molecules required for cellular function through the process of
photosynthesis, animals obtain their nutrients by the consumption of other organisms. At
the cellular level, the biological molecules necessary for animal function are amino acids,
lipid molecules, nucleotides, and simple sugars. However, the food consumed consists of
protein, fat, and complex carbohydrates. Animals must convert these macromolecules into
the simple molecules required for maintaining cellular function. The conversion of the food
consumed to the nutrients required is a multistep process involving digestion and
absorption. During digestion, food particles are broken down to smaller components, which
are later absorbed by the body. This happens by both physical means, such as chewing, and
by chemical means.
One of the challenges in human nutrition is maintaining a balance between food intake,
storage, and energy expenditure. Taking in more food energy than is used in activity leads to
storage of the excess in the form of fat deposits. The rise in obesity and the resulting
diseases like type 2 diabetes makes understanding the role of diet and nutrition in
maintaining good health all the more important.

The Human Digestive System


The process of digestion begins in the mouth with the intake of food (Figure 9.15). The teeth
play an important role in masticating (chewing) or physically breaking food into smaller
particles. The enzymes present in saliva also begin to chemically break down food. The food
is then swallowed and enters the esophagus—a long tube that connects the mouth to the
stomach. Using peristalsis, or wave-like smooth-muscle contractions, the muscles of the
esophagus push the food toward the stomach. The stomach contents are extremely acidic,
with a pH between 1.5 and 2.5. This acidity kills microorganisms, breaks down food tissues,
and activates digestive enzymes. Further breakdown of food takes place in the small
intestine where bile produced by the liver, and enzymes produced by the small intestine and
the pancreas, continue the process of digestion. The smaller molecules are absorbed into
the blood stream through the epithelial cells lining the walls of the small intestine. The
waste material travels on to the large intestine where water is absorbed and the drier waste
material is compacted into faeces; it is stored until it is excreted through the anus.

Figure 9.15 The components of the human digestive system are shown.
Oral Cavity
Both physical and chemical digestion begin in the mouth or oral cavity, which is the point of
entry of food into the digestive system. The food is broken into smaller particles by
mastication, the chewing action of the teeth. All mammals have teeth and can chew their
food to begin the process of physically breaking it down into smaller particles.
The chemical process of digestion begins during chewing as food mixes with saliva,
produced by the salivary glands (Figure 9.16). Saliva contains mucus that moistens food and
buffers the pH of the food. Saliva also contains lysozyme, which has antibacterial action. It
also contains an enzyme called salivary amylase that begins the process of converting
starches in the food into a disaccharide called maltose. Another enzyme called lipase is
produced by cells in the tongue to break down fats. The chewing and wetting action
provided by the teeth and saliva prepare the food into a mass called the bolus for
swallowing. The tongue helps in swallowing—moving the bolus from the mouth into the
pharynx. The pharynx opens to two passageways: the esophagus and the trachea. The
esophagus leads to the stomach and the trachea leads to the lungs. The epiglottis is a flap of
tissue that covers the tracheal opening during swallowing to prevent food from entering the
lungs.

Figure 9.16 (a) Digestion of food begins in the mouth. (b) Food is masticated by teeth and moistened by saliva
secreted from the salivary glands. Enzymes in the saliva begin to digest starches and fats. With the help of the
tongue, the resulting bolus is moved into the esophagus by swallowing. (credit: modification of work by
Mariana Ruiz Villareal)
Oesophagus
The esophagus is a tubular organ that connects the mouth to the stomach. The chewed and
softened food passes through the esophagus after being swallowed. The smooth muscles of
the esophagus undergo peristalsis that pushes the food toward the stomach. The peristaltic
wave is unidirectional—it moves food from the mouth the stomach, and reverse movement
is not possible, except in the case of the vomit reflex. The peristaltic movement of the
esophagus is an involuntary reflex; it takes place in response to the act of swallowing.
Ring-like muscles called sphincters form valves in the digestive system. The gastro-
esophageal sphincter (or cardiac sphincter) is located at the stomach end of the esophagus.
In response to swallowing and the pressure exerted by the bolus of food, this sphincter
opens, and the bolus enters the stomach. When there is no swallowing action, this sphincter
is shut and prevents the contents of the stomach from traveling up the esophagus. Acid
reflux or “heartburn” occurs when the acidic digestive juices escape into the esophagus.

Stomach
A large part of protein digestion occurs in the stomach (Figure 9.17). The stomach is a
saclike organ that secretes gastric digestive juices.
Protein digestion is carried out by an enzyme called pepsin in the stomach chamber. The
highly acidic environment kills many microorganisms in the food and, combined with the
action of the enzyme pepsin, results in the catabolism of protein in the food. Chemical
digestion is facilitated by the churning action of the stomach caused by contraction and
relaxation of smooth muscles. The partially digested food and gastric juice mixture is
called chyme. Gastric emptying occurs within two to six hours after a meal. Only a small
amount of chyme is released into the small intestine at a time. The movement of chyme
from the stomach into the small intestine is regulated by hormones, stomach distension and
muscular reflexes that influence the pyloric sphincter.
The stomach lining is unaffected by pepsin and the acidity because pepsin is released in an
inactive form and the stomach has a thick mucus lining that protects the underlying tissue.

Small Intestine
Chyme moves from the stomach to the small intestine. The small intestine is the organ
where the digestion of protein, fats, and carbohydrates is completed. The small intestine is a
long tube-like organ with a highly folded surface containing finger-like projections called the
villi. The top surface of each villus has many microscopic projections called microvilli. The
epithelial cells of these structures absorb nutrients from the digested food and release them
to the bloodstream on the other side. The villi and microvilli, with their many folds, increase
the surface area of the small intestine and increase absorption efficiency of the nutrients.
Figure 9.17 The stomach has an extremely acidic environment where most of the protein gets digested.
(credit: modification of work by Mariana Ruiz Villareal)

The human small intestine is over 6 m (19.6 ft) long and is divided into three parts: the
duodenum, the jejunum and the ileum. The duodenum is separated from the stomach by
the pyloric sphincter. The chyme is mixed with pancreatic juices, an alkaline solution rich in
bicarbonate that neutralizes the acidity of chyme from the stomach. Pancreatic juices
contain several digestive enzymes that break down starches, disaccharides, proteins, and
fats. Bile is produced in the liver and stored and concentrated in the gallbladder; it enters
the duodenum through the bile duct. Bile contains bile salts, which make lipids accessible to
the water-soluble enzymes. The monosaccharides, amino acids, bile salts, vitamins, and
other nutrients are absorbed by the cells of the intestinal lining.
The undigested food is sent to the colon from the ileum via peristaltic movements. The
ileum ends and the large intestine begins at the ileocecal valve. The vermiform, “worm-like,”
appendix is located at the ileocecal valve. The appendix of humans has a minor role in
immunity.

Large Intestine
The large intestine reabsorbs the water from indigestible food material and processes the
waste material (Figure 9.18). The human large intestine is much smaller in length compared
to the small intestine but larger in diameter. It has three parts: the cecum, the colon, and
the rectum. The cecum joins the ileum to the colon and is the receiving pouch for the waste
matter. The colon is home to many bacteria or “intestinal flora” that aid in the digestive
Figure 9.18 The large intestine reabsorbs water from undigested food and stores waste until it is eliminated.
(credit: modification of work by Mariana Ruiz Villareal)

processes. The colon has four regions, the ascending colon, the transverse colon, the
descending colon and the sigmoid colon. The main functions of the colon are to extract the
water and mineral salts from undigested food, and to store waste material.
The rectum (Figure 9.18) stores faeces until defecation. The faeces are propelled using
peristaltic movements during elimination. The anus is an opening at the far-end of the
digestive tract and is the exit point for the waste material. Two sphincters regulate the exit
of faeces, the inner sphincter is involuntary and the outer sphincter is voluntary.

Accessory Organs
The organs discussed above are the organs of the digestive tract through which food passes.
Accessory organs add secretions and enzymes that break down food into nutrients.
Accessory organs include the salivary glands, the liver, the pancreas, and the gall bladder.
The secretions of the liver, pancreas, and gallbladder are regulated by hormones in
response to food consumption.
The liver is the largest internal organ in humans and it plays an important role in digestion
of fats and detoxifying blood. The liver produces bile, a digestive juice that is required for
the breakdown of fats in the duodenum. The liver also processes the absorbed vitamins and
fatty acids and synthesizes many plasma proteins. The gallbladder is a small organ that aids
the liver by storing bile and concentrating bile salts.
The pancreas secretes bicarbonate that neutralizes the acidic chyme and a variety of
enzymes for the digestion of protein and carbohydrates.
Ok, so we beat to get blood to flow around the body. We breathe to get gases in to and out
of the blood and we eat to get nutrients from food into the blood. Do you remember what
we do with the gases and nutrients? Here's a hint: metabolism.
Try the review questions below to test what you have learned from this chapter.
REVIEW QUESTIONS
Question 9.1
The heart is located in the:
A thoracic cavity
B mediastinum
C pericardial cavity
D all of the above are true

Question 9.2
The approximate weight of the female heart is?
A 150 - 200 g
B 200 - 250 g
C 250 - 300 g
D 300-350 g

Question 9.3
The Left side of the heart pumps blood to:
A the systemic circuit
B the right side of the heart
C the pulmonary circuit
D right ventricle
Question 9.4
The right side of the heart pumps blood to:
A the systemic circuit
B the left side of the heart
C the pulmonary circuit
D left ventricle

Question 9.5
Arteries are blood vessels which carry:
A deoxygenated blood
B blood toward the heart
C oxygenated blood
D blood away from the heart

Question 9.6
Veins are blood vessels which carry:
A deoxygenated blood
B blood toward the heart
C oxygenated blood
D blood away from the heart

Question 9.7
Nutrient and waste exchange occurs in the:
A arteries
B arterioles
C capillaries
D venules
E veins
Question 9.8
Blood pressure can be altered by:
A changing the diameter of the blood vessels
B changing the rate of the heart beat
C both of the above are true

Question 9.9
The main purpose of the respiratory system is to provide a mechanism by which carbon
dioxide can enter and oxygen leave, the body
A. True
B. False

Question 9.10
The conducting zone of the respiratory system:
A. Is where gas exchange occurs
B. Includes respiratory bronchioles
C. Warms and humidifies air
D. Includes alveoli

Question 9.11
The respiratory zone of the respiratory system:
A. Is where gas exchange occurs
B. Includes terminal bronchioles
C. Warms and humidifies air
D. Includes the bronchi
Question 9.12
The respiratory membrane:
A. Is formed by the alveolar and capillary membranes
B. Allows gas diffusion
C. Is approximately 0.5 mm thick
D. All of the above are true

