1
Unraveling the Mystery of How
Geckos Defy Gravity
Science is most fun when it tickles your imagination. This is
particularly true when you see something your common
sense tells you just can’t be true. Imagine, for example, you
are lying on a bed in a tropical hotel room. A little lizard, a
blue gecko about the size of a toothbrush, walks up the wall
beside you and upside down across the ceiling, stopping for
a few moments over your head to look down at you, and
then trots over to the far wall and down.
There is nothing at all unusual in what you have just
imagined. Geckos are famous for strolling up walls in this
fashion. How do geckos perform this gripping feat? Investi-
gators have puzzled over the adhesive properties of geckos
for decades. What force prevents gravity from dropping the
gecko on your nose?
The most reasonable hypothesis seemed suction—
salamanders’ feet form suction cups that let them climb
walls, so maybe geckos’ do too. The way to test this is to see
if the feet adhere in a vacuum, with no air to create suction.
Salamander feet don’t, but gecko feet do. It’s not suction.
How about friction? Cockroaches climb using tiny hooks
that grapple onto irregularities in the surface, much as rock-
climbers use crampons. Geckos, however, happily run up
walls of smooth polished glass that no cockroach can climb.
It’s not friction.
Electrostatic attraction? Clothes in a dryer stick together
because of electrical charges created by their rubbing to-
gether. You can stop this by adding a “static remover” like a
Cling-free sheet that is heavily ionized. But a gecko’s feet
still adhere in ionized air. It’s not electrostatic attraction.
Could it be glue? Many insects use adhesive secretions
from glands in their feet to aid climbing. But there are no
glands cells in the feet of a gecko, no secreted chemicals, no
footprints left behind. It’s not glue.
There is one tantalizing clue, however, the kind that ex-
perimenters love. Gecko feet seem to get stickier on some
surfaces than others. They are less sticky on low-energy
surfaces like Teflon, and more sticky on surfaces made of
polar molecules. This suggests that geckos are tapping
directly into the molecular structure of the surfaces they
walk on!
Tracking down this clue, Kellar Autumn of Lewis &
Clark College in Portland, Oregon, and Robert Full of the
University of California, Berkeley, took a closer look at
gecko feet. Geckos have rows of tiny hairs called setae on
the bottoms of their feet, like the bristles of some trendy
toothbrush. When you look at these hairs under the micro-
scope, the end of each seta is divided into 400 to 1000 fine
projections called spatulae. There are about half a million of
these setae on each foot, each only one-tenth the diameter
of a human hair.
Autumn and Full put together an interdisciplinary team
of scientists and set out to measure the force produced by a
single seta. To do this, they had to overcome two significant
experimental challenges:
Isolating a single seta. No one had ever isolated a single
seta before. They succeeded in doing this by surgically
plucking a hair from a gecko foot under a microscope and
bonding the hair onto a microprobe. The microprobe
was fitted into a specially designed micromanipulator that
can move the mounted hair in various ways.
Measuring a very small force. Previous research had
shown that if you pull on a whole gecko, the adhesive
force sticking each of the gecko’s feet to the wall is about
10 Newtons (N), which is like supporting 1 kg. Because
each foot has half a million setae, this predicts that a sin-
gle seta would produce about 20 microNewtons of force.
That’s a very tiny amount to measure. To attempt the
measurement, Autumn and Full recruited a mechanical
engineer from Stanford, Thomas Kenny. Kenny is an ex-
pert at building instruments that can measure forces at
the atomic level.
Part
I
The Origin of Living
Things
Defying gravity. This gecko lizard is able to climb walls and
walk upside down across ceilings. Learning how geckos do this is
a fascinating bit of experimental science.
Real People Doing Real Science
The Experiment
Once this team had isolated a seta and placed it in Kenny’s
device, “We had a real nasty surprise,” says Autumn. For
two months, pushing individual seta against a surface, they
couldn’t get the isolated hair to stick at all!
This forced the research team to stand back and think a
bit. Finally it hit them. Geckos don’t walk by pushing their
feet down, like we do. Instead, when a gecko takes a step, it
pushes the palm of the foot into the surface, then uncurls
its toes, sliding them backwards onto the surface. This
shoves the forest of tips sideways against the surface.
Going back to their instruments, they repeated their ex-
periment, but this time they oriented the seta to approach
the surface from the side rather than head-on. This had the
effect of bringing the many spatulae on the tip of the seta
into direct contact with the surface.
To measure these forces on the seta from the side, as well
as the perpendicular forces they had already been measur-
ing, the researchers constructed a micro-electromechanical
cantilever. The apparatus consisted of two piezoresistive
layers deposited on a silicon cantilever to detect force in
both parallel and perpendicular angles.
The Results
With the seta oriented properly, the experiment yielded re-
sults. Fantastic results. The attachment force measured by
the machine went up 600-fold from what the team had
been measuring before. A single seta produced not the 20
microNewtons of force predicted by the whole-foot mea-
surements, but up to an astonishing 200 microNewtons
(see graph above)! Measuring many individual seta, adhe-
sive forces averaged 194+25 microNewtons.
Two hundred microNewtons is a tiny force, but stupen-
dous for a single hair only 100 microns long. Enough to hold
up an ant. A million hairs could support a small child. A little
gecko, ceiling walking with 2 million of them (see photos
above), could theoretically carry a 90-pound backpack—talk
about being over-engineered.
If a gecko’s feet stick that good, how do geckos ever
become unstuck? The research team experimented with
unattaching individual seta; they used yet another micro-
instrument, this one designed by engineer Ronald Fearing
also from U.C. Berkeley, to twist the hair in various ways.
They found that tipped past a critical angle, 30 degrees,
the attractive forces between hair and surface atoms
weaken to nothing. The trick is to tip a foot hair until its
projections let go. Geckos release their feet by curling up
each toe and peeling it off, just the way we remove tape.
What is the source of the powerful adhesion of gecko feet?
The experiments do not reveal exactly what the attractive
force is, but it seems almost certain to involve interactions at
the atomic level. For a gecko’s foot to stick, the hundreds of
spatulae at the tip of each seta must butt up squarely against
the surface, so the individual atoms of each spatula can come
into play. When two atoms approach each other very
closely—closer than the diameter of an atom—a subtle nu-
clear attraction called Van der Waals forces comes into play.
These forces are individually very weak, but when lots of
them add their little bits, the sum can add up to quite a lot.
Might robots be devised with feet tipped with artificial
setae, able to walk up walls? Autumn and Full are working
with a robotics company to find out. Sometimes science is
not only fun, but can lead to surprising advances.
To explore this experiment further,
go to the Virtual Lab at
www.mhhe.com/raven6/vlab1.mhtml
12
Time (s)
345
20
0
-20
40
60
Force (
μ
N)
80
0
Begin parallel
pulling
Seta pulled
off sensor
The sliding step experiment. The adhesive force of a single seta
was measured. An initial push perpendicularly put the seta in
contact with the sensor. Then, with parallel pulling, the force
continued to increase over time to a value of 60 microNewtons
(after this, the seta began to slide and pulled off the sensor). In a
large number of similar experiments, adhesion forces typically
approach 200 microNewtons.
Closeup look at a gecko’s foot. The setae on a gecko’s foot are
arranged in rows, and point backwards, away from the toenail.
Each seta branches into several hundred spatulae (inset photo).
3
1
The Science of Biology
Concept Outline
1.1 Biology is the science of life.
Properties of Life. Biology is the science that studies
living organisms and how they interact with one another and
their environment.
1.2 Scientists form generalizations from observations.
The Nature of Science. Science employs both deductive
reasoning and inductive reasoning.
How Science Is Done. Scientists construct hypotheses
from systematically collected objective data. They then
perform experiments designed to disprove the hypotheses.
1.3 Darwin’s theory of evolution illustrates how science
works.
Darwin’s Theory of Evolution. On a round-the-world
voyage Darwin made observations that eventually led him to
formulate the hypothesis of evolution by natural selection.
Darwin’s Evidence. The fossil and geographic patterns of
life he observed convinced Darwin that a process of evolution
had occurred.
Inventing the Theory of Natural Selection. The
Malthus idea that populations cannot grow unchecked led
Darwin, and another naturalist named Wallace, to propose
the hypothesis of natural selection.
Evolution After Darwin: More Evidence. In the century
since Darwin, a mass of experimental evidence has supported
his theory of evolution, now accepted by practically all prac-
ticing biologists.
1.4 This book is organized to help you learn biology.
Core Principles of Biology. The first half of this text is
devoted to general principles that apply to all organisms, the
second half to an examination of particular organisms.
Y
ou are about to embark on a journey—a journey of
discovery about the nature of life. Nearly 180 years
ago, a young English naturalist named Charles Darwin set
sail on a similar journey on board H.M.S. Beagle (figure
1.1 shows a replica of the Beagle). What Darwin learned on
his five-year voyage led directly to his development of the
theory of evolution by natural selection, a theory that has
become the core of the science of biology. Darwin’s voyage
seems a fitting place to begin our exploration of biology,
the scientific study of living organisms and how they have
evolved. Before we begin, however, let’s take a moment to
think about what biology is and why it’s important.
FIGURE 1.1
A replica of the Beagle, off the southern coast of South
America. The famous English naturalist, Charles Darwin,
set forth on H.M.S. Beagle in 1831, at the age of 22.
4 Part I The Origin of Living Things
Properties of Life
In its broadest sense, biology is the study of living things—the
science of life. Living things come in an astounding variety of
shapes and forms, and biologists study life in many differ-
ent ways. They live with gorillas, collect fossils, and listen
to whales. They isolate viruses, grow mushrooms, and ex-
amine the structure of fruit flies. They read the messages
encoded in the long molecules of heredity and count how
many times a hummingbird’s wings beat each second.
What makes something “alive”? Anyone could deduce
that a galloping horse is alive and a car is not, but why? We
cannot say, “If it moves, it’s alive,” because a car can move,
and gelatin can wiggle in a bowl. They certainly are not
alive. What characteristics do define life? All living organ-
isms share five basic characteristics:
1. Order. All organisms consist of one or more cells
with highly ordered structures: atoms make up mole-
cules, which construct cellular organelles, which are
contained within cells. This hierarchical organization
continues at higher levels in multicellular organisms
and among organisms (figure 1.2).
2. Sensitivity. All organisms respond to stimuli. Plants
grow toward a source of light, and your pupils dilate
when you walk into a dark room.
3. Growth, development, and reproduction. All or-
ganisms are capable of growing and reproducing, and
they all possess hereditary molecules that are passed to
their offspring, ensuring that the offspring are of the
same species. Although crystals also “grow,” their
growth does not involve hereditary molecules.
4. Regulation. All organisms have regulatory mecha-
nisms that coordinate the organism’s internal func-
tions. These functions include supplying cells with nu-
trients, transporting substances through the organism,
and many others.
5. Homeostasis. All organisms maintain relatively
constant internal conditions, different from their envi-
ronment, a process called homeostasis.
All living things share certain key characteristics: order,
sensitivity, growth, development and reproduction,
regulation, and homeostasis.
1.1 Biology is the science of life.
FIGURE 1.2
Hierarchical organization of living things. Life is highly orga-
nized—from small and simple to large and complex, within cells,
within multicellular organisms, and among organisms.
Organelle
Macromolecule
Molecule
Cell
WITHIN CELLS
Chapter 1 The Science of Biology 5
AMONG ORGANISMS
Ecosystem
Community
Species
Population
WITHIN MULTICELLULAR ORGANISMS
Tissue
Organ
Organ system
Organism
6 Part I The Origin of Living Things
The Nature of Science
Biology is a fascinating and important subject, because it
dramatically affects our daily lives and our futures. Many
biologists are working on problems that critically affect our
lives, such as the world’s rapidly expanding population and
diseases like cancer and AIDS. The knowledge these biolo-
gists gain will be fundamental to our ability to manage the
world’s resources in a suitable manner, to prevent or cure
diseases, and to improve the quality of our lives and those
of our children and grandchildren.
Biology is one of the most successful of the “natural sci-
ences,” explaining what our world is like. To understand
biology, you must first understand the nature of science.
The basic tool a scientist uses is thought. To understand
the nature of science, it is useful to focus for a moment on
how scientists think. They reason in two ways: deductively
and inductively.
Deductive Reasoning
Deductive reasoning applies general principles to predict
specific results. Over 2200 years ago, the Greek Era-
tosthenes used deductive reasoning to accurately estimate
the circumference of the earth. At high noon on the longest
day of the year, when the sun’s rays hit the bottom of a
deep well in the city of Syene, Egypt, Eratosthenes mea-
sured the length of the shadow cast by a tall obelisk in Al-
exandria, about 800 kilometers to the north. Because he
knew the distance between the two cities and the height of
the obelisk, he was able to employ the principles of Euclid-
ean geometry to correctly deduce the circumference of the
earth (figure 1.3). This sort of analysis of specific cases us-
ing general principles is an example of deductive reasoning.
It is the reasoning of mathematics and philosophy and is
used to test the validity of general ideas in all branches of
knowledge. General principles are constructed and then
used as the basis for examining specific cases.
Inductive Reasoning
Inductive reasoning uses specific observations to construct
general scientific principles. Webster’s Dictionary defines sci-
ence as systematized knowledge derived from observation
and experiment carried on to determine the principles un-
derlying what is being studied. In other words, a scientist
determines principles from observations, discovering gen-
eral principles by careful examination of specific cases. In-
ductive reasoning first became important to science in the
1600s in Europe, when Francis Bacon, Isaac Newton, and
others began to use the results of particular experiments to
infer general principles about how the world operates. If
you release an apple from your hand, what happens? The
apple falls to the ground. From a host of simple, specific
observations like this, Newton inferred a general principle:
all objects fall toward the center of the earth. What New-
ton did was construct a mental model of how the world
works, a family of general principles consistent with what
he could see and learn. Scientists do the same today. They
use specific observations to build general models, and then
test the models to see how well they work.
Science is a way of viewing the world that focuses on
objective information, putting that information to work
to build understanding.
1.2 Scientists form generalizations from observations.
FIGURE 1.3
Deductive reasoning: How Eratosthenes estimated the cir-
cumference of the earth using deductive reasoning. 1. On a
day when sunlight shone straight down a deep well at Syene in
Egypt, Eratosthenes measured the length of the shadow cast by a
tall obelisk in the city of Alexandria, about 800 kilometers away.
2. The shadow’s length and the obelisk’s height formed two sides
of a triangle. Using the recently developed principles of Euclidean
geometry, he calculated the angle, a, to be 7° and 12′, exactly
1
50
of
a circle (360°). 3. If angle a =
1
50
of a circle, then the distance
between the obelisk (in Alexandria) and the well (in Syene) must
equal
1
50
of the circumference of the earth. 4. Eratosthenes had
heard that it was a 50-day camel trip from Alexandria to Syene.
Assuming that a camel travels about 18.5 kilometers per day, he
estimated the distance between obelisk and well as 925 kilometers
(using different units of
measure, of course).
5. Eratosthenes thus de-
duced the circumference
of the earth to be 50 H11003
925 H11005 46,250
kilometers. Modern
measurements put the
distance from the well to
the obelisk at just over
800 kilometers. Employ-
ing a distance of 800
kilometers, Era-
tosthenes’s value would
have been 50 × 800 H11005
40,000 kilometers. The
actual circumference is
40,075 kilometers.
Sunlight
at midday
Dista
nc
e
b
e
tw
e
e
n
cities
=
80
0
k
m
Well
Light rays
parallel
Height of
obelisk
Length of
shadow
a
a
How Science Is Done
How do scientists establish which general principles are
true from among the many that might be true? They do
this by systematically testing alternative proposals. If these
proposals prove inconsistent with experimental observa-
tions, they are rejected as untrue. After making careful ob-
servations concerning a particular area of science, scien-
tists construct a hypothesis, which is a suggested
explanation that accounts for those observations. A hy-
pothesis is a proposition that might be true. Those hy-
potheses that have not yet been disproved are retained.
They are useful because they fit the known facts, but they
are always subject to future rejection if, in the light of new
information, they are found to be incorrect.
Testing Hypotheses
We call the test of a hypothesis an experiment (figure
1.4). Suppose that a room appears dark to you. To under-
stand why it appears dark, you propose several hypotheses.
The first might be, “There is no light in the room because
the light switch is turned off.” An alternative hypothesis
might be, “There is no light in the room because the light-
bulb is burned out.” And yet another alternative hypothe-
sis might be, “I am going blind.” To evaluate these hy-
potheses, you would conduct an experiment designed to
eliminate one or more of the hypotheses. For example, you
might test your hypotheses by reversing the position of the
light switch. If you do so and the light does not come on,
you have disproved the first hypothesis. Something other
than the setting of the light switch must be the reason for
the darkness. Note that a test such as this does not prove
that any of the other hypotheses are true; it merely dem-
onstrates that one of them is not. A successful experiment
is one in which one or more of the alternative hypotheses
is demonstrated to be inconsistent with the results and is
thus rejected.
As you proceed through this text, you will encounter
many hypotheses that have withstood the test of experiment.
Many will continue to do so; others will be revised as new
observations are made by biologists. Biology, like all science,
is in a constant state of change, with new ideas appearing
and replacing old ones.
Chapter 1 The Science of Biology 7
FIGURE 1.4
How science is done. This diagram il-
lustrates the way in which scientific in-
vestigations proceed. First, scientists
make observations that raise a
particular question. They develop a
number of potential explanations
(hypotheses) to answer the question.
Next, they carry out experiments in an
attempt to eliminate one or more of
these hypotheses. Then, predictions are
made based on the remaining
hypotheses, and further experiments
are carried out to test these predictions.
As a result of this process, the least
unlikely hypothesis is selected.
Observation
Question
Experiment
Hypothesis 1
Hypothesis 2
Hypothesis 3
Hypothesis 4
Hypothesis 5
Potential
hypotheses
Remaining
possible
hypotheses
Last remaining
possible hypothesis
Reject
hypotheses
1 and 4
Reject
hypotheses
2 and 3
Experiment
Experiment 1
Hypothesis 2
Hypothesis 3
Hypothesis 5
Hypothesis 5
Predictions
Predictions
confirmed
Experiment 2 Experiment 3 Experiment 4
Establishing Controls
Often we are interested in learning about processes that are
influenced by many factors, or variables. To evaluate alter-
native hypotheses about one variable, all other variables
must be kept constant. This is done by carrying out two ex-
periments in parallel: in the first experiment, one variable is
altered in a specific way to test a particular hypothesis; in the
second experiment, called the control experiment, that
variable is left unaltered. In all other respects the two exper-
iments are identical, so any difference in the outcomes of
the two experiments must result from the influence of the
variable that was changed. Much of the challenge of experi-
mental science lies in designing control experiments that
isolate a particular variable from other factors that might in-
fluence a process.