Question 9.13
Lungs house structures of both the conducting and respiratory zones. The main function of
the lungs is to perform the exchange of oxygen and carbon dioxide. They exchange these
gases across a very large epithelial surface area of about 70 square metres that is
highly permeable to gases.
A. True
B. False

Question 9.14
Features of the pleura include
A. They surround the lungs
B. They produce pleural fluid
C. They help form a vacuum which sucks the lungs to the chest cavity
D. A and B only are true
E. A, B and C are all true

Question 9.15
Boyle’s Law states that increasing the volume of a gas will increase its pressure
A. True
B. False
Question 9.16
During inspiration the chest cavity and lungs expand, increasing the volume of the lung and
increasing intra-alveolar pressure. When intra-alveolar pressure is higher than atmospheric
pressure, air flows in to the lung.
A. True
B. False

Question 9.17
While all living things need nutrients to survive, plants are able to obtain nutrients from
their roots and via photosynthesis, while humans (and other animals) obtain their nutrients
by the consumption of other organisms.
A. True
B. False

Question 9.18
Which of the following lists contains the structures of the digestive system in the correct
order?
A. Mouth, stomach, oesophagus, large intestine, small intestine, anus
B. Mouth, large intestine, oesophagus, stomach, small intestine, anus
C. Mouth, oesophagus, small intestine, stomach, large intestine, anus
D. Mouth, oesophagus, stomach, small intestine, large intestine, anus
E. Mouth, oesophagus, stomach, large intestine, small intestine, anus

Question 9.19
Physical digestion occurs in the
A. Mouth
B. Stomach
C. Intestines
D. Mouth and Stomach
E. Mouth, Stomach, and Intestines
Question 9.20
Chemical digestion occurs in the
A. Mouth
B. Stomach
C. Intestines
D. Mouth and Stomach
E. Mouth, Stomach, and Intestines

Question 9.21
Absorption of nutrients occurs in the
A. Mouth
B. Stomach
C. Intestines
D. Mouth and Stomach
E. Mouth, Stomach, and Intestines

Question 9.22
The accessory organs of the digestive system include the liver, gall bladder and pancreas.
A. True
B. False
ANSWERS
Question 9.1
D

Question 9.2
C

Question 9.3
B

Question9.4
B

Question 9.5
C, D

Question 9.6
A, B

Question 9.7
D

Question9.8
C

Question 9.9
B

Question 9.10
C

Question 9.11
A

Question 9.12
D

Question 9.13
A

Question 9.14
E
Question 9.15
B

Question 9.16
B

Question 9.17
A

Question 9.18
D

Question 9.19
D
Question 9.20
E

Question 9.21
C

Question 9.22
A
10 | COMMUNICATION IN
THE BODY

Figure 10.0: The nervous system communicates using electrical signals

By the end of this chapter you will be able to:


• Identify the importance of cellular communication
• Describe the steps in cellular communication
• Indicate the general characteristics of excitable nervous tissue.
• Briefly describe the organisation of the nervous system
• Describe the 3 basic functions of the nervous system.
• Name examples of diseases where function of the nervous system is compromised.
• Compare and contrast the actions of the endocrine and nervous systems in
maintaining homeostasis
• Draw, using a flow chart, a generalised endocrine system.
• Recognise blood as a conduit of endocrine communication.
• Compare how endocrine disorders can arise from too much or too little hormone
activity. (Give an example of each)

10.1 Cell Communication


Cellular communication is an important process that allows cells to pass signals to one
another.

Remember playing the game 'telephone' as a kid? One person started with a message that
got passed down the line through whispers. As the message made its way from person to
person, it changed into something different and by the end may have had nothing to do
with the original statement. This happens because there is a communication problem along
the line. It shows you just how important good communication is, even with something as
simple as a child's game.

Your cells also rely on good communication, but unlike a game of telephone, if there is a
miscommunication, the consequences can be quite severe!

Cell-to-cell signaling is a critical component of coordinating cellular activities. Through this


communication, messages are carried from signaling cells to receiving cells, also known as
target cells. This signaling occurs with proteins and other types of signaling molecules. What
kinds of processes and activities depend on cell communication? Just about everything you
can imagine that happens in your body - growth and development, cellular reproduction,
tissue repair, sensing pain, and much more.

How Cells Communicate


The pathway along which cellular communication occurs is called the signal transduction
pathway. It's a series of cellular and molecular changes that help a target cell receive a
message and respond accordingly. These pathways are crucial to cellular function because
they're the main mode of communication between cells.

Here's how it works. First, a signaling cell secretes a signaling molecule, which then travels
with its message to the target cell. Once it reaches the target cell, it binds with a receptor
protein in the target cell's membrane. This is the first step of cell signaling, called reception.
This receptor protein is like the security guard at an entrance gate - it's the designated
'check-in' point for the signaling molecule. When the receptor protein security guard
accepts the signaling molecule and binds to it, a relay begins, which is the second step of cell
signaling, called transduction. Just like a baton is passed from person to person in a relay
race, so is the signal in the target cell through the process of transduction as it passes from
relay molecule to relay molecule.

The final relay molecule activates a protein that triggers the final stage of cell signaling,
called response. Each response is different because it depends on the message that was
carried through the cell.

Cell Communication Is Universal


Just how important is cell communication? It seems that this process is a fairly ancient one
and one that most organisms have in common. For example, cell communication is
necessary for reproduction, not just of individual cells, but also for entire organisms.

Let's look at yeast as an example. Some yeast cells not only make delicious bread and beer,
but they also help us understand how chemical signals can be used to communicate
complex information between cells. Even without the ability to make flirty eye contact,
these organisms can tell which cells will help them reproduce and create new offspring, and
they do this by identifying their 'mates' through chemical signals.

The most amazing part? This signaling pathway found in the yeast cells is incredibly similar
to the signaling pathways found in other organisms, like mammals. Yeast and mammals
aren't closely related at all. In fact, yeast, which are single-celled organisms, were around a
long time before multicellular organisms (like mammals), so the similarity between the
pathways tells us that this is both an ancient and important mechanism for sustaining life on
Earth because it has endured through many different evolutionary steps.

Summary
Cellular activity and function relies on good communication between cells. Through
signaling molecules and pathways, cells can 'talk' to each other. These communications help
cells develop, repair tissues, reproduce, and much more.

The signals are transmitted from a signaling cell, which sends out a signaling molecule.
During this first stage of cell signaling, called reception, the signaling molecule binds to the
receptor protein. The second step, called transduction, occurs when the signal from the
signaling molecule travels along the transduction pathway, like a baton through a relay race.
It's during the final stage, called response, that the signal finally gets converted into a
message the target cell can understand.
You should be glad that your cells are better at communicating than a group of people
playing telephone. It would be pretty detrimental to your body if the messages inside got as
distorted as they do when we whisper them in someone else's ear!

10.2 Excitable Cells


In the early 19th century, the scientist Luigi Galvani accidentally touched his metal scalpel to
a muscle in a dissected frog leg and observed a subtle twitch. That small observation gave
birth to an entire field of study dedicated to understanding how excitable cells like muscle
and neurons generate, propagate, and send electrical signals. Throughout the 19th century,
curious scientists attempted to build on Galvani’s work by electrocuting limbs, heads, and
even entire cadavers to better understand the link between electricity and the physiology of
muscle and nervous tissues. If this sounds like the plot of a horror story, it is. Many people
credit these early “reanimation” experiments with being the inspiration for Mary Shelley’s
famous novel Frankenstein.
Originally called Galvanism, the study of excitability is referred to as electrophysiology by
modern scientists. Nearly 200 years since the field began, we now understand that
electricity is a major method of signalling and cell to cell communication used by the
nervous system and muscle tissue. While this electricity is similar to the electricity carried
within a wire conceptually, there are some differences that will be explored later. In this
chapter, we will explore how electrical signals are generated in a biological system and then
talk specifically about excitability in neurons (muscle excitability depends on these same
principles, but is covered in its own chapter later in the text). Excitability is often a
challenging subject for students because it is based on things that are impossible to see and
difficult to imagine. Don’t be discouraged – it has been a challenging area of study for
scientists too! However, understanding many of the systems we will talk about later in the
textbook will require applying your knowledge of excitability. As we move through this
chapter we will briefly review some key points about membrane transport that you learned
about in the cell biology chapter, but if you don’t have a good grasp on the concepts of
diffusion and carrier mediated transport, then it’s recommended you revisit that chapter
before working through this one.

10.3 Creating Membrane Potential and an


Action Potential in a Biological System
The functions of the nervous system—sensation, integration, and response—depend on the
functions of the neurons underlying these pathways. To understand how neurons are able
to communicate, it is necessary to describe the role of an excitable membrane in generating
these signals. The basis of this communication is the action potential, which demonstrates
how changes in the membrane can constitute a signal. Looking at the way these signals
work in more variable circumstances involves a look at graded potentials, which will be
covered in the next section.

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 of the 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 assistance (Figure 10.1).
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.

Figure 10.1 Cell Membrane and Transmembrane Proteins. The cell membrane is composed of a
phospholipid bilayer and has many transmembrane proteins, including different types of channel proteins that
serve as ion channels.
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 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 opposed 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 categorised 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 (Figure 10.2).
Figure 10.2 Ligand-Gated Channels When the ligand, in this case the neurotransmitter acetylcholine, binds to
a specific location on the extracellular surface of the channel protein, the pore opens to allow select ions
through. The ions, in this case, are cations of sodium, calcium, and potassium.

A mechanically gated channel opens because of a physical distortion of the cell membrane.
Many channels associated with the sense of touch (somatosensation) are mechanically
gated. For example, as pressure is applied to the skin, these channels open and allow ions to
enter the 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 (Figure 10.3).

Figure 10.3 Mechanically Gated Channels When a mechanical change occurs in the surrounding tissue, such
as pressure or touch, the channel is physically opened. Thermoreceptors work on a similar principle. When the
local tissue temperature changes, the protein reacts by physically opening the channel.
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 (Figure 10.4).

Figure 10.4 Voltage-Gated Channels Voltage-gated channels open when the transmembrane voltage
changes around them. Amino acids in the structure of the protein are sensitive to charge and cause the pore to
open to the selected ion.

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 (Figure 10.5).

Figure 10.5 Leakage Channels In certain situations, ions need to move across the membrane randomly.
The particular electrical properties of certain cells are modified by the presence of this type of channel.
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 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 (Figure 10.6).

Figure 10.6 Measuring Charge across a Membrane with a Voltmeter A recording electrode is inserted into
the cell and a reference electrode is outside the cell. By comparing the charge measured by these two
electrodes, the transmembrane voltage is determined. It is conventional to express that value for the cytosol
relative to the outside.