Using Predictions
A successful scientific hypothesis needs to be not only valid
but useful—it needs to tell you something you want to
know. A hypothesis is most useful when it makes predic-
tions, because those predictions provide a way to test the va-
lidity of the hypothesis. If an experiment produces results
inconsistent with the predictions, the hypothesis must be re-
jected. On the other hand, if the predictions are supported
by experimental testing, the hypothesis is supported. The
more experimentally supported predictions a hypothesis
makes, the more valid the hypothesis is. For example, Ein-
stein’s hypothesis of relativity was at first provisionally ac-
cepted because no one could devise an experiment that in-
validated it. The hypothesis made a clear prediction: that
the sun would bend the path of light passing by it. When
this prediction was tested in a total eclipse, the light from
background stars was indeed bent. Because this result was
unknown when the hypothesis was being formulated, it pro-
vided strong support for the hypothesis, which was then ac-
cepted with more confidence.
Developing Theories
Scientists use the word theory in two main ways. A “theo-
ry” is a proposed explanation for some natural phenome-
non, often based on some general principle. Thus one
speaks of the principle first proposed by Newton as the
“theory of gravity.” Such theories often bring together
concepts that were previously thought to be unrelated,
and offer unified explanations of different phenomena.
Newton’s theory of gravity provided a single explanation
for objects falling to the ground and the orbits of planets
around the sun. “Theory” is also used to mean the body
of interconnected concepts, supported by scientific rea-
soning and experimental evidence, that explains the facts
in some area of study. Such a theory provides an indis-
pensable framework for organizing a body of knowledge.
For example, quantum theory in physics brings together a
set of ideas about the nature of the universe, explains ex-
perimental facts, and serves as a guide to further questions
and experiments.
To a scientist, such theories are the solid ground of sci-
ence, that of which we are most certain. In contrast, to the
general public, theory implies just the opposite—a lack of
knowledge, or a guess. Not surprisingly, this difference
often results in confusion. In this text, theory will always be
used in its scientific sense, in reference to an accepted gen-
eral principle or body of knowledge.
To suggest, as many critics outside of science do, that
evolution is “just a theory” is misleading. The hypothesis
that evolution has occurred is an accepted scientific fact; it is
supported by overwhelming evidence. Modern evolutionary
theory is a complex body of ideas whose importance spreads
far beyond explaining evolution; its ramifications permeate
all areas of biology, and it provides the conceptual frame-
work that unifies biology as a science.
Research and the Scientific Method
It used to be fashionable to speak of the “scientific meth-
od” as consisting of an orderly sequence of logical “ei-
ther/or” steps. Each step would reject one of two mutually
incompatible alternatives, as if trial-and-error testing
would inevitably lead one through the maze of uncertain-
ty that always impedes scientific progress. If this were in-
deed so, a computer would make a good scientist. But sci-
ence is not done this way. As British philosopher Karl
Popper has pointed out, successful scientists without ex-
ception design their experiments with a pretty fair idea of
how the results are going to come out. They have what
Popper calls an “imaginative preconception” of what the
truth might be. A hypothesis that a successful scientist
tests is not just any hypothesis; rather, it is an educated
guess or a hunch, in which the scientist integrates all that
he or she knows and allows his or her imagination full
play, in an attempt to get a sense of what might be true
(see Box: How Biologists Do Their Work). It is because
insight and imagination play such a large role in scientific
progress that some scientists are so much better at science
than others, just as Beethoven and Mozart stand out
among most other composers.
Some scientists perform what is called basic research,
which is intended to extend the boundaries of what we
know. These individuals typically work at universities, and
their research is usually financially supported by their in-
stitutions and by external sources, such as the government,
industry, and private foundations. Basic research is as di-
verse as its name implies. Some basic scientists attempt to
find out how certain cells take up specific chemicals, while
others count the number of dents in tiger teeth. The infor-
mation generated by basic research contributes to the
growing body of scientific knowledge, and it provides the
scientific foundation utilized by applied research. Scien-
tists who conduct applied research are often employed in
8 Part I The Origin of Living Things
some kind of industry. Their work may involve the manu-
facturing of food additives, creating of new drugs, or test-
ing the quality of the environment.
After developing a hypothesis and performing a series of
experiments, a scientist writes a paper carefully describing
the experiment and its results. He or she then submits the
paper for publication in a scientific journal, but before it is
published, it must be reviewed and accepted by other scien-
tists who are familiar with that particular field of research.
This process of careful evaluation, called peer review, lies at
the heart of modern science, fostering careful work, precise
description, and thoughtful analysis. When an important
discovery is announced in a paper, other scientists attempt
to reproduce the result, providing a check on accuracy and
honesty. Nonreproducible results are not taken seriously
for long.
The explosive growth in scientific research during the
second half of the twentieth century is reflected in the
enormous number of scientific journals now in existence.
Although some, such as Science and Nature, are devoted to
a wide range of scientific disciplines, most are extremely
specialized: Cell Motility and the Cytoskeleton, Glycoconju-
gate Journal, Mutation Research, and Synapse are just a few
examples.
The scientific process involves the rejection of
hypotheses that are inconsistent with experimental
results or observations. Hypotheses that are consistent
with available data are conditionally accepted. The
formulation of the hypothesis often involves creative
insight.
Chapter 1 The Science of Biology 9
How Biologists Do
Their Work
learn why the ginkgo trees drop all their
leaves simultaneously, a scientist would
first formulate several possible answers,
called hypotheses:
Hypothesis 1: Ginkgo trees possess an inter-
nal clock that times the release of leaves to
match the season. On the day Nemerov de-
scribes, this clock sends a “drop” signal
(perhaps a chemical) to all the leaves at the
same time.
Hypothesis 2: The individual leaves of ginkgo
trees are each able to sense day length, and
when the days get short enough in the fall,
each leaf responds independently by falling.
Hypothesis 3: A strong wind arose the night
before Nemerov made his observation,
blowing all the leaves off the ginkgo trees.
Next, the scientist attempts to eliminate
one or more of the hypotheses by conduct-
ing an experiment. In this case, one might
cover some of the leaves so that they can-
not use light to sense day length. If hypoth-
esis 2 is true, then the covered leaves
should not fall when the others do, because
they are not receiving the same informa-
tion. Suppose, however, that despite the
covering of some of the leaves, all the
leaves still fall together. This result would
eliminate hypothesis 2 as a possibility. Ei-
ther of the other hypotheses, and many
others, remain possibilities.
This simple experiment with ginkgoes
points out the essence of scientific
progress: science does not prove that cer-
tain explanations are true; rather, it proves
that others are not. Hypotheses that are
inconsistent with experimental results are
rejected, while hypotheses that are not
proven false by an experiment are provi-
sionally accepted. However, hypotheses
may be rejected in the future when more
information becomes available, if they are
inconsistent with the new information. Just
as finding the correct path through a maze
by trying and eliminating false paths, sci-
entists work to find the correct explana-
tions of natural phenomena by eliminating
false possibilities.
The Consent
Late in November, on a single night
Not even near to freezing, the ginkgo trees
That stand along the walk drop all their leaves
In one consent, and neither to rain nor to wind
But as though to time alone: the golden and
green
Leaves litter the lawn today, that yesterday
Had spread aloft their fluttering fans of light.
What signal from the stars? What senses took it
in?
What in those wooden motives so decided
To strike their leaves, to down their leaves,
Rebellion or surrender? And if this
Can happen thus, what race shall be exempt?
What use to learn the lessons taught by time,
If a star at any time may tell us: Now.
Howard Nemerov
What is bothering the poet Howard Nem-
erov is that life is influenced by forces he
cannot control or even identify. It is the job
of biologists to solve puzzles such as the one
he poses, to identify and try to understand
those things that influence life.
Nemerov asks why ginkgo trees (figure
1.A) drop all their leaves at once. To find
an answer to questions such as this, biolo-
gists and other scientists pose possible an-
swers and then try to determine which an-
swers are false. Tests of alternative
possibilities are called experiments. To
FIGURE 1.A
A ginkgo tree.
10 Part I The Origin of Living Things
Darwin’s Theory of
Evolution
Darwin’s theory of evolution explains
and describes how organisms on earth
have changed over time and acquired a
diversity of new forms. This famous
theory provides a good example of how
a scientist develops a hypothesis and
how a scientific theory grows and wins
acceptance.
Charles Robert Darwin (1809–1882;
figure 1.5) was an English naturalist
who, after 30 years of study and obser-
vation, wrote one of the most famous
and influential books of all time. This
book, On the Origin of Species by Means
of Natural Selection, or The Preservation
of Favoured Races in the Struggle for Life,
created a sensation when it was pub-
lished, and the ideas Darwin expressed
in it have played a central role in the
development of human thought ever
since.
In Darwin’s time, most people be-
lieved that the various kinds of organ-
isms and their individual structures re-
sulted from direct actions of the Creator
(and to this day many people still believe
this to be true). Species were thought to
be specially created and unchangeable,
or immutable, over the course of time.
In contrast to these views, a number of
earlier philosophers had presented the
view that living things must have
changed during the history of life on
earth. Darwin proposed a concept he
called natural selection as a coherent,
logical explanation for this process, and
he brought his ideas to wide public at-
tention. His book, as its title indicates,
presented a conclusion that differed
sharply from conventional wisdom. Al-
though his theory did not challenge the
existence of a Divine Creator, Darwin
argued that this Creator did not simply
create things and then leave them forev-
er unchanged. Instead, Darwin’s God
expressed Himself through the operation of natural laws
that produced change over time, or evolution. These
views put Darwin at odds with most people of his time,
who believed in a literal interpretation of the Bible and ac-
cepted the idea of a fixed and constant world. His revolu-
tionary theory deeply troubled not only many of his con-
temporaries but Darwin himself.