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 specialised 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 localised 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 depolarisation, meaning the membrane potential moves toward zero.
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 repolarisation, meaning that the membrane voltage moves back toward the -70 mV
value of the resting membrane potential. Repolarisation 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 hyperpolarisation occurs while the K+ channels are open. Those K+ channels are slightly
delayed in closing, accounting for this short overshoot.
Figure 10.7 Graph of Action Potential Plotting voltage measured across the cell membrane against time,
the action potential begins with depolarisation, followed by repolarisation, which goes past the resting
potential into hyperpolarisation, and finally the membrane returns to rest.

What has been described here is the action potential, which is presented as a graph of
voltage over time in Figure 10.7. 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 depolarisation 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.
What happens across the membrane of an electrically active cell is a dynamic process that is
hard to visualise with static images or through text descriptions. View this animation
(http://openstaxcollege.org/l/dynamic1) to learn more about this process. What is the
difference between the driving force for Na+ and K+? And what is similar about the
movement of these two ions?
10.4 The Nervous System
Summary
The nervous system is made up of neurons and glia. Neurons are specialized cells that are
capable of sending electrical as well as chemical signals. Most neurons contain dendrites,
which receive these signals, and axons that send signals to other neurons or tissues. Glia are
non-neuronal cells in the nervous system that support neuronal development and signaling.
There are several types of glia that serve different functions.
Neurons have a resting potential across their membranes and when they are stimulated by
a strong enough signal from another neuron an action potential may carry an
electrochemical signal along the neuron to a synapse with another neuron.
Neurotransmitters carry signals across synapses to initiate a response in another neuron.
The vertebrate central nervous system contains the brain and the spinal cord, which are
covered and protected by three meninges. The brain contains structurally and functionally
defined regions. In mammals, these include the cortex (which can be broken down into four
primary functional lobes: frontal, temporal, occipital, and parietal), basal ganglia, thalamus,
hypothalamus, limbic system, cerebellum, and brainstem—although structures in some of
these designations overlap. While functions may be primarily localized to one structure in
the brain, most complex functions, like language and sleep, involve neurons in multiple
brain regions. The spinal cord is the information superhighway that connects the brain with
the rest of the body through its connections with peripheral nerves. It transmits sensory and
motor input and also controls motor reflexes.
The peripheral nervous system contains both the autonomic and sensory-somatic nervous
systems. The autonomic nervous system provides unconscious control over visceral
functions and has two divisions: the sympathetic and parasympathetic nervous systems. The
sympathetic nervous system is activated in stressful situations to prepare the animal for a
“fight-or-flight” response. The parasympathetic nervous system is active during restful
periods. The sensory-somatic nervous system is made of cranial and spinal nerves that
transmit sensory information from skin and muscle to the CNS and motor commands from
the CNS to the muscles.
As you read this, your nervous system is performing several functions simultaneously. The
visual system is processing what is seen on the page; the motor system controls your eye
movements and the turn of the pages (or click of the mouse); the prefrontal cortex
maintains attention. Even fundamental functions, like breathing and regulation of body
temperature, are controlled by the nervous system. The nervous system is one of two
systems that exert control over all the organ systems of the body; the other is the endocrine
system. The nervous system’s control is much more specific and rapid than the hormonal
system. It communicates signals through cells and the tiny gaps between them rather than
through the circulatory system as in the endocrine system. It uses a combination of chemical
and electrochemical signals, rather than purely chemical signals used by the endocrine
system to cover long distances quickly. The nervous system acquires information from
sensory organs, processes it and then may initiate a response either through motor
function, leading to movement, or in a change in the organism’s physiological state.
Nervous systems throughout the animal kingdom vary in structure and complexity. Some
organisms, like sea sponges, lack a true nervous system. Others, like jellyfish, lack a true
brain and instead have a system of separate but connected nerve cells (neurons) called a
“nerve net.” Flatworms have both a central nervous system (CNS), made up of a ganglion
(clusters of connected neurons) and two nerve cords, and a peripheral nervous system
(PNS) containing a system of nerves that extend throughout the body. The insect nervous
system is more complex but also fairly decentralized. It contains a brain, ventral nerve cord,
and ganglia. These ganglia can control movements and behaviors without input from the
brain.
Compared to invertebrates, vertebrate nervous systems are more complex, centralized, and
specialized. While there is great diversity among different vertebrate nervous systems, they
all share a basic structure: a CNS that contains a brain and spinal cord and a PNS made up of
peripheral sensory and motor nerves. One interesting difference between the nervous
systems of invertebrates and vertebrates is that the nerve cords of many invertebrates are
located ventrally (toward the stomach) whereas the vertebrate spinal cords are located
dorsally (toward the back). There is debate among evolutionary biologists as to whether
these different nervous system plans evolved separately or whether the invertebrate body
plan arrangement somehow “flipped” during the evolution of vertebrates.
The nervous system is made up of neurons, specialized cells that can receive and transmit
chemical or electrical signals, and glia, cells that provide support functions for the neurons.
There is great diversity in the types of neurons and glia that are present in different parts of
the nervous system.

Basic Functions of the Nervous System


The nervous system is involved in receiving information about the environment around us
(sensation) and generating responses to that information (motor responses). The nervous
system can be divided into regions that are responsible for sensation (sensory functions)
and for the response (motor functions). But there is a third function that needs to
be included. Sensory input needs to be integrated with other sensations, as well as with
memories, emotional state, or learning (cognition). Some regions of the nervous system are
termed integration or association areas. The process of integration combines sensory
perceptions and higher cognitive functions such as memories, learning, and emotion to
produce a response.
Sensation. The first major function of the nervous system is sensation—receiving
information about the environment to gain input about what is happening outside the body
(or, sometimes, within the body). The sensory functions of the nervous system register the
presence of a change from homeostasis or a particular event in the environment, known as
a stimulus. The senses we think of most are the “big five”: taste, smell, touch, sight, and
hearing. The stimuli for taste and smell are both chemical substances (molecules,
compounds, ions, etc.), touch is physical or mechanical stimuli that interact with the skin,
sight is light stimuli, and hearing is the perception of sound, which is a physical stimulus
similar to some aspects of touch. There are actually more senses than just those, but that
list represents the major senses. Those five are all senses that receive stimuli from the
outside world, and of which there is conscious perception. Additional sensory stimuli might
be from the internal environment (inside the body), such as the stretch of an organ wall or
the concentration of certain ions in the blood.
Response. The nervous system produces a response on the basis of the stimuli perceived by
sensory structures. An obvious response would be the movement of muscles, such as
withdrawing a hand from a hot stove, but there are broader uses of the term. The nervous
system can cause the contraction of all three types of muscle tissue. For example, skeletal
muscle contracts to move the skeleton, cardiac muscle is influenced as heart rate increases
during exercise, and smooth muscle contracts as the digestive system moves food along the
digestive tract. Responses also include the neural control of glands in the body as well, such
as the production and secretion of sweat by the eccrine and merocrine sweat glands found
in the skin to lower body temperature.
Responses can be divided into those that are voluntary or conscious (contraction of skeletal
muscle) and those that are involuntary (contraction of smooth muscles, regulation of
cardiac muscle, activation of glands). Voluntary responses are governed by the somatic
nervous system and involuntary responses are governed by the autonomic nervous system,
which are discussed in the next section.
Integration. Stimuli that are received by sensory structures are communicated to the
nervous system where that information is processed. This is called integration. Stimuli are
compared with, or integrated with, other stimuli, memories of previous stimuli, or the state
of a person at a particular time. This leads to the specific response that will be generated.
Seeing a baseball pitched to a batter will not automatically cause the batter to swing. The
trajectory of the ball and its speed will need to be considered. Maybe the count is three
balls and one strike, and the batter wants to let this pitch go by in the hope of getting a walk
to first base. Or maybe the batter’s team is so far ahead, it would be fun to just swing away.

Neurons and Glial Cells


The nervous system of the common laboratory fly, Drosophila melanogaster, contains
around 100,000 neurons, the same number as a lobster. This number compares to 75
million in the mouse and 300 million in the octopus. A human brain contains around 86
billion neurons. Despite these very different numbers, the nervous systems of these animals
control many of the same behaviors—from basic reflexes to more complicated behaviors
like finding food and courting mates. The ability of neurons to communicate with each other
as well as with other types of cells underlies all of these behaviors.
Most neurons share the same cellular components. But neurons are also highly
specialized—different types of neurons have different sizes and shapes that relate to their
functional roles.
Like other cells, each neuron has a cell body (or soma) that contains a nucleus, smooth and
rough endoplasmic reticulum, Golgi apparatus, mitochondria, and other cellular
components. Neurons also contain unique structures for receiving and sending the electrical
signals that make communication between neurons possible (Figure 10.8). Dendrites are
tree-like structures that extend away from the cell body to receive messages from other
neurons at specialized junctions called synapses. Although some neurons do not have any
dendrites, most have one or many dendrites.
The bilayer lipid membrane that surrounds a neuron is impermeable to ions. To enter or exit
the neuron, ions must pass through ion channels that span the membrane. Some ion
channels need to be activated to open and allow ions to pass into or out of the cell. These
ion channels are sensitive to the environment and can change their shape accordingly. Ion
channels that change their structure in response to voltage changes are called voltage-gated
ion channels. The difference in total charge between the inside and outside of the cell is
called the membrane potential. A neuron at rest is negatively charged: the inside of a cell is
approximately 70 millivolts more negative than the outside (–70 mV). This voltage is called
the resting membrane potential as discussed earlier in this chapter.
A neuron can receive input from other neurons and, if this input is strong enough, send the
signal to downstream neurons. Transmission of a signal between neurons is generally
carried by a chemical, called a neurotransmitter, which diffuses from the axon of one
neuron to the dendrite of a second neuron. When neurotransmitter molecules bind to
receptors located on a neuron’s dendrites, the neurotransmitter opens ion channels in the
dendrite’s plasma membrane. This opening allows sodium ions to enter the neuron and
results in depolarization of the membrane—a decrease in the voltage across the neuron
membrane. Once a signal is received by the dendrite, it then travels passively to the cell
body. A large enough signal from neurotransmitters will reach the axon. If it is strong
enough (that is, if the threshold of excitation, a depolarization to around –60mV is
reached), then depolarization creates a positive feedback loop: as more Na+ ions enter the
cell, the axon becomes further depolarized, opening even more sodium channels at further
distances from the cell body. This will cause voltage dependent Na+ channels further down
the axon to open and more positive ions to enter the cell. In the axon, this “signal” will
become a self-propagating brief reversal of the resting membrane potential called an action
potential. An action potential is an all-or-nothing event; it either happens or it does not. The
threshold of excitation must be reached for the neuron to “fire” an action potential. It
travels in only one direction because the sodium channels have been inactivated and
unavailable until the membrane potential is near the resting potential again; at this point
they are reset to closed and can be opened again.
An axon is a tube-like structure that propagates the signal from the cell body to specialized
endings called axon terminals. These terminals in turn then synapse with other neurons,
muscle, or target organs. When the action potential reaches the axon terminal, this causes
the release of neurotransmitter onto the dendrite of another neuron. Neurotransmitters
released at axon terminals allow signals to be communicated to these other cells, and the
process begins again. Neurons usually have one or two axons, but some neurons do not
contain any axons.