The story of Darwin and his theory begins in 1831, when
he was 22 years old. On the recommendation of one of his
professors at Cambridge University, he was selected to serve
1.3 Darwin’s theory of evolution illustrates how science works.
FIGURE 1.5
Charles Darwin. This newly rediscovered photograph taken in 1881, the year before
Darwin died, appears to be the last ever taken of the great biologist.
as naturalist on a five-year navigational mapping expedition
around the coasts of South America (figure 1.6), aboard
H.M.S. Beagle (figure 1.7). During this long voyage, Darwin
had the chance to study a wide variety of plants and animals
on continents and islands and in distant seas. He was able to
explore the biological richness of the tropical forests, exam-
ine the extraordinary fossils of huge extinct mammals in
Patagonia at the southern tip of South America, and observe
the remarkable series of related but distinct forms of life on
the Galápagos Islands, off the west coast of South America.
Such an opportunity clearly played an important role in the
development of his thoughts about the nature of life on
earth.
When Darwin returned from the voyage at the age of 27,
he began a long period of study and contemplation. During
the next 10 years, he published important books on several
different subjects, including the formation of oceanic islands
from coral reefs and the geology of South America. He also
devoted eight years of study to barnacles, a group of small
marine animals with shells that inhabit rocks and pilings,
eventually writing a four-volume work on their classification
and natural history. In 1842, Darwin and his family moved
out of London to a country home at Down, in the county of
Kent. In these pleasant surroundings, Darwin lived, studied,
and wrote for the next 40 years.
Darwin was the first to propose natural selection as an
explanation for the mechanism of evolution that
produced the diversity of life on earth. His hypothesis
grew from his observations on a five-year voyage around
the world.
Chapter 1 The Science of Biology 11
British Isles
Western
Isles
Europe
Africa
Indian
Ocean
Madagascar
Mauritius
Bourbon Island
Cape of
Good Hope
King George’s
Sound
Hobart
Sydney
Australia
New
Zealand
Friendly
Islands
Phillippine
Islands
Equator
North Pacific
Ocean
Asia
North Atlantic
Ocean
Cape Verde
Marquesas
Galápagos
Islands
Valparaiso
Society
Islands
Straits of Magellan
Tierra del FuegoCape Horn
Falkland
Islands
Port Desire
South Atlantic
Ocean
Montevideo
Buenos Aires
Rio de Janeiro
St. Helena
Ascension
North America
Canary
Islands
Keeling
Islands
South
America
Bahia
FIGURE 1.6
The five-year voyage of H.M.S. Beagle. Most of the time was spent exploring the coasts and coastal islands of South America,
such as the Galápagos Islands. Darwin’s studies of the animals of the Galápagos Islands played a key role in his eventual
development of the theory of evolution by means of natural selection.
FIGURE 1.7
Cross section of the
Beagle. A 10-gun brig of
242 tons, only 90 feet in
length, the Beagle had a
crew of 74 people! After he
first saw the ship, Darwin
wrote to his college
professor Henslow: “The
absolute want of room is an
evil that nothing can
surmount.”
12 Part I The Origin of Living Things
Darwin’s Evidence
One of the obstacles that had blocked the acceptance of
any theory of evolution in Darwin’s day was the incorrect
notion, widely believed at that time, that the earth was
only a few thousand years old. Evidence discovered during
Darwin’s time made this assertion seem less and less likely.
The great geologist Charles Lyell (1797–1875), whose
Principles of Geology (1830) Darwin read eagerly as he
sailed on the Beagle, outlined for the first time the story of
an ancient world of plants and animals in flux. In this
world, species were constantly becoming extinct while oth-
ers were emerging. It was this world that Darwin sought to
explain.
What Darwin Saw
When the Beagle set sail, Darwin was fully convinced that
species were immutable. Indeed, it was not until two or
three years after his return that he began to consider seri-
ously the possibility that they could change. Nevertheless,
during his five years on the ship, Darwin observed a number
of phenomena that were of central importance to him in
reaching his ultimate conclusion (table 1.1). For example, in
the rich fossil beds of southern South America, he observed
fossils of extinct armadillos similar to the armadillos that
still lived in the same area (figure 1.8). Why would similar
living and fossil organisms be in the same area unless the
earlier form had given rise to the other?
Repeatedly, Darwin saw that the characteristics of simi-
lar species varied somewhat from place to place. These
geographical patterns suggested to him that organismal lin-
eages change gradually as species migrate from one area to
another. On the Galápagos Islands, off the coast of Ecua-
dor, Darwin encountered giant land tortoises. Surprisingly,
these tortoises were not all identical. In fact, local residents
and the sailors who captured the tortoises for food could
tell which island a particular tortoise had come from just by
looking at its shell. This distribution of physical variation
suggested that all of the tortoises were related, but that
they had changed slightly in appearance after becoming
isolated on different islands.
In a more general sense, Darwin was struck by the fact
that the plants and animals on these relatively young vol-
canic islands resembled those on the nearby coast of
South America. If each one of these plants and animals
had been created independently and simply placed on the
Galápagos Islands, why didn’t they resemble the plants
and animals of islands with similar climates, such as those
off the coast of Africa, for example? Why did they resem-
ble those of the adjacent South American coast instead?
The fossils and patterns of life that Darwin observed on
the voyage of the Beagle eventually convinced him that
evolution had taken place.
Table 1.1 Darwin’s Evidence
that Evolution Occurs
FOSSILS
1. Extinct species, such as the fossil armadillo in figure 1.8,
most closely resemble living ones in the same area,
suggesting that one had given rise to the other.
2. In rock strata (layers), progressive changes in characteristics
can be seen in fossils from earlier and earlier layers.
GEOGRAPHICAL DISTRIBUTION
3. Lands with similar climates, such as Australia, South Africa,
California, and Chile, have unrelated plants and animals,
indicating that diversity is not entirely influenced by climate
and environment.
4. The plants and animals of each continent are distinctive;
all South American rodents belong to a single group,
structurally similar to the guinea pigs, for example, while
most of the rodents found elsewhere belong to other
groups.
OCEANIC ISLANDS
5. Although oceanic islands have few species, those they do
have are often unique (endemic) and show relatedness to
one another, such as the Galápagos tortoises. This suggests
that the tortoises and other groups of endemic species
developed after their mainland ancestors reached the islands
and are, therefore, more closely related to one another.
6. Species on oceanic islands show strong affinities to those on
the nearest mainland. Thus, the finches of the Galápagos
Islands closely resemble a finch seen on the western coast of
South America. The Galápagos finches do not resemble the
birds on the Cape Verde Islands, islands in the Atlantic
Ocean off the coast of Africa that are similar to the
Galápagos. Darwin visited the Cape Verde Islands and
many other island groups personally and was able to make
such comparisons on the basis of his own observations.
FIGURE 1.8
Fossil evidence of evolution. The now-extinct glyptodont (a)
was a 2000-kilogram South American armadillo, much larger than
the modern armadillo (b), which weighs an average of about 4.5
kilograms. (Drawings are not to scale.)
(a) Glyptodont
(b) Armadillo
Inventing the Theory
of Natural Selection
It is one thing to observe the results of evolution, but
quite another to understand how it happens. Darwin’s
great achievement lies in his formulation of the hypothe-
sis that evolution occurs because of natural selection.
Darwin and Malthus
Of key importance to the development of Darwin’s in-
sight was his study of Thomas Malthus’s Essay on the
Principle of Population (1798). In his book, Malthus
pointed out that populations of plants and animals (in-
cluding human beings) tend to increase geometrically,
while the ability of humans to increase their food supply
increases only arithmetically. A geometric progression is
one in which the elements increase by a constant factor;
for example, in the progression 2, 6, 18, 54, ..., each
number is three times the preceding one. An arithmetic
progression, in contrast, is one in which the elements in-
crease by a constant difference; in the progression 2, 6, 10,
14,..., each number is four greater than the preced-
ing one (figure 1.9).
Because populations increase geometrically, virtually
any kind of animal or plant, if it could reproduce un-
checked, would cover the entire surface of the world
within a surprisingly short time. Instead, populations of
species remain fairly constant year after year, because
death limits population numbers. Malthus’s conclusion
provided the key ingredient that was necessary for Dar-
win to develop the hypothesis that evolution occurs by
natural selection.
Sparked by Malthus’s ideas, Darwin saw that although
every organism has the potential to produce more off-
spring than can survive, only a limited number actually
do survive and produce further offspring. Combining
this observation with what he had seen on the voyage of
the Beagle, as well as with his own experiences in breed-
ing domestic animals, Darwin made an important associ-
ation (figure 1.10): Those individuals that possess supe-
rior physical, behavioral, or other attributes are more
likely to survive than those that are not so well endowed.
By surviving, they gain the opportunity to pass on their
favorable characteristics to their offspring. As the fre-
quency of these characteristics increases in the popula-
tion, the nature of the population as a whole will gradu-
ally change. Darwin called this process selection. The
driving force he identified has often been referred to as
survival of the fittest.