Figure 10.8 Neurons contain organelles common to other cells, such as a nucleus and mitochondria. They also
have more specialized structures, including dendrites and axons.

Some axons are covered with a special structure called a myelin sheath, which acts as an
insulator to keep the electrical signal from dissipating as it travels down the axon. This
insulation is important, as the axon from a human motor neuron can be as long as a meter
(3.2 ft)—from the base of the spine to the toes. The myelin sheath is produced by glial cells.
Along the axon there are periodic gaps in the myelin sheath. These gaps are called nodes of
Ranvier and are sites where the signal is “recharged” as it travels along the axon.
It is important to note that a single neuron does not act alone—neuronal communication
depends on the connections that neurons make with one another (as well as with other
cells, like muscle cells). Dendrites from a single neuron may receive synaptic contact from
many other neurons. For example, dendrites from a Purkinje cell in the cerebellum are
thought to receive contact from as many as 200,000 other neurons.
Several diseases can result from the demyelination of axons. The causes of these diseases
are not the same; some have genetic causes, some are caused by pathogens, and others are
the result of autoimmune disorders. Though the causes are varied, the results are largely
similar. The myelin insulation of axons is compromised, making electrical signalling slower.
Multiple sclerosis (MS) is one such disease. It is an example of an autoimmune disease. The
antibodies produced by lymphocytes (a type of white blood cell) mark myelin as something
that should not be in the body. This causes inflammation and the destruction of the myelin
in the central nervous system. As the insulation around the axons is destroyed by the
disease, scarring becomes obvious. This is where the name of the disease comes from;
sclerosis means hardening of tissue, which is what a scar is. Multiple scars are found in the
white matter of the brain and spinal cord. The symptoms of MS include both somatic and
autonomic deficits. Control of the musculature is compromised, as is control of organs such
as the bladder.
Guillain-Barré (pronounced gee-YAN bah-RAY) syndrome is an example of a demyelinating
disease of the peripheral nervous system. It is also the result of an autoimmune reaction,
but the inflammation is in peripheral nerves. Sensory symptoms or motor deficits are
common, and autonomic failures can lead to changes in the heart rhythm or a drop in blood
pressure, especially when standing, which causes dizziness.
While glial cells are often thought of as the supporting cast of the nervous system, the
number of glial cells in the brain actually outnumbers the number of neurons by a factor of
10. Neurons would be unable to function without the vital roles that are fulfilled by these
glial cells. Glia guide developing neurons to their destinations, buffer ions and chemicals
that would otherwise harm neurons, and provide myelin sheaths around axons. When glia
do not function properly, the result can be disastrous—most brain tumors are caused by
mutations in glia.

How Neurons Communicate


All functions performed by the nervous system—from a simple motor reflex to more
advanced functions like making a memory or a decision—require neurons to communicate
with one another. Neurons communicate between the axon of one neuron and the
dendrites, and sometimes the cell body, of another neuron across the gap between them,
known as the synaptic cleft. When an action potential reaches the end of an axon it
stimulates the release of neurotransmitter molecules into the synaptic cleft between the
synaptic knob of the axon and the post-synaptic membrane of the dendrite or soma of the
next cell (Figure 10.9). The neurotransmitter is released through exocytosis of vesicles
containing the neurotransmitter molecules. The neurotransmitter diffuses across the
synaptic cleft and binds to receptors in the post-synaptic membrane. These receptor
molecules are chemically regulated ion channels and will open, allowing sodium to enter the
cell. If sufficient neurotransmitter has been released an action potential may be initiated in
the next cell, but this is not guaranteed. If insufficient neurotransmitter is released the
nerve signal will die at this point. There are a number of different neurotransmitters that are
specific to neuron types that have specific functions.
Figure 10.9. The axon terminal is the distal region of the axon, and contains vesicles filled with
neurotransmitter.

The Central Nervous System


The central nervous system (CNS) is made up of the brain and spinal cord and is covered
with three layers of protective coverings called meninges (“meninges” is derived from the
Greek and means “membranes”) (Figure 10.10). The outermost layer is the dura mater, the
middle layer is the web-like arachnoid mater, and the inner layer is the pia mater, which
directly contacts and covers the brain and spinal cord. The space between the arachnoid
and pia maters is filled with cerebrospinal fluid (CSF). The brain floats in CSF, which acts as a
cushion and shock absorber.
Figure 10.10 The cerebral cortex is covered by three layers of meninges: the dura, arachnoid, and pia maters.
(credit: modification of work by Gray's Anatomy)

The Brain
The brain is the part of the central nervous system that is contained in the cranial cavity of
the skull. It includes the cerebral cortex, limbic system, basal ganglia, thalamus,
hypothalamus, cerebellum, brainstem, and retinas. The outermost part of the brain is a
thick piece of nervous system tissue called the cerebral cortex. The cerebral cortex, limbic
system, and basal ganglia make up the two cerebral hemispheres. A thick fiber bundle called
the corpus callosum (corpus = “body”; callosum = “tough”) connects the two hemispheres.
Although there are some brain functions that are localized more to one hemisphere than
the other, the functions of the two hemispheres are largely redundant. In fact, sometimes
(very rarely) an entire hemisphere is removed to treat severe epilepsy. While patients do
suffer some deficits following the surgery, they can have surprisingly few problems,
especially when the surgery is performed on children who have very immature nervous
systems.
In other surgeries to treat severe epilepsy, the corpus callosum is cut instead of removing an
entire hemisphere. This causes a condition called split-brain, which gives insights into
unique functions of the two hemispheres. For example, when an object is presented to
patients’ left visual field, they may be unable to verbally name the object (and may claim to
not have seen an object at all). This is because the visual input from the left visual field
crosses and enters the right hemisphere and cannot then signal to the speech center, which
generally is found in the left side of the brain. Remarkably, if a split-brain patient is asked to
pick up a specific object out of a group of objects with the left hand, the patient will be able
to do so but will still be unable to verbally identify it.
Visit the following website (http://openstaxcollege.org/l/split-brain2) to learn more about
split-brain patients and to play a game where you can model split-brain experiments
yourself.
Each hemisphere contains regions called lobes that are involved in different functions. Each
hemisphere of the mammalian cerebral cortex can be broken down into four functionally
and spatially defined lobes: frontal, parietal, temporal, and occipital (Figure 10.11).

Figure 10.11 The human cerebral cortex includes the frontal, parietal, temporal, and occipital lobes.
The frontal lobe is located at the front of the brain, over the eyes. This lobe contains the
olfactory bulb, which processes smells. The frontal lobe also contains the motor cortex,
which is important for planning and implementing movement. Areas within the motor
cortex map to different muscle groups. Neurons in the frontal lobe also control cognitive
functions like maintaining attention, speech, and decision-making. Studies of humans who
have damaged their frontal lobes show that parts of this area are involved in personality,
socialization, and assessing risk. The parietal lobe is located at the top of the brain. Neurons
in the parietal lobe are involved in speech and also reading. Two of the parietal lobe’s main
functions are processing somatosensation—touch sensations like pressure, pain, heat,
cold—and processing proprioception—the sense of how parts of the body are oriented in
space. The parietal lobe contains a somatosensory map of the body similar to the motor
cortex. The occipital lobe is located at the back of the brain. It is primarily involved in
vision—seeing, recognizing, and identifying the visual world. The temporal lobe is located at
the base of the brain and is primarily involved in processing and interpreting sounds. It also
contains the hippocampus (named from the Greek for “seahorse,” which it resembles in
shape) a structure that processes memory formation. The role of the hippocampus in
memory was partially determined by studying one famous epileptic patient, HM, who had
both sides of his hippocampus removed in an attempt to cure his epilepsy. His seizures went
away, but he could no longer form new memories (although he could remember some facts
from before his surgery and could learn new motor tasks).
Interconnected brain areas called the basal ganglia play important roles in movement
control and posture. The basal ganglia also regulate motivation.
The thalamus acts as a gateway to and from the cortex. It receives sensory and motor inputs
from the body and also receives feedback from the cortex. This feedback mechanism can
modulate conscious awareness of sensory and motor inputs depending on the attention and
arousal state of the animal. The thalamus helps regulate consciousness, arousal, and sleep
states.
Below the thalamus is the hypothalamus. The hypothalamus controls the endocrine system
by sending signals to the pituitary gland. Among other functions, the hypothalamus is the
body’s thermostat—it makes sure the body temperature is kept at appropriate levels.
Neurons within the hypothalamus also regulate circadian rhythms, sometimes called sleep
cycles.
The limbic system is a connected set of structures that regulates emotion, as well as
behaviors related to fear and motivation. It plays a role in memory formation and includes
parts of the thalamus and hypothalamus as well as the hippocampus. One important
structure within the limbic system is a temporal lobe structure called the amygdala. The
two amygdala (one on each side) are important both for the sensation of fear and for
recognizing fearful faces.
The cerebellum (cerebellum = “little brain”) sits at the base of the brain on top of the
brainstem. The cerebellum controls balance and aids in coordinating movement and
learning new motor tasks. The cerebellum of birds is large compared to other vertebrates
because of the coordination required by flight.
The brainstem connects the rest of the brain with the spinal cord and regulates some of the
most important and basic functions of the nervous system including breathing, swallowing,
digestion, sleeping, walking, and sensory and motor information integration.

Spinal cord
Connecting to the brainstem and extending down the body through the spinal column is the
spinal cord. The spinal cord is a thick bundle of nerve tissue that carries information about
the body to the brain and from the brain to the body. The spinal cord is contained within the
meninges and the bones of the vertebral column but is able to communicate signals to and
from the body through its connections with spinal nerves (part of the peripheral nervous
system). A cross-section of the spinal cord looks like a white oval containing a gray butterfly-
shape (Figure 10.12). Axons make up the “white matter” and neuron and glia cell bodies
(and interneurons) make up the “gray matter.” Axons and cell bodies in the dorsa spinal
cord convey mostly sensory information from the body to the brain. Axons and cell bodies in
the spinal cord primarily transmit signals controlling movement from the brain to the body.
The spinal cord also controls motor reflexes. These reflexes are quick, unconscious
movements—like automatically removing a hand from a hot object. Reflexes are so fast
because they involve local synaptic connections. For example, the knee reflex that a doctor
tests during a routine physical is controlled by a single synapse between a sensory neuron
and a motor neuron. While a reflex may only require the involvement of one or two
synapses, synapses with interneurons in the spinal column transmit information to the brain
to convey what happened (the knee jerked, or the hand was hot).