Chapter 1 The Science of Biology 13
Geometric
progression
Arithmetic
progression
2
6
18
54
4
6
8
FIGURE 1.9
Geometric and arithmetic progressions. A geometric progression
increases by a constant factor (e.g., H11003 2 or H11003 3 or H11003 4), while an
arithmetic progression increases by a constant difference (e.g.,
units of 1 or 2 or 3) . Malthus contended that the human growth
curve was geometric, but the human food production curve was
only arithmetic. Can you see the problems this difference would
cause?
FIGURE 1.10
An excerpt from Charles Darwin’s On the Origin of Species.
Natural Selection
Darwin was thoroughly familiar with
variation in domesticated animals and
began On the Origin of Species with a
detailed discussion of pigeon breeding.
He knew that breeders selected certain
varieties of pigeons and other animals,
such as dogs, to produce certain char-
acteristics, a process Darwin called ar-
tificial selection. Once this had been
done, the animals would breed true for
the characteristics that had been select-
ed. Darwin had also observed that the
differences purposely developed be-
tween domesticated races or breeds
were often greater than those that sep-
arated wild species. Domestic pigeon
breeds, for example, show much
greater variety than all of the hundreds
of wild species of pigeons found
throughout the world. Such relation-
ships suggested to Darwin that evolu-
tionary change could occur in nature
too. Surely if pigeon breeders could
foster such variation by “artificial selec-
tion,” nature could do the same, play-
ing the breeder’s role in selecting the
next generation—a process Darwin
called natural selection.
Darwin’s theory thus incorporates
the hypothesis of evolution, the pro-
cess of natural selection, and the mass of new evidence
for both evolution and natural selection that Darwin
compiled. Thus, Darwin’s theory provides a simple and
direct explanation of biological diversity, or why animals
are different in different places: because habitats differ in
their requirements and opportunities, the organisms with
characteristics favored locally by natural selection will
tend to vary in different places.
Darwin Drafts His Argument
Darwin drafted the overall argument for evolution by natu-
ral selection in a preliminary manuscript in 1842. After
showing the manuscript to a few of his closest scientific
friends, however, Darwin put it in a drawer, and for
16 years turned to other research. No one knows for sure
why Darwin did not publish his initial manuscript—it is
very thorough and outlines his ideas in detail. Some histo-
rians have suggested that Darwin was shy of igniting public
criticism of his evolutionary ideas—there could have been
little doubt in his mind that his theory of evolution by nat-
ural selection would spark controversy. Others have pro-
posed that Darwin was simply refining
his theory all those years, although
there is little evidence he altered his
initial manuscript in all that time.
Wallace Has the Same Idea
The stimulus that finally brought Dar-
win’s theory into print was an essay he
received in 1858. A young English nat-
uralist named Alfred Russel Wallace
(1823–1913) sent the essay to Darwin
from Malaysia; it concisely set forth
the theory of evolution by means of
natural selection, a theory Wallace had
developed independently of Darwin.
Like Darwin, Wallace had been
greatly influenced by Malthus’s 1798
essay. Colleagues of Wallace, knowing
of Darwin’s work, encouraged him to
communicate with Darwin. After re-
ceiving Wallace’s essay, Darwin ar-
ranged for a joint presentation of their
ideas at a seminar in London. Darwin
then completed his own book, expand-
ing the 1842 manuscript which he had
written so long ago, and submitted it
for publication.
Publication of Darwin’s Theory
Darwin’s book appeared in November 1859 and caused an
immediate sensation. Many people were deeply disturbed by
the suggestion that human beings were descended from the
same ancestor as apes (figure 1.11). Darwin did not actually
discuss this idea in his book, but it followed directly from the
principles he outlined. In a subsequent book, The Descent of
Man, Darwin presented the argument directly, building a
powerful case that humans and living apes have common an-
cestors. Although people had long accepted that humans
closely resembled apes in many characteristics, the possibility
that there might be a direct evolutionary relationship was un-
acceptable to many. Darwin’s arguments for the theory of
evolution by natural selection were so compelling, however,
that his views were almost completely accepted within the in-
tellectual community of Great Britain after the 1860s.
The fact that populations do not really expand
geometrically implies that nature acts to limit
population numbers. The traits of organisms that
survive to produce more offspring will be more
common in future generations—a process Darwin
called natural selection.
14 Part I The Origin of Living Things
FIGURE 1.11
Darwin greets his monkey ancestor. In
his time, Darwin was often portrayed
unsympathetically, as in this drawing from
an 1874 publication.
Evolution After Darwin:
More Evidence
More than a century has elapsed since Darwin’s death in
1882. During this period, the evidence supporting his the-
ory has grown progressively stronger. There have also
been many significant advances in our understanding of
how evolution works. Although these advances have not
altered the basic structure of Darwin’s theory, they have
taught us a great deal more about the mechanisms by
which evolution occurs. We will briefly explore some of
this evidence here; in chapter 21 we will return to the the-
ory of evolution and examine the evidence in more detail.
The Fossil Record
Darwin predicted that the fossil record would yield inter-
mediate links between the great groups of organisms, for
example, between fishes and the amphibians thought to
have arisen from them, and between reptiles and birds. We
now know the fossil record to a degree that was unthink-
able in the nineteenth century. Recent discoveries of mi-
croscopic fossils have extended the known history of life on
earth back to about 3.5 billion years ago. The discovery of
other fossils has supported Darwin’s predictions and has
shed light on how organisms have, over this enormous time
span, evolved from the simple to the complex. For verte-
brate animals especially, the fossil record is rich and exhib-
its a graded series of changes in form, with the evolutionary
parade visible for all to see (see Box: Why Study Fossils?).
The Age of the Earth
In Darwin’s day, some physicists argued that the earth was
only a few thousand years old. This bothered Darwin, be-
cause the evolution of all living things from some single
original ancestor would have required a great deal more
time. Using evidence obtained by studying the rates of ra-
dioactive decay, we now know that the physicists of Dar-
win’s time were wrong, very wrong: the earth was formed
about 4.5 billion years ago.
The Mechanism of Heredity
Darwin received some of his sharpest criticism in the area of
heredity. At that time, no one had any concept of genes or
of how heredity works, so it was not possible for Darwin to
explain completely how evolution occurs. Theories of he-
redity in Darwin’s day seemed to rule out the possibility of
genetic variation in nature, a critical requirement of Dar-
win’s theory. Genetics was established as a science only at
the start of the twentieth century, 40 years after the publica-
tion of Darwin’s On the Origin of Species. When scientists
began to understand the laws of inheritance (discussed in
chapter 13), the heredity problem with Darwin’s theory
vanished. Genetics accounts in a neat and orderly way for
the production of new variations in organisms.
Comparative Anatomy
Comparative studies of animals have provided strong evi-
dence for Darwin’s theory. In many different types of verte-
brates, for example, the same bones are present, indicating
their evolutionary past. Thus, the forelimbs shown in figure
1.12 are all constructed from the same basic array of bones,
modified in one way in the wing of a bat, in another way in
the fin of a porpoise, and in yet another way in the leg of a
horse. The bones are said to be homologous in the differ-
ent vertebrates; that is, they have the same evolutionary ori-
gin, but they now differ in structure and function. This con-
trasts with analogous structures, such as the wings of birds
and butterflies, which have similar structure and function
but different evolutionary origins.
Chapter 1 The Science of Biology 15
Human Cat Bat Porpoise Horse
FIGURE 1.12
Homology among vertebrate
limbs. The forelimbs of these
five vertebrates show the ways
in which the relative
proportions of the forelimb
bones have changed in relation
to the particular way of life of
each organism.
Molecular Biology
Biochemical tools are now of major importance in efforts to
reach a better understanding of how evolution occurs.
Within the last few years, for example, evolutionary biolo-
gists have begun to “read” genes, much as you are reading
this page. They have learned to recognize the order of the
“letters” of the long DNA molecules, which are present in
every living cell and which provide the genetic information
for that organism. By comparing the sequences of “letters”
in the DNA of different groups of animals or plants, we can
specify the degree of relationship among the groups more
precisely than by any other means. In many cases, detailed
family trees can then be constructed. The consistent pattern
emerging from a growing mountain of data is one of pro-
gressive change over time, with more distantly related
species showing more differences in their DNA than closely
related ones, just as Darwin’s theory predicts. By measuring
the degree of difference in the genetic coding, and by inter-
preting the information available from the fossil record, we
can even estimate the rates at which evolution is occurring
in different groups of organisms.
Development
Twentieth-century knowledge about growth and develop-
ment further supports Darwin’s theory of evolution. Strik-
ing similarities are seen in the developmental stages of
many organisms of different species. Human embryos, for
example, go through a stage in which they possess the
same structures that give rise to the gills in fish, a tail, and
even a stage when the embryo has fur! Thus, the develop-
ment of an organism (its ontogeny) often yields informa-
tion about the evolutionary history of the species as a
whole (its phylogeny).
Since Darwin’s time, new discoveries of the fossil
record, genetics, anatomy, and development all support
Darwin’s theory.
16 Part I The Origin of Living Things
by studying modern organisms. But history
is complex and unpredictable—and princi-
ples of evolution (like natural selection)
cannot specify the pathway that life’s histo-
ry has actually followed. Paleontology holds
the archives of the pathway—the fossil
record of past life, with its fascinating histo-
ry of mass extinctions, periods of rapid
change, long episodes of stability, and con-
stantly changing patterns of dominance and
diversity. Humans represent just one tiny,
largely fortuitous, and late-arising twig on
the enormously arborescent bush of life.