Figure 10.12 A cross-section of the spinal cord shows gray matter (containing cell bodies and interneurons) and
white matter (containing myelinated axons).
The Peripheral Nervous System
The peripheral nervous system (PNS) is the connection between the central nervous system
and the rest of the body. The PNS can be broken down into the autonomic nervous system,
which controls bodily functions without conscious control, and the sensory-somatic nervous
system, which transmits sensory information from the skin, muscles, and sensory organs to
the CNS and sends motor commands from the CNS to the muscles.

Figure 10.13 In the autonomic nervous system, a preganglionic neuron (originating in the CNS) synapses to a
neuron in a ganglion that, in turn, synapses on a target organ. Activation of the sympathetic nervous system
causes release of norepinephrine on the target organ.
Activation of the parasympathetic nervous system causes release of acetylcholine on the
target organ.
The autonomic nervous system serves as the relay between the CNS and the internal
organs. It controls the lungs, the heart, smooth muscle, and exocrine and endocrine glands.
The autonomic nervous system controls these organs largely without conscious control; it
can continuously monitor the conditions of these different systems and implement changes
as needed. Signaling to the target tissue usually involves two synapses: a preganglionic
neuron (originating in the CNS) synapses to a neuron in a ganglion that, in turn, synapses on
the target organ (Figure 10.13). There are two divisions of the autonomic nervous system
that often have opposing effects: the sympathetic nervous system and the parasympathetic
nervous system.
The sympathetic nervous system is responsible for the immediate responses an animal
makes when it encounters a dangerous situation. One way to remember this is to think of
the “fight-or-flight” response a person feels when encountering a snake (“snake” and
“sympathetic” both begin with “s”). Examples of functions controlled by the sympathetic
nervous system include an accelerated heart rate and inhibited digestion. These functions

Figure 10.14 The sympathetic and parasympathetic nervous systems often have opposing effects on target
organs.
help prepare an organism’s body for the physical strain required to escape a potentially
dangerous situation or to fend off a predator. While the sympathetic nervous system is
activated in stressful situations, the parasympathetic nervous system allows an animal to
“rest and digest.” One way to remember this is to think that during a restful situation like a
picnic, the parasympathetic nervous system is in control (“picnic” and “parasympathetic”
both start with “p”). Parasympathetic preganglionic neurons have cell bodies located in the
brainstem and in the sacral (toward the bottom) spinal cord (Figure 10.14). The
parasympathetic nervous system resets organ function after the sympathetic nervous
system is activated including slowing of heart rate, lowered blood pressure, and stimulation
of digestion.
The sensory-somatic nervous system is made up of cranial and spinal nerves and contains
both sensory and motor neurons. Sensory neurons transmit sensory information from the
skin, skeletal muscle, and sensory organs to the CNS. Motor neurons transmit messages
about desired movement from the CNS to the muscles to make them contract. Without its
sensory-somatic nervous system, an animal would be unable to process any information
about its environment (what it sees, feels, hears, and so on) and could not control motor
movements. Unlike the autonomic nervous system, which usually has two synapses
between the CNS and the target organ, sensory and motor neurons usually have only one
synapse—one ending of the neuron is at the organ and the other directly contacts a CNS
neuron.

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. 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 (Figure
10.15), the cell membrane of the thermoreceptors changes its electrical state (voltage). The
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 signalling 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 signalling molecule called
a neurotransmitter.

Figure 10.15 The Sensory Input Receptors in the skin sense the temperature of the water.

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 (Figure 10.16).
Figure 10.16 The Motor Response On the basis of the sensory input and the integration in the CNS, a
motor response is formulated and executed.

A region of the cortex is specialised for sending signals down to the spinal cord for
movement. The upper motor neuron is in this region that extends all the way down the
spinal cord where its 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 fibres 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 fibres at the neuromuscular junction. Acetylcholine is released at this
specialised 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 fibre, it contracts. All of this occurs in a fraction of a second, but this story is the
basis of how the nervous system functions.

Neurophysiologist
Understanding how the nervous system works could be a driving force in your career.
Studying neurophysiology is a very rewarding path to follow. It means that there is a lot of
work to do, but the rewards are worth the effort.
The career path of a research scientist can be straightforward: college, graduate school,
postdoctoral research, academic research position at a university. A Bachelor’s degree in
science will get you started, and for neurophysiology that might be in biology, psychology,
computer science, engineering, or neuroscience. But the real specialisation comes in
graduate school. There are many different programs out there to study the nervous system,
not just neuroscience itself. Most graduate programs are doctoral, meaning that a Master’s
degree is not part of the work. These are usually considered five-year programs, with the
first two years dedicated to course work and finding a research mentor, and the last three
years dedicated to finding a research topic and pursuing that with a near single-mindedness.
The research will usually result in a few publications in scientific journals, which will make
up the bulk of a doctoral dissertation. After graduating with a Ph.D., researchers will go on
to find specialised work called a postdoctoral fellowship within established labs. In this
position, a researcher starts to establish their own research career with the hopes of finding
an academic position at a research university.
Other options are available if you are interested in how the nervous system works.
Especially for neurophysiology, a medical degree might be more suitable so you can learn
about the clinical applications of neurophysiology and possibly work with human subjects.
An academic career is not a necessity. Biotechnology firms are eager to find
motivated scientists ready to tackle the tough questions about how the nervous system
works so that therapeutic chemicals can be tested on some of the most challenging
disorders such as Alzheimer’s disease or Parkinson’s disease, or spinal cord injury.
Others with a medical degree and a specialization in neuroscience go on to work directly
with patients, diagnosing and treating mental disorders. You can do this as a psychiatrist, a
neuropsychologist, a neuroscience nurse, or a neurodiagnostic technician, among other
possible career paths.
KEY TERMS NERVOUS SYSTEM
action potential change in voltage of a cell membrane in response to a stimulus that results
in transmission of an electrical signal; unique to neurons and muscle fibres
autonomic nervous system (ANS) functional division of the nervous system that is
responsible for homeostatic reflexes that coordinate control of cardiac and smooth muscle,
as well as glandular tissue
axon single process of the neuron that carries an electrical signal (action potential) away
from the cell body toward a target cell
blood-brain barrier (BBB) physiological barrier between the circulatory system and the
central nervous system that establishes a privileged blood supply, restricting the flow of
substances into the CNS
brain the large organ of the central nervous system composed of white and gray matter,
contained within the cranium and continuous with the spinal cord
central nervous system (CNS) anatomical division of the nervous system located within the
cranial and vertebral cavities, namely the brain and spinal cord
cerebral cortex outermost layer of gray matter in the brain, where conscious perception
takes place
cerebrospinal fluid (CSF) circulatory medium within the CNS that is produced by ependymal
cells in the choroid plexus filtering the blood
chemical synapse connection between two neurons, or between a neuron and its target,
where a neurotransmitter diffuses across a very short distance
dendrite one of many branchlike processes that extends from the neuron cell body and
functions as a contact for incoming signals (synapses) from other neurons or sensory cells
depolarisation change in a cell membrane potential from rest toward zero
excitable membrane cell membrane that regulates the movement of ions so that an
electrical signal can be generated
frontal lobe the part of the cerebral cortex that contains the motor cortex and areas
involved in planning, attention, and language
glial cell one of the various types of neural tissue cells responsible for maintenance of the
tissue, and largely responsible for supporting neurons
gray matter regions of the nervous system containing cell bodies of neurons with few or no
myelinated axons; actually may be more pink or tan in colour, but called gray in contrast to
white matter
membrane potential distribution of charge across the cell membrane, based on the charges
of ions
myelin lipid-rich insulating substance surrounding the axons of many neurons, allowing for
faster transmission of electrical signals
neuron neural tissue cell that is primarily responsible for generating and propagating
electrical signals into, within, and out of the nervous system
neurotransmitter chemical signal that is released from the synaptic end bulb of a neuron to
cause a change in the target cell
occipital lobe the part of the cerebral cortex that contains visual cortex and processes
visual stimuli
parietal lobe the part of the cerebral cortex involved in processing touch and the sense of
the body in space
peripheral nervous system (PNS) anatomical division of the nervous system that is largely
outside the cranial and vertebral cavities, namely all parts except the brain and spinal cord
repolarization return of the membrane potential to its normally negative voltage at the end
of the action potential
resting membrane potential the difference in voltage measured across a cell membrane
under steady-state conditions, typically -70 mV
sensation nervous system function that receives information from the environment and
translates it into the electrical signals of nervous tissue
somatic nervous system (SNS) functional division of the nervous system that is concerned
with conscious perception, voluntary movement, and skeletal muscle reflexes
spinal cord organ of the central nervous system found within the vertebral cavity and
connected with the periphery through spinal nerves; mediates reflex behaviours
synapse narrow junction across which a chemical signal passes from neuron to the next,
initiating a new electrical signal in the target cell
temporal lobe the part of the cerebral cortex that processes auditory input; parts of the
temporal lobe are involved in speech, memory, and emotion processing
voltage-gated channel ion channel that opens because of a change in the charge distributed
across the membrane where it is located
white matter regions of the nervous system containing mostly myelinated axons, making
the tissue appear white because of the high lipid content of myelin
10.5 The Endocrine System
Summary
Communication is a process in which a sender transmits signals to one or more receivers to
control and coordinate actions. In the human body, two major organ systems participate in
relatively “long distance” communication: the nervous system and the endocrine system.
Together, these two systems are primarily responsible for maintaining homeostasis in the
body.
The nervous system uses two types of intercellular communication—electrical and chemical
signaling—either by the direct action of an electrical potential, or in the latter case, through
the action of chemical neurotransmitters such as serotonin or norepinephrine.
Neurotransmitters act locally and rapidly.
In contrast, the endocrine system uses just one method of communication:
chemical signaling. These signals are sent by the endocrine organs, which secrete
chemicals—the hormone—into the extracellular fluid. Hormones are transported primarily
via the bloodstream throughout the body, where they bind to receptors on target cells,
inducing a characteristic response. The number of receptors on a target cell can increase or
decrease in response to hormone activity.
Hormone levels are primarily controlled through negative feedback, in which rising levels of
a hormone inhibit its further release. The endocrine system produces hormones that
function to control and regulate many different body processes. Many of the hormones are
secreted in response to signals from the nervous system, thus the two systems act in
concert to effect changes in the body.