Paleontology is the study of this grandest of
all bushes.
geological time, occur by a natural process
of evolutionary transformation—“descent
with modification,” in Darwin’s words. I
was thrilled to learn that humans had arisen
from apelike ancestors, who had themselves
evolved from the tiny mouselike mammals
that had lived in the time of dinosaurs and
seemed then so inconspicuous, so unsuc-
cessful, and so unpromising.
Now, at mid-career (I was born in 1941)
I remain convinced that I made the right
choice, and committed to learn and convey,
as much as I can as long as I can, about evo-
lution and the history of life. We can learn
a great deal about the process of evolution
I grew up on the streets of New York City,
in a family of modest means and little for-
mal education, but with a deep love of
learning. Like many urban kids who be-
come naturalists, my inspiration came
from a great museum—in particular, from
the magnificent dinosaurs on display at the
American Museum of Natural History. As
we all know from Jurassic Park and other
sources, dinomania in young children (I
was five when I saw my first dinosaur) is
not rare—but nearly all children lose the
passion, and the desire to become a pale-
ontologist becomes a transient moment
between policeman and fireman in a chro-
nology of intended professions. But I per-
sisted and became a professional paleontol-
ogist, a student of life’s history as revealed
by the evidence of fossils (though I ended
up working on snails rather than dino-
saurs!). Why?
I remained committed to paleontology
because I discovered, still as a child, the
wonder of one of the greatest transforming
ideas ever discovered by science: evolution.
I learned that those dinosaurs, and all crea-
tures that have ever lived, are bound to-
gether in a grand family tree of physical re-
lationships, and that the rich and fascinating
changes of life, through billions of years in
Why Study Fossils?
Flight has evolved
three separate
times among ver-
tebrates. Birds and
bats are still with
us, but pterosaurs,
such as the one
pictured, became
extinct with the di-
nosaurs about 65
million years ago.
Stephen Jay Gould
Harvard University
Chapter 1 The Science of Biology 17
Core Principles
of Biology
From centuries of biological observation and inquiry, one
organizing principle has emerged: biological diversity re-
flects history, a record of success, failure, and change ex-
tending back to a period soon after the formation of the
earth. The explanation for this diversity, the theory of evo-
lution by natural selection, will form the backbone of your
study of biological science, just as the theory of the covalent
bond is the backbone of chemistry, or the theory of quan-
tum mechanics is that of physics. Evolution by natural selec-
tion is a thread that runs through everything you will learn
in this book.
Basic Principles
The first half of this book is devoted to a description of the
basic principles of biology, introduced through a levels-of-
organization framework (see figure 1.2). At the molecular,
organellar, and cellular levels of organization, you will be in-
troduced to cell biology. You will learn how cells are con-
structed and how they grow, divide, and communicate. At
the organismal level, you will learn the principles of genetics,
which deal with the way that individual traits are transmit-
ted from one generation to the next. At the population level,
you will examine evolution, the gradual change in popula-
tions from one generation to the next, which has led
through natural selection to the biological diversity we see
around us. Finally, at the community and ecosystem levels,
you will study ecology, which deals with how organisms in-
teract with their environments and with one another to pro-
duce the complex communities characteristic of life on
earth.
Organisms
The second half of the book is devoted to an examination of
organisms, the products of evolution. It is estimated that at
least 5 million different kinds of plants, animals, and micro-
organisms exist, and their diversity is incredible (figure 1.13).
Later in the book, we will take a particularly detailed look at
the vertebrates, the group of animals of which we are mem-
bers. We will consider the vertebrate body and how it func-
tions, as this information is of greatest interest and impor-
tance to most students.
As you proceed through this book, what you learn at one
stage will give you the tools to understand the next. The
core principle of biology is that biological diversity is the
result of a long evolutionary journey.
1.4 This book is organized to help you learn biology.
Plantae
Animalia
Fungi
Eubacteria
Archaebacteria
Protista
FIGURE 1.13
The diversity of life. Biologists categorize
all living things into six major groups
called kingdoms: archaebacteria,
eubacteria, protists, fungi, plants, and
animals.
Chapter 1
Summary Questions Media Resources
1.1 Biology is the science of life.
18 Part I The Origin of Living Things
? Living things are highly organized, whether as single
cells or as multicellular organisms, with several hier-
archical levels.
1. What are the characteristics
of living things?
1.2 Scientists form generalizations from observations.
? Science is the determination of general principles
from observation and experimentation.
? Scientists select the best hypotheses by using
controlled experiments to eliminate alternative
hypotheses that are inconsistent with observations.
? A group of related hypotheses supported by a large
body of evidence is called a theory. In science, a
theory represents what we are most sure about.
However, there are no absolute truths in science, and
even theories are accepted only conditionally.
? Scientists conduct basic research, designed to gain
information about natural phenomena in order to
contribute to our overall body of knowledge, and
applied research, devoted to solving specific problems
with practical applications.
2. What is the difference be-
tween deductive and inductive
reasoning? What is a hypothesis?
3. What are variables? How are
control experiments used in test-
ing hypotheses?
4. How does a hypothesis
become a theory? At what point
does a theory become accepted
as an absolute truth, no longer
subject to any uncertainty?
5. What is the difference
between basic and applied
research?
6. Describe the evidence that led
Darwin to propose that evolu-
tion occurs by means of natural
selection. What evidence
gathered since the publication of
Darwin’s theory has lent further
support to the theory?
7. What is the difference be-
tween homologous and analo-
gous structures? Give an
example of each.
8. Can you think of any alterna-
tives to levels-of-organization as
ways of organizing the mass of
information in biology?
? One of the central theories of biology is Darwin’s
theory that evolution occurs by natural selection. It
states that certain individuals have heritable traits that
allow them to produce more offspring in a given kind
of environment than other individuals lacking those
traits. Consequently, those traits will increase in
frequency through time.
? Because environments differ in their requirements
and opportunities, the traits favored by natural
selection will vary in different environments.
? This theory is supported by a wealth of evidence ac-
quired over more than a century of testing and
questioning.
? Biological diversity is the result of a long history of
evolutionary change. For this reason evolution is the
core of the science of biology.
? Considered in terms of levels-of-organization, the
science of biology can be said to consist of subdisci-
plines focusing on particular levels. Thus one speaks
of molecular biology, cell biology, organismal biolo-
gy, population biology, and community biology.
1.3 Darwin’s theory of evolution illustrates how science works.
1.4 This book is organized to help you learn biology.
? Art Activity: Biological
organization
? Scientists on Science:
Why Paleonthology?
? Experiments:
Probability and
Hypothesis Testing in
Biology
? Introduction to
Evolution
? Before Darwin
? Voyage of the Beagle
? Natural Selection
? The Process of Natural
Selection
? Evidence for Evolution
? Student Research: The
Search for Medicinal
Plants on Science
Articles
? 140 Years Without
Darwin Are Enough
? Bird-Killing Cats:
Nature’s Way of
Making Better Bids
http://www.mhhe.com/raven6e http://www.biocourse.com
19
2
The Nature
of Molecules
Concept Outline
2.1 Atoms are nature’s building material.
Atoms. All substances are composed of tiny particles called
atoms, each a positively charged nucleus around which orbit
negative electrons.
Electrons Determine the Chemical Behavior of Atoms.
Electrons orbit the nucleus of an atom; the closer an
electron’s orbit to the nucleus, the lower its energy level.
2.2 The atoms of living things are among the smallest.
Kinds of Atoms. Of the 92 naturally occurring elements,
only 11 occur in organisms in significant amounts.
2.3 Chemical bonds hold molecules together.
Ionic Bonds Form Crystals. Atoms are linked together
into molecules, joined by chemical bonds that result from
forces like the attraction of opposite charges or the sharing of
electrons.
Covalent Bonds Build Stable Molecules. Chemical
bonds formed by the sharing of electrons can be very strong,
and require much energy to break.
2.4 Water is the cradle of life.
Chemistry of Water. Water forms weak chemical
associations that are responsible for much of the organization
of living chemistry.
Water Atoms Act Like Tiny Magnets. Because electrons
are shared unequally by the hydrogen and oxygen atoms of
water, a partial charge separation occurs. Each water atom
acquires a positive and negative pole and is said to be “polar.”
Water Clings to Polar Molecules. Because the opposite
partial charges of polar molecules attract one another, water
tends to cling to itself and other polar molecules and to
exclude nonpolar molecules.
Water Ionizes. Because its covalent bonds occasionally
break, water contains a low concentration of hydrogen (H
+
)
and hydroxide (OH
–
) ions, the fragments of broken water
molecules.
A
bout 10 to 20 billion years ago, an enormous explo-
sion likely marked the beginning of the universe.
With this explosion began the process of evolution, which
eventually led to the origin and diversification of life on
earth. When viewed from the perspective of 20 billion
years, life within our solar system is a recent development,
but to understand the origin of life, we need to consider
events that took place much earlier. The same processes
that led to the evolution of life were responsible for the
evolution of molecules (figure 2.1). Thus, our study of life
on earth begins with physics and chemistry. As chemical
machines ourselves, we must understand chemistry to
begin to understand our origins.
FIGURE 2.1
Cells are made of molecules. Specific, often simple, combina-
tions of atoms yield an astonishing diversity of molecules within
the cell, each with unique functional characteristics.
weight will be greater on the earth because the earth’s grav-
itational force is greater than the moon’s. The atomic
mass of an atom is equal to the sum of the masses of its
protons and neutrons. Atoms that occur naturally on earth
contain from 1 to 92 protons and up to 146 neutrons.
The mass of atoms and subatomic particles is measured
in units called daltons. To give you an idea of just how small
these units are, note that it takes 602 million million billion
(6.02 × 10
23
) daltons to make 1 gram! A proton weighs ap-
proximately 1 dalton (actually 1.009 daltons), as does a neu-
tron (1.007 daltons). In contrast, electrons weigh only
1
1840
of
a dalton, so their contribution to the overall mass of an atom
is negligible.
20 Part I The Origin of Living Things
Atoms
Any substance in the universe that has
mass (see below) and occupies space is
defined as matter. All matter is com-
posed of extremely small particles
called atoms. Because of their size,
atoms are difficult to study. Not until
early in this century did scientists
carry out the first experiments sug-
gesting what an atom is like.
The Structure of Atoms
Objects as small as atoms can be
“seen” only indirectly, by using very
complex technology such as tunneling
microcopy. We now know a great
deal about the complexities of atomic
structure, but the simple view put
forth in 1913 by the Danish physicist
Niels Bohr provides a good starting
point. Bohr proposed that every atom
possesses an orbiting cloud of tiny
subatomic particles called electrons
whizzing around a core like the plan-
ets of a miniature solar system. At the
center of each atom is a small, very
dense nucleus formed of two other
kinds of subatomic particles, protons
and neutrons (figure 2.2).
Within the nucleus, the cluster of
protons and neutrons is held together
by a force that works only over short
subatomic distances. Each proton car-
ries a positive (+) charge, and each
electron carries a negative (–) charge.
Typically an atom has one electron
for each proton. The number of protons (the atom’s
atomic number) determines the chemical character of the
atom, because it dictates the number of electrons orbiting
the nucleus which are available for chemical activity. Neu-
trons, as their name implies, possess no charge.
Atomic Mass
The terms mass and weight are often used interchangeably,
but they have slightly different meanings. Mass refers to the
amount of a substance, while weight refers to the force
gravity exerts on a substance. Hence, an object has the
same mass whether it is on the earth or the moon, but its
2.1 Atoms are nature’s building material.
FIGURE 2.2
Basic structure of atoms. All atoms have a nucleus consisting of protons and neutrons,
except hydrogen, the smallest atom, which has only one proton and no neutrons in its
nucleus. Oxygen, for example, has eight protons and eight neutrons in its nucleus. Electrons
spin around the nucleus a far distance away from the nucleus.
Proton
(Positive charge)
(No charge) (Negative charge)
Neutron Electron
Hydrogen
1 Proton
1 Electron
Oxygen
8 Protons
8 Neutrons
8 Electrons
Isotopes
Atoms with the same atomic number (that is, the same num-
ber of protons) have the same chemical properties and are
said to belong to the same element. Formally speaking, an
element is any substance that cannot be broken down to any
other substance by ordinary chemical means. However, while
all atoms of an element have the same number of protons,
they may not all have the same number of neutrons. Atoms of
an element that possess different numbers of neutrons are
called isotopes of that element. Most elements in nature exist
as mixtures of different isotopes. Carbon (C), for example,
has three isotopes, all containing six protons (figure 2.3).
Over 99% of the carbon found in nature exists as an isotope
with six neutrons. Because its total mass is 12 daltons (6 from
protons plus 6 from neutrons), this isotope is referred to as
carbon-12, and symbolized
12
C. Most of the rest of the natu-
rally occurring carbon is carbon-13, an isotope with seven
neutrons. The rarest carbon isotope is carbon-14, with eight
neutrons. Unlike the other two isotopes, carbon-14 is unsta-
ble: its nucleus tends to break up into elements with lower
atomic numbers. This nuclear breakup, which emits a signifi-
cant amount of energy, is called radioactive decay, and iso-
topes that decay in this fashion are radioactive isotopes.
Some radioactive isotopes are more unstable than others
and therefore decay more readily. For any given isotope,
however, the rate of decay is constant. This rate is usually
expressed as the half-life, the time it takes for one half of the
atoms in a sample to decay. Carbon-14, for example, has a
half-life of about 5600 years. A sample of carbon containing
1 gram of carbon-14 today would contain 0.5 gram of car-
bon-14 after 5600 years, 0.25 gram 11,200 years from now,
0.125 gram 16,800 years from now, and so on. By determin-
ing the ratios of the different isotopes of carbon and other
elements in biological samples and in rocks, scientists are
able to accurately determine when these materials formed.
While there are many useful applications of radioactivity,
there are also harmful side effects that must be considered in
any planned use of radioactive substances. Radioactive sub-
stances emit energetic subatomic particles that have the po-
tential to severely damage living cells, producing mutations in
their genes, and, at high doses, cell death. Consequently, ex-
posure to radiation is now very carefully controlled and regu-
lated. Scientists who work with radioactivity (basic re-
searchers as well as applied scientists such as X-ray
technologists) wear radiation-sensitive badges to monitor the
total amount of radioactivity to which they are exposed. Each
month the badges are collected and scrutinized. Thus, em-
ployees whose work places them in danger of excessive radio-
active exposure are equipped with an “early warning system.”
Electrons
The positive charges in the nucleus of an atom are counter-
balanced by negatively charged electrons orbiting at vary-
ing distances around the nucleus. Thus, atoms with the
same number of protons and electrons are electrically neu-
tral, having no net charge.
Electrons are maintained in their orbits by their attrac-
tion to the positively charged nucleus. Sometimes other
forces overcome this attraction and an atom loses one or
more electrons. In other cases, atoms may gain additional
electrons. Atoms in which the number of electrons does
not equal the number of protons are known as ions, and
they carry a net electrical charge. An atom that has more
protons than electrons has a net positive charge and is
called a cation. For example, an atom of sodium (Na) that
has lost one electron becomes a sodium ion (Na
+
), with a
charge of +1. An atom that has fewer protons than elec-
trons carries a net negative charge and is called an anion. A
chlorine atom (Cl) that has gained one electron becomes a
chloride ion (Cl
–
), with a charge of –1.
An atom consists of a nucleus of protons and neutrons
surrounded by a cloud of electrons. The number of its
electrons largely determines the chemical properties of
an atom. Atoms that have the same number of protons
but different numbers of neutrons are called isotopes.
Isotopes of an atom differ in atomic mass but have
similar chemical properties.
Chapter 2 The Nature of Molecules 21
Carbon-12
6 Protons
6 Neutrons
6 Electrons
Carbon-13
6 Protons
7 Neutrons
6 Electrons
Carbon-14
6 Protons
8 Neutrons
6 Electrons
FIGURE 2.3
The three most abundant
isotopes of carbon. Isotopes
of a particular atom have
different numbers of
neutrons.
Electrons Determine the Chemical
Behavior of Atoms
The key to the chemical behavior of an atom lies in the ar-
rangement of its electrons in their orbits. It is convenient to
visualize individual electrons as following discrete circular
orbits around a central nucleus, as in the Bohr model of the
atom. However, such a simple picture is not realistic. It is
not possible to precisely locate the position of any individual
electron precisely at any given time. In fact, a particular
electron can be anywhere at a given instant, from close to
the nucleus to infinitely far away from it.
However, a particular electron is more likely to be locat-
ed in some positions than in others. The area around a nu-
cleus where an electron is most likely to be found is called
the orbital of that electron (figure 2.4). Some electron or-
bitals near the nucleus are spherical (s orbitals), while oth-
ers are dumbbell-shaped (p orbitals). Still other orbitals,
more distant from the nucleus, may have different shapes.
Regardless of its shape, no orbital may contain more than
two electrons.
Almost all of the volume of an atom is empty space, be-
cause the electrons are quite far from the nucleus relative
to its size. If the nucleus of an atom were the size of an ap-
ple, the orbit of the nearest electron would be more than
1600 meters away. Consequently, the nuclei of two atoms
never come close enough in nature to interact with each
other. It is for this reason that an atom’s electrons, not its
protons or neutrons, determine its chemical behavior. This
also explains why the isotopes of an element, all of which
have the same arrangement of electrons, behave the same
way chemically.
Energy within the Atom
All atoms possess energy, defined as the ability to do work.
Because electrons are attracted to the positively charged
nucleus, it takes work to keep them in orbit, just as it takes
work to hold a grapefruit in your hand against the pull of
gravity. The grapefruit is said to possess potential energy,
the ability to do work, because of its position; if you were
to release it, the grapefruit would fall and its energy would
be reduced. Conversely, if you were to move the grapefruit
to the top of a building, you would increase its potential
energy. Similarly, electrons have potential energy of posi-
tion. To oppose the attraction of the nucleus and move the
electron to a more distant orbital requires an input of en-
ergy and results in an electron with greater potential ener-
gy. This is how chlorophyll captures energy from light
during photosynthesis (chapter 10)—the light excites elec-
trons in the chlorophyll. Moving an electron closer to the
nucleus has the opposite effect: energy is released, usually
as heat, and the electron ends up with less potential energy
(figure 2.5).
A given atom can possess only certain discrete amounts
of energy. Like the potential energy of a grapefruit on a step
of a staircase, the potential energy contributed by the posi-
tion of an electron in an atom can have only certain values.
22 Part I The Origin of Living Things
1s Orbital
x
x
y
z
Orbital for energy level K:
one spherical orbital (1s)
2s Orbital
2p Orbitals
Composite of
all p orbitals
Orbitals for energy level L:
one spherical orbital (2s) and
three dumbbell-shaped orbitals (2p)
z
y
FIGURE 2.4
Electron orbitals. The lowest energy level or electron shell, which is nearest the nucleus, is level K. It is occupied by a single s orbital,
referred to as 1s. The next highest energy level, L, is occupied by four orbitals: one s orbital (referred to as the 2s orbital) and three p
orbitals (each referred to as a 2p orbital). The four L-level orbitals compactly fill the space around the nucleus, like two pyramids set base-
to-base.