Table 10.1: Endocrine and Nervous Systems


Hormones
Maintaining homeostasis within the body requires the coordination of many different
systems and organs. One mechanism of communication between neighbouring cells, and
between cells and tissues in distant parts of the body, occurs through the release of
chemicals called hormones. Hormones are released into body fluids, usually blood, which
carries them to their target cells where they elicit a response. The cells that secrete
hormones are often located in specific organs, called endocrine glands, and the cells,
tissues, and organs that secrete hormones make up the endocrine system. Examples of
endocrine organs include the pancreas, which produces the hormones insulin and glucagon
to regulate blood-glucose levels, the adrenal glands, which produce hormones such as
epinephrine and norepinephrine that regulate responses to stress, and the thyroid gland,
which produces thyroid hormones that regulate metabolic rates.
The endocrine glands differ from the exocrine glands. Exocrine glands secrete chemicals
through ducts that lead outside the gland (not to the blood). For example, sweat produced
by sweat glands is released into ducts that carry sweat to the surface of the skin. The
pancreas has both endocrine and exocrine functions because besides releasing hormones
into the blood. It also produces digestive juices, which are carried by ducts into the small
intestine.

How Hormones Work


Hormones cause changes in target cells by binding to specific cell-surface or intracellular
hormone receptors, molecules embedded in the cell membrane or floating in the cytoplasm
with a binding site that matches a binding site on the hormone molecule. In this way, even
though hormones circulate throughout the body and come into contact with many different
cell types, they only affect cells that possess the necessary receptors. Receptors for a
specific hormone may be found on or in many different cells or may be limited to a small
number of specialized cells. For example, thyroid hormones act on many different tissue
types, stimulating metabolic activity throughout the body. Cells can have many receptors for
the same hormone but often also possess receptors for different types of hormones. The
number of receptors that respond to a hormone determines the cell’s sensitivity to that
hormone, and the resulting cellular response. Additionally, the number of receptors
available to respond to a hormone can change over time, resulting in increased or
decreased cell sensitivity. In up-regulation, the number of receptors increases in response
to rising hormone levels, making the cell more sensitive to the hormone and allowing for
more cellular activity. When the number of receptors decreases in response to rising
hormone levels, called down-regulation, cellular activity is reduced. This process allows cells
to be more sensitive to the hormone that is present. Cells can also alter the sensitivity of the
receptors themselves to various hormones.
Visit this link (https://www.dnatube.com/video/5707/Epinephrine-action) to watch an
animation of the events that occur when a hormone binds to a cell membrane receptor.
Regulation of Hormone Production
To prevent abnormal hormone levels and a potential disease state, hormone levels must be
tightly controlled. The body maintains this control by balancing hormone production and
degradation. Feedback loops govern the initiation and maintenance of most hormone
secretion in response to various stimuli.
Hormone production and release are primarily controlled by negative feedback, as
described previously in the discussion on homeostasis. Negative feedback is characterized
by the inhibition of further secretion of a hormone in response to adequate levels of that
hormone. This allows blood levels of the hormone to be regulated within a narrow range.
For example, the anterior pituitary signals the thyroid to release thyroid hormones.
Increasing levels of these hormones in the blood then give feedback to the hypothalamus
and anterior pituitary to inhibit further signaling to the thyroid gland (Figure 10.17).

Figure 10.17 The anterior pituitary stimulates the thyroid gland to release thyroid hormones T3 and T4.
Increasing levels of these hormones in the blood result in feedback to the hypothalamus and anterior pituitary
to inhibit further signaling to the thyroid gland. (credit: modification of work by Mikael Häggström)
Positive feedback loops are characterized by the release of additional hormone in response
to an original hormone release. The release of oxytocin during childbirth is a positive
feedback loop. The initial release of oxytocin begins to signal the uterine muscles to
contract, which pushes the fetus toward the cervix, causing it to stretch. This, in turn, signals
the pituitary gland to release more oxytocin, causing labor contractions to intensify. The
release of oxytocin decreases after the birth of the child.

Endocrine Glands
The endocrine glands secrete hormones into the surrounding interstitial fluid; those
hormones then diffuse into blood and are carried to various organs and tissues within the
body (Figure 10.18). The ductless endocrine glands are not to be confused with the
body’s exocrine system, whose glands release their secretions through ducts. Examples of
exocrine glands include the sebaceous and sweat glands of the skin.

Figure 10.18 Endocrine System Endocrine glands and cells are located throughout the body and play an
important role in homeostasis.
Table 10.2 Endocrine glands and their associated hormones

The hypothalamus–pituitary complex can be thought of as the “command center” of the


endocrine system. This complex secretes several hormones that directly produce responses
in target tissues, as well as hormones that regulate the synthesis and secretion of hormones
of other glands. In addition, the hypothalamus–pituitary complex coordinates the
messages of the endocrine and nervous systems. In many cases, a stimulus received by the
nervous system must pass through the hypothalamus–pituitary complex to be translated
into hormones that can initiate a response.
The hypothalamus is a structure of the diencephalon of the brain located anterior and
inferior to the thalamus (Figure 10.19). It has both neural and endocrine functions,
producing and secreting many hormones. In addition, the hypothalamus is anatomically and
functionally related to the pituitary gland (or hypophysis), a bean-sized organ suspended
from it by a stem called the infundibulum (or pituitary stalk). Visit this link
(http://openstaxcollege.org/l/roleofhypo) to watch an animation showing the role of the
hypothalamus and the pituitary gland.

Figure 10.19 Hypothalamus–Pituitary Complex The hypothalamus region lies inferior and anterior to the
thalamus. It connects to the pituitary gland by the stalk-like infundibulum. The pituitary gland consists of an
anterior and posterior lobe, with each lobe secreting different hormones in response to signals from the
hypothalamus.

The pituitary gland, sometimes called the hypophysis, is located at the base of the
brain (Figure 10.19). It is attached to the hypothalamus. The posterior pituitary gland is
significantly different in structure from the anterior pituitary, it is a part of the brain,
extending down from the hypothalamus, and contains mostly nerve fibers that extend from
the hypothalamus to the posterior pituitary which stores and releases the hormones
oxytocin and antidiuretic hormone (ADH) that were produced by the hypothalamus.

Oxytocin
When fetal development is complete, the peptide-derived hormone oxytocin (tocia- =
“childbirth”) stimulates uterine contractions and dilation of the cervix. Throughout most of
pregnancy, oxytocin hormone receptors are not expressed at high levels in the uterus.
Toward the end of pregnancy, the synthesis of oxytocin receptors in the uterus increases,
and the smooth muscle cells of the uterus become more sensitive to its effects. Oxytocin is
continually released throughout childbirth through a positive feedback mechanism.

Antidiuretic Hormone (ADH)


In response to high blood osmolarity, which can occur during dehydration or following a
very salty meal, the osmoreceptors signal the posterior pituitary to release antidiuretic
hormone (ADH). The target cells of ADH are located in the tubular cells of the kidneys. Its
effect is to increase epithelial permeability to water, allowing increased water reabsorption.
The more water reabsorbed from the filtrate, the greater the amount of water that is
returned to the blood and the less that is excreted in the urine.
The anterior pituitary gland responds to hormones produced by the hypothalamus by
producing its own hormones, most of which regulate other hormone-producing glands.
The anterior pituitary produces six hormones: growth hormone, prolactin, thyroid-
stimulating hormone, adrenocorticotropic hormone, follicle-stimulating hormone, and
luteinizing hormone. Growth hormone stimulates cellular activities like protein synthesis
that promote growth. Prolactin stimulates the production of milk by the mammary glands.
The other hormones produced by the anterior pituitary regulate the production of
hormones by other endocrine tissues (Table 10.2).

Growth Hormone
The endocrine system regulates the growth of the human body, protein synthesis, and
cellular replication. A major hormone involved in this process is growth hormone (GH), also
called somatotropin—a protein hormone produced and secreted by the anterior pituitary
gland. Its primary function is anabolic; it promotes protein synthesis and tissue building
through direct and indirect mechanisms. GH levels are controlled by the release of GHRH
and GHIH (also known as somatostatin) from the hypothalamus.
GH indirectly mediates growth and protein synthesis by triggering the liver and other tissues
to produce a group of proteins called insulin-like growth factors (IGFs). These proteins
enhance cellular proliferation and inhibit apoptosis, or programmed cell death. IGFs
stimulate cells to increase their uptake of amino acids from the blood for protein
synthesis. Skeletal muscle and cartilage cells are particularly sensitive to stimulation from
IGFs.
Dysfunction of the endocrine system’s control of growth can result in several disorders. For
example, gigantism is a disorder in children that is caused by the secretion of abnormally
large amounts of GH, resulting in excessive growth. A similar condition in adults
is acromegaly, a disorder that results in the growth of bones in the face, hands, and feet
in response to excessive levels of GH in individuals who have stopped growing. Abnormally
low levels of GH in children can cause growth impairment—a disorder called pituitary
dwarfism (also known as growth hormone deficiency).

The thyroid gland is located in the neck, just below the larynx and in front of the
trachea (Figure 10.20). It is a butterfly-shaped gland with two lobes that are connected. The
thyroid follicle cells synthesize the hormone thyroxine, which is also known as T4 because it
contains four atoms of iodine, and triiodothyronine, also known as T3 because it contains
three atoms of iodine. T3 and T4 are released by the thyroid in response to thyroid-
stimulating hormone produced by the anterior pituitary, and both T3 and T4 have the effect
of stimulating metabolic activity in the body and increasing energy use. A third hormone,
calcitonin, is also produced by the thyroid. Calcitonin is released in response to rising
calcium ion concentrations in the blood and has the effect of reducing those levels.

Figure 10.20 Thyroid Gland The thyroid gland is located in the neck where it wraps around the trachea. (a)
Anterior view of the thyroid gland. (b) Posterior view of the thyroid gland. (Micrograph provided by the
Regents of University of Michigan Medical School © 2012)
Most people have four parathyroid glands; however, the number can vary from two to six.
These glands are located on the posterior surface of the thyroid gland (Figure 10.20).

Iodine Deficiency, Hypothyroidism, and Hyperthyroidism


As discussed above, dietary iodine is required for the synthesis of T3 and T4. But for much of
the world’s population, foods do not provide adequate levels of this mineral, because the
amount varies according to the level in the soil in which the food was grown, as well as the
irrigation and fertilizers used. Marine fish and shrimp tend to have high levels because they
concentrate iodine from seawater, but many people in landlocked regions lack access to
seafood. Thus, the primary source of dietary iodine in many countries is iodized salt.
Dietary iodine deficiency can result in the impaired ability to synthesize T3 and T4, leading to
a variety of severe disorders. When T3 and T4 cannot be produced, TSH is secreted in
increasing amounts. As a result of this hyperstimulation, thyroglobulin accumulates in the
thyroid gland follicles, increasing their deposits of colloid. The accumulation of colloid
increases the overall size of the thyroid gland, a condition called a goiter (Figure 10.21). A
goiter is only a visible indication of the deficiency. Other iodine deficiency disorders include
impaired growth and development, decreased fertility, and prenatal and infant death.
Moreover, iodine deficiency is the primary cause of preventable mental retardation
worldwide.

Figure 10.21 Goiter (credit: “Almazi”/Wikimedia Commons)

In areas of the world with access to iodized salt, dietary deficiency is rare. Instead,
inflammation of the thyroid gland is the more common cause of low blood levels of thyroid
hormones. Called hypothyroidism, the condition is characterized by a low metabolic rate,
weight gain, cold extremities, constipation, reduced libido, menstrual irregularities, and
reduced mental activity. In contrast, hyperthyroidism—an abnormally elevated blood level
of thyroid hormones—is often caused by a pituitary or thyroid tumor. In Graves’ disease,
the hyperthyroid state results from an autoimmune reaction in which antibodies
overstimulate the follicle cells of the thyroid gland. Hyperthyroidism can lead to an
increased metabolic rate, excessive body heat and sweating, diarrhea, weight loss, tremors,
and increased heart rate. The person’s eyes may bulge (called exophthalmos) as antibodies
produce inflammation in the soft tissues of the orbits. The person may also develop a goiter.
The adrenal glands are located on top of each kidney (Figure 10.22). The adrenal glands
consist of an outer adrenal cortex and an inner adrenal medulla. These regions secrete
different hormones.

Figure 10.22 Adrenal Glands Both adrenal glands sit atop the kidneys and are composed of an outer cortex and
an inner medulla, all surrounded by a connective tissue capsule. The cortex can be subdivided into additional
zones, all of which produce different types of hormones. LM × 204. (Micrograph provided by the Regents of
University of Michigan Medical School © 2012)

The adrenal cortex produces mineralocorticoids, glucocorticoids, and androgens. The main
mineralocorticoid is aldosterone, which regulates the concentration of ions in urine, sweat,
and saliva. Aldosterone release from the adrenal cortex is stimulated by a decrease in blood
concentrations of sodium ions, blood volume, or blood pressure, or by an increase in blood
potassium levels. The glucocorticoids maintain proper blood-glucose levels between meals.
They also control a response to stress by increasing glucose synthesis from fats and proteins
and interact with epinephrine to cause vasoconstriction. Androgens are sex hormones that
are produced in small amounts by the adrenal cortex. They do not normally affect sexual
characteristics and may supplement sex hormones released from the gonads. The adrenal
medulla contains two types of secretory cells: one that produces epinephrine (adrenaline)
and another that produces norepinephrine (noradrenaline). Epinephrine and
norepinephrine cause immediate, short-term changes in response to stressors, inducing the
so-called fight-or-flight response. The responses include increased heart rate, breathing
rate, cardiac muscle contractions, and blood-glucose levels. They also accelerate the
breakdown of glucose in skeletal muscles and stored fats in adipose tissue, and redirect
blood flow toward skeletal muscles and away from skin and viscera. The release of
epinephrine and norepinephrine is stimulated by neural impulses from the sympathetic
nervous system that originate from the hypothalamus.

Table 10.3: Hormones of the adrenal glands

The pancreas is an elongate organ located between the stomach and the proximal portion
of the small intestine (Figure 10.23). It contains both exocrine cells that excrete digestive
enzymes and endocrine cells that release hormones.

Figure 10.23 Pancreas endocrine function involves the secretion of insulin (produced by beta cells) and
glucagon (produced by alpha cells) within the pancreatic islets. These two hormones regulate the rate of
glucose metabolism in the body. The micrograph reveals pancreatic islets. LM × 760. (Micrograph provided by
the Regents of University of Michigan Medical School © 2012)

View the University of Michigan WebScope


at http://141.214.65.171/Histology/Digestive%20System/ Liver%20and%20Pancreas/188B
_HISTO_40X.svs/view.apml (http://openstaxcollege.org/l/pancreaticislet) to explore the
tissue sample in greater detail.
The endocrine cells of the pancreas form clusters called pancreatic islets or the islets of
Langerhans. Among the cell types in each pancreatic islet are the alpha cells, which produce
the hormone glucagon, and the beta cells, which produce the hormone insulin. These
hormones regulate blood-glucose levels. Alpha cells release glucagon as blood-glucose
levels decline. When blood-glucose levels rise, beta cells release insulin. Glucagon causes
the release of glucose to the blood from the liver, and insulin facilitates the uptake of
glucose by the body’s cells (Figure 10.24).

Figure 10.24 Homeostatic Regulation of Blood Glucose Levels Blood glucose concentration is tightly
maintained between 70 mg/dL and 110 mg/dL. If blood glucose concentration rises above this range, insulin is
released, which stimulates body cells to remove glucose from the blood. If blood glucose concentration drops
below this range, glucagon is released, which stimulates body cells to release glucose into the blood.
Endocrine System: Diabetes Mellitus
Dysfunction of insulin production and secretion, as well as the target cells’ responsiveness
to insulin, can lead to a condition called diabetes mellitus. An increasingly common disease,
diabetes mellitus has been diagnosed in more than 18 million adults in the United States,
and more than 200,000 children. It is estimated that up to 7 million more adults have the
condition but have not been diagnosed. In addition, approximately 79 million people in the
US are estimated to have pre-diabetes, a condition in which blood glucose levels are
abnormally high, but not yet high enough to be classified as diabetes.
There are two main forms of diabetes mellitus. Type 1 diabetes is an autoimmune disease
affecting the beta cells of the pancreas. Certain genes are recognized to increase
susceptibility. The beta cells of people with type 1 diabetes do not produce insulin; thus,
synthetic insulin must be administered by injection or infusion. This form of diabetes
accounts for less than five percent of all diabetes cases.
Type 2 diabetes accounts for approximately 95 percent of all cases. It is acquired, and
lifestyle factors such as poor diet, inactivity, and the presence of pre-diabetes greatly
increase a person’s risk. About 80 to 90 percent of people with type 2 diabetes are
overweight or obese. In type 2 diabetes, cells become resistant to the effects of insulin.
In response, the pancreas increases its insulin secretion, but over time, the beta cells
become exhausted. In many cases, type 2 diabetes can be reversed by moderate weight
loss, regular physical activity, and consumption of a healthy diet; however, if blood glucose
levels cannot be controlled, the diabetic will eventually require insulin.
Two of the early manifestations of diabetes are excessive urination and excessive thirst.
They demonstrate how the out-of-control levels of glucose in the blood affect kidney
function. The kidneys are responsible for filtering glucose from the blood. Excessive blood
glucose draws water into the urine, and as a result the person eliminates an
abnormally large quantity of sweet urine. The use of body water to dilute the urine leaves
the body dehydrated, and so the person is unusually and continually thirsty. The person may
also experience persistent hunger because the body cells are unable to access the glucose in
the bloodstream.
Over time, persistently high levels of glucose in the blood injure tissues throughout the
body, especially those of the blood vessels and nerves. Inflammation and injury of the lining
of arteries lead to atherosclerosis and an increased risk of heart attack and stroke. Damage
to the microscopic blood vessels of the kidney impairs kidney function and can lead to
kidney failure. Damage to blood vessels that serve the eyes can lead to blindness. Blood
vessel damage also reduces circulation to the limbs, whereas nerve damage leads to a loss
of sensation, called neuropathy, particularly in the hands and feet. Together, these changes
increase the risk of injury, infection, and tissue death (necrosis), contributing to a high rate
of toe, foot, and lower leg amputations in people with diabetes. Uncontrolled diabetes can
also lead to a dangerous form of metabolic acidosis called ketoacidosis. Deprived of glucose,
cells increasingly rely on fat stores for fuel. However, in a glucose-deficient state, the liver is
forced to use an alternative lipid metabolism pathway that results in the increased
production of ketone bodies (or ketones), which are acidic. The build-up of ketones in the
blood causes ketoacidosis, which—if left untreated—may lead to a life-threatening “diabetic
coma.” Together, these complications make diabetes the seventh leading cause of death in
the United States.
Diabetes is diagnosed when lab tests reveal that blood glucose levels are higher than
normal, a condition called hyperglycemia. The treatment of diabetes depends on the type,
the severity of the condition, and the ability of the patient to make lifestyle changes. As
noted earlier, moderate weight loss, regular physical activity, and consumption of a
healthful diet can reduce blood glucose levels. Some patients with type 2 diabetes may be
unable to control their disease with these lifestyle changes, and will require medication.
Historically, the first-line treatment of type 2 diabetes was insulin. Research advances have
resulted in alternative options, including medications that enhance pancreatic function.
Visit this link (http://openstaxcollege.org/l/insulin) to view an animation describing the
role of insulin and the pancreas in diabetes.
The gonads—the male testes and female ovaries—produce steroid hormones. The testes
produce androgens, testosterone being the most prominent, which allow for the
development of secondary sex characteristics and the production of sperm cells. The ovaries
produce estrogen and progesterone, which cause secondary sex characteristics, regulate
production of eggs, control pregnancy, and prepare the body for childbirth.

Table 10.4: Reproductive hormones

There are several organs whose primary functions are non-endocrine but that also possess
endocrine functions. These include the heart, kidneys, intestines, thymus, and adipose
tissue. The heart has endocrine cells in the walls of the atria that release a hormone in
response to increased blood volume. It causes a reduction in blood volume and blood
pressure, and reduces the concentration of Na+ in the blood.
The gastrointestinal tract produces several hormones that aid in digestion. The endocrine
cells are located in the mucosa of the GI tract throughout the stomach and small intestine.
They trigger the release of gastric juices, which help to break down and digest food in the GI
tract.
The kidneys also possess endocrine function. Two of these hormones regulate ion
concentrations and blood volume or pressure. Erythropoietin (EPO) is released by kidneys in
response to low oxygen levels. EPO triggers the formation of red blood cells in the bone
marrow. EPO has been used by athletes to improve performance. But EPO doping has its
risks, since it thickens the blood and increases strain on the heart; it also increases the risk
of blood clots and therefore heart attacks and stroke.

Endocrinologist
An endocrinologist is a medical doctor who specializes in treating endocrine disorders. An
endocrine surgeon specializes in the surgical treatment of endocrine diseases and glands.
Some of the diseases that are managed by endocrinologists include disorders of the
pancreas (diabetes mellitus), disorders of the pituitary (gigantism, acromegaly, and pituitary
dwarfism), disorders of the thyroid gland (goiter and Graves’ disease), and disorders of the
adrenal glands (Cushing’s disease and Addison’s disease).
Endocrinologists are required to assess patients and diagnose endocrine disorders through
extensive use of laboratory tests. Many endocrine diseases are diagnosed using tests that
stimulate or suppress endocrine organ functioning. Blood samples are then drawn to
determine the effect of stimulating or suppressing an endocrine organ on the production of
hormones. For example, to diagnose diabetes mellitus, patients are required to fast for 12
to 24 hours. They are then given a sugary drink, which stimulates the pancreas to produce
insulin to decrease blood-glucose levels. A blood sample is taken one to two hours after the
sugar drink is consumed. If the pancreas is functioning properly, the blood-glucose level will
be within a normal range. Another example is the A1C test, which can be performed during
blood screening. The A1C test measures average blood-glucose levels over the past two to
three months. The A1C test is an indicator of how well blood glucose is being managed over
a long time.
Once a disease such as diabetes has been diagnosed, endocrinologists can prescribe lifestyle
changes and medications to treat the disease. Some cases of diabetes mellitus can be
managed by exercise, weight loss, and a healthy diet; in other cases, medications may be
required to enhance insulin’s production or effect. If the disease cannot be controlled by
these means, the endocrinologist may prescribe insulin injections.
In addition to clinical practice, endocrinologists may also be involved in primary research
and development activities. For example, ongoing islet transplant research is investigating
how healthy pancreas islet cells may be transplanted into diabetic patients. Successful islet
transplants may allow patients to stop taking insulin injections.
KEY TERMS ENDOCRINE
acromegaly disorder in adults caused when abnormally high levels of GH trigger growth of
bones in the face, hands, and feet
adrenal glands endocrine glands located at the top of each kidney that are important for
the regulation of the stress response, blood pressure and blood volume, water homeostasis,
and electrolyte levels
aldosterone hormone produced and secreted by the adrenal cortex that stimulates sodium
and fluid retention and increases blood volume and blood pressure
antidiuretic hormone (ADH) hypothalamic hormone that is stored by the posterior pituitary
and that signals the kidneys to reabsorb water
autocrine chemical signal that elicits a response in the same cell that secreted it
beta cell pancreatic islet cell type that produces the hormone insulin
calcitonin peptide hormone produced and secreted by the parafollicular cells (C cells) of the
thyroid gland that functions to decrease blood calcium levels
cortisol glucocorticoid important in gluconeogenesis, the catabolism of glycogen, and
downregulation of the immune system
cyclic adenosine monophosphate (cAMP) second messenger that, in response to adenylyl
cyclase activation, triggers a phosphorylation cascade
diabetes mellitus condition caused by destruction or dysfunction of the beta cells of the
pancreas or cellular resistance to insulin that results in abnormally high blood glucose levels
downregulation decrease in the number of hormone receptors, typically in response to
chronically excessive levels of a hormone
endocrine gland tissue or organ that secretes hormones into the blood and lymph without
ducts such that they may be transported to organs distant from the site of secretion
endocrine system cells, tissues, and organs that secrete hormones as a primary or
secondary function and play an integral role in normal bodily processes
epinephrine primary and most potent catecholamine hormone secreted by the adrenal
medulla in response to short term stress; also called adrenaline
exocrine system cells, tissues, and organs that secrete substances directly to target tissues
via glandular ducts
first messenger hormone that binds to a cell membrane hormone receptor and triggers
activation of a second messenger system
gigantism disorder in children caused when abnormally high levels of GH prompt excessive
growth
glucagon pancreatic hormone that stimulates the catabolism of glycogen to glucose,
thereby increasing blood glucose levels
goiter enlargement of the thyroid gland either as a result of iodine deficiency or
hyperthyroidism
growth hormone (GH) anterior pituitary hormone that promotes tissue building and
influences nutrient metabolism (also called somatotropin)
hormone secretion of an endocrine organ that travels via the bloodstream or lymphatics to
induce a response in target cells or tissues in another part of the body
hormone receptor protein within a cell or on the cell membrane that binds a hormone,
initiating the target cell response
hyperglycemia abnormally high blood glucose levels
hyperthyroidism clinically abnormal, elevated level of thyroid hormone in the blood;
characterized by an increased metabolic rate, excess body heat, sweating, diarrhea, weight
loss, and increased heart rate
hypothalamus region of the diencephalon inferior to the thalamus that functions in neural
and endocrine signaling
hypothyroidism clinically abnormal, low level of thyroid hormone in the blood;
characterized by low metabolic rate, weight gain, cold extremities, constipation, and
reduced mental activity
insulin pancreatic hormone that enhances the cellular uptake and utilization of glucose,
thereby decreasing blood glucose levels
norepinephrine secondary catecholamine hormone secreted by the adrenal medulla in
response to short-term stress; also called noradrenaline
pancreas organ with both exocrine and endocrine functions located posterior to the
stomach that is important for digestion and the regulation of blood glucose
pancreatic islets specialized clusters of pancreatic cells that have endocrine functions; also
called islets of Langerhans
paracrine chemical signal that elicits a response in neighboring cells; also called paracrine
factor
parathyroid glands small, round glands embedded in the posterior thyroid gland that
produce parathyroid hormone (PTH)
pituitary dwarfism disorder in children caused when abnormally low levels of GH result in
growth retardation
pituitary gland bean-sized organ suspended from the hypothalamus that produces, stores,
and secretes hormones in response to hypothalamic stimulation (also called hypophysis)
second messenger molecule that initiates a signaling cascade in response to hormone
binding on a cell membrane receptor and activation of a G protein
thyroid gland large endocrine gland responsible for the synthesis of thyroid hormones
thyroxine (also, tetraiodothyronine, T4) amino acid–derived thyroid hormone that is more
abundant but less potent than T3 and often converted to T3 by target cells
upregulation increase in the number of hormone receptors, typically in response to
chronically reduced levels of a hormone
REVISION QUESTIONS
Question 10.1
Cells may communicate with one another by __________.

A locally acting chemicals, called paracrines, autocrines or neuromodulators


B long-distance means, which rely on combinations of electrical and chemical signals
C A and B

Question 10.2
The difference between an endocrine and an exocrine is __________.

A the cell that releases it


B the cell that responds to it
C the method of transport
D the route of transport
E there are no differences they are the same

Question 10.3
Receptor molecules for chemical signaling are located

A in the membrane
B in the cytosol
C in the nucleus
D a, b & c

Question 10.4
Which of the following types of cells display the property of electrical excitability?
A Muscle cells.
B Neurons.
C Both muscle and neurons.
D None of these choices.
Question 10.5
In an action potential, the current that flows down the axon of a neuron is generated by the
movement of _____ across the membrane.

A electrons
B protons
C ions
D free radicals
E neutrons

Question 10.6
The resting membrane potential (RMP) of a typical nerve cell is

A +50 mV
B +70 mV
C0
D -50 mV
E -70 mV

Question 10.7
Which of the following types of electrical signals allow rapid long-distance communication
within the nervous system?

A Resting membrane potential


B Nerve action potential
C Muscle action potential
D Graded potential

The nervous system


Question 10.8
Which of the following is NOT a function of the nervous system?

A Sensory function
B Integrative function
C Response (Motor) function
D All are functions of the nervous system
Question 10.9
This part of a neuron receives information from other neurons

AC
BD
CE
DA
EG
Question 10.10
Which portions of the spinal cord below represent the gray matter?

A B, D, I
B C, F, H
CA
DJ
EG

Question 10.11
Which of the following lists the components of a reflex arc in the proper sequence of
activation?

A Receptor - sensory neuron - integrating centre - motor neuron – effector.


B Receptor - motor neuron - integrating centre - sensory neuron - effector.
C Effector - sensory neuron - integrating centre - motor neuron - receptor.
D Integrating centre - receptor - sensory neuron - motor neuron - effector.
E Receptor - sensory neuron - motor neuron - effector - integrating centre.
Question 10.12
Hearing your cell phone ring in an otherwise quiet lecture hall is an example of which of the
following types of nervous system functions?

A Sensory function.
B Integrative function.
C Response (Motor) function.
D More than one of the types of nervous system functions.

Question 10.13
Which type of sensory receptor responds to stimuli resulting from changes in temperature?

A Photoreceptors.
B Mechanoreceptors.
C Proprioceptors.
D. Nociceptors.
E Thermoreceptors.

Question 10.14
Which of the following disorders is the most common brain disorder?

A Alzheimer's disease.
B Cerebrovascular accident (stroke).
C Brain tumour.
D Attention deficit hyperactivity disorder.
Question 10.15
Which of the labelled structures in the diagram points to the front lobe of the cerebral
cortex?

AA
BB
CD
DE
EF

Endocrine

Question 10.16
The endocrine system __________.

A releases chemicals into the bloodstream for distribution throughout the body
B releases hormones that alter the metabolic activities of many different tissues and organs
simultaneously
C produces effects that can last for hours, days and even longer
D functions to control ongoing metabolic processes
E all of the above
Question 10.17
Endocrine cells __________.

A are a type of nerve cell


B release their secretions onto an epithelial surface
C release their secretions into the blood

Question 10.18
The most complex endocrine responses involve the __________.

A thyroid gland
B pancreas
C adrenal glands
D hypothalamus

Question 10.19
When a hormone is present in excessive levels, the number of target-cell receptors may
decrease. This is called:

A receptor recognition.
B sensory adaptation.
C paracrine regulation.
D up-regulation.
E down-regulation.

Question 10.20
Which of the following are a group of lipid-soluble hormones derived from cholesterol?

A Steroids.
B Thyroid hormones.
C Nitric oxide.
D Amine hormones.
E Peptide hormones.
Question 10.21
What controls hormone release from the anterior pituitary gland?

A Muscle contraction.
B The peripheral nervous system.
C Hormones released from the hypothalamus.
D Action potentials from the thalamus.
E Chemical changes in the cerebrospinal fluid (CSF).

Question 10.22
Which of the following endocrine disorders is the individual shown in this picture suffering
from?

A Acromegaly.
B Grave’s Disease.
C Goiter.
D Cushing's syndrome.
E Gigantism.
Question 10.23
Which of the following anterior pituitary hormones stimulates general body growth?

A Growth hormone.
B Prolactin.
C Thyrotropin.
D Luteinising hormone.
E Adrenocorticotropic hormone.

Question 10.24
What endocrine abnormality does the individual on the left in the picture of identical twins
have?

A Goiter.
B Graves' Disease.
C Cushing's syndrome.
D Gigantism.

Question 10.25
Which blood glucose-lowering hormone is produced by the pancreatic islet cells?

A Insulin.
B Glucagon.
D Thyroid hormones.
E Calcitonin.
ANSWERS
Question 10.1
C

Question 10.2
C

Question 10.3
D

Question 10.4
C

Question 10.5
C

Question 10.6
E

Question 10.7
B

Question 10.8
D

Question 10.9
D

Question 10.10
B

Question 10.11
A

Question 10.12
A

Question 10.13
E

Question 10.14
B
Question 10.15
A

Question 10.16
E

Question 10.17
C

Question 10.18
D

Question 10.19
E

Question 10.20
A

Question 10.21
C

Question 10.22
C

Question 10.23
A

Question 10.24
D

Question 10.25
A

You might also like