Every atom exhibits a ladder of potential energy values,
rather than a continuous spectrum of possibilities, a discrete
set of orbits at particular distances from the nucleus.
During some chemical reactions, electrons are trans-
ferred from one atom to another. In such reactions, the loss
of an electron is called oxidation, and the gain of an elec-
tron is called reduction (figure 2.6). It is important to real-
ize that when an electron is transferred in this way, it keeps
its energy of position. In organisms, chemical energy is
stored in high-energy electrons that are transferred from
one atom to another in reactions involving oxidation and
reduction.
Because the amount of energy an electron possesses is
related to its distance from the nucleus, electrons that are
the same distance from the nucleus have the same energy,
even if they occupy different orbitals. Such electrons are
said to occupy the same energy level. In a schematic dia-
gram of an atom (figure 2.7), the nucleus is represented as a
small circle and the electron energy levels are drawn as con-
centric rings, with the energy level increasing with distance
from the nucleus. Be careful not to confuse energy levels,
which are drawn as rings to indicate an electron’s energy,
with orbitals, which have a variety of three-dimensional
shapes and indicate an electron’s most likely location.
Electrons orbit a nucleus in paths called orbitals. No
orbital can contain more than two electrons, but many
orbitals may be the same distance from the nucleus and,
thus, contain electrons of the same energy.
Chapter 2 The Nature of Molecules 23
Energy released
Energy
level
3
Energy
level
2
Energy
level
1
–
ML K
Energy
level
1
Energy absorbed
Energy
level
2
Energy
level
3
+
–
+
+
+
+
+ +
MLK
FIGURE 2.5
Atomic energy levels. When an electron
absorbs energy, it moves to higher energy
levels farther from the nucleus. When an
electron releases energy, it falls to lower
energy levels closer to the nucleus.
FIGURE 2.6
Oxidation and reduction. Oxidation is the loss of an electron;
reduction is the gain of an electron.
Oxidation Reduction
H11001
–
H11001
Helium Nitrogen
7H11001
7n
2H11001
2n
KKL
Nucleus
L
M
N
K
Energy le
ve
l
FIGURE 2.7
Electron energy levels for helium and nitrogen. Gold balls
represent the electrons. Each concentric circle represents a
different distance from the nucleus and, thus, a different electron
energy level.
24 Part I The Origin of Living Things
1
H
1
H
3
Li
4
Be
19
K
12
Mg
93
Np
94
Pu
95
Am
96
Cm
97
Bk
98
Cf
99
Es
100
Fm
101
Md
102
No
103
Lr
37
Rb
38
Sr
39
Y
42
Mo
45
Rh
46
Pd
47
Ag
48
Cd
49
In
50
Sn
51
Sb
52
Te
53
I
54
Xe
21
Sc
40
Zr
22
Ti
23
V
24
Cr
25
Mn
27
Co
28
Ni
29
Cu
30
Zn
36
Kr
5
B
6
C
6
C
8
O
2
He
55
Cs
56
Ba
72
Hf
73
Ta
74
W
75
Re
76
Os
77
Ir
78
Pt
79
Au
80
Hg
81
Tl
82
Pb
83
Bi
84
Po
85
At
86
Rn
87
Fr
88
Ra
57
La
89
Ac
104 105 106 107 108 109
58
Ce
59
Pr
60
Nd
61
Pm
62
Sm
63
Eu
64
Gd
65
Tb
66
Dy
67
Ho
68
Er
69
Tm
70
Yb
71
Lu
90
Th
91
Pa
92
U
(Lanthanide series)
(Actinide series)
11
Na
20
Ca
41
Nb
43
Tc
44
Ru
26
Fe
13
Al
31
Ga
32
Ge
14
Si
7
N
15
P
33
As
16
S
35
Br
34
Se
9
F
18
Ar
10
Ne
17
Cl
110
FIGURE 2.8
Periodic table of the elements. In this representation, the frequency of elements that occur in the earth’s crust is indicated by the height
of the block. Elements found in significant amounts in living organisms are shaded in blue.
Kinds of Atoms
There are 92 naturally occurring elements, each with a dif-
ferent number of protons and a different arrangement of
electrons. When the nineteenth-century Russian chemist
Dmitri Mendeleev arranged the known elements in a table
according to their atomic mass (figure 2.8), he discovered
one of the great generalizations in all of science. Mendeleev
found that the elements in the table exhibited a pattern of
chemical properties that repeated itself in groups of eight el-
ements. This periodically repeating pattern lent the table its
name: the periodic table of elements.
The Periodic Table
The eight-element periodicity that Mendeleev found is
based on the interactions of the electrons in the outer en-
ergy levels of the different elements. These electrons are
called valence electrons and their interactions are the
basis for the differing chemical properties of the elements.
For most of the atoms important to life, an outer energy
level can contain no more than eight electrons; the chemi-
cal behavior of an element reflects how many of the eight
positions are filled. Elements possessing all eight elec-
trons in their outer energy level (two for helium) are
inert, or nonreactive; they include helium (He), neon
(Ne), argon (Ar), krypton (Kr), xenon (Xe), and radon
(Rn). In sharp contrast, elements with seven electrons (one
fewer than the maximum number of eight) in their outer
energy level, such as fluorine (F), chlorine (Cl), and
bromine (Br), are highly reactive. They tend to gain the
extra electron needed to fill the energy level. Elements
with only one electron in their outer energy level, such as
lithium (Li), sodium (Na), and potassium (K), are also
very reactive; they tend to lose the single electron in their
outer level.
Mendeleev’s periodic table thus leads to a useful generali-
zation, the octet rule (Latin octo, “eight”) or rule of eight:
atoms tend to establish completely full outer energy levels.
Most chemical behavior can be predicted quite accurately
from this simple rule, combined with the tendency of at-
oms to balance positive and negative charges.
2.2 The atoms of living things are among the smallest.
Distribution of the Elements
Of the 92 naturally occurring elements on earth, only 11 are
found in organisms in more than trace amounts (0.01% or
higher). These 11 elements have atomic numbers less than
21 and, thus, have low atomic masses. Table 2.1 lists the
levels of various elements in the human body; their levels in
other organisms are similar. Inspection of this table suggests
that the distribution of elements in living systems is by no
means accidental. The most common elements inside or-
ganisms are not the elements that are most abundant in the
earth’s crust. For example, silicon, aluminum, and iron con-
stitute 39.2% of the earth’s crust, but they exist in trace
amounts in the human body. On the other hand, carbon at-
oms make up 18.5% of the human body but only 0.03% of
the earth’s crust.
Ninety-two elements occur naturally on earth; only
eleven of them are found in significant amounts in living
organisms. Four of them—oxygen, hydrogen, carbon,
nitrogen—constitute 96.3% of the weight of your body.
Chapter 2 The Nature of Molecules 25
Table 2.1 The Most Common Elements on Earth and Their Distribution in the Human Body
Approximate
Percent of Percent of
Earth’s Crust Human Body
Element Symbol Atomic Number by Weight by Weight Importance or Function
Oxygen
Silicon
Aluminum
Iron
Calcium
Sodium
Potassium
Magnesium
Hydrogen
Manganese
Fluorine
Phosphorus
Carbon
Sulfur
Chlorine
Vanadium
Chromium
Copper
Nitrogen
Boron
Cobalt
Zinc
Selenium
Molybdenum
Tin
Iodine
O
Si
Al
Fe
Ca
Na
K
Mg
H
Mn
F
P
C
S
Cl
V
Cr
Cu
N
B
Co
Zn
Se
Mo
Sn
I
8
14
13
26
20
11
19
12
1
25
9
15
6
16
17
23
24
29
7
5
27
30
34
42
50
53
46.6
27.7
6.5
5.0
3.6
2.8
2.6
2.1
0.14
0.1
0.07
0.07
0.03
0.03
0.01
0.01
0.01
0.01
Trace
Trace
Trace
Trace
Trace
Trace
Trace
Trace
65.0
Trace
Trace
Trace
1.5
0.2
0.4
0.1
9.5
Trace
Trace
1.0
18.5
0.3
0.2
Trace
Trace
Trace
3.3
Trace
Trace
Trace
Trace
Trace
Trace
Trace
Required for cellular respiration;
component of water
Critical component of hemoglobin in
the blood
Component of bones and teeth; trig-
gers muscle contraction
Principal positive ion outside cells;
important in nerve function
Principal positive ion inside cells; im-
portant in nerve function
Critical component of many energy-
transferring enzymes
Electron carrier; component of water
and most organic molecules
Backbone of nucleic acids; important
in energy transfer
Backbone of organic molecules
Component of most proteins
Principal negative ion outside cells
Key component of many enzymes
Component of all proteins and nucleic
acids
Key component of some enzymes
Key component of many enzymes
Component of thyroid hormone
26 Part I The Origin of Living Things
Na
Sodium atom
Sodium ion
Chlorine atom
Chloride ion+ –
Na
+
Cl
Cl
H11002
(a)
FIGURE 2.9
The formation of ionic bonds by sodium chloride. (a) When a sodium atom donates an electron to a chlorine atom, the sodium atom
becomes a positively charged sodium ion, and the chlorine atom becomes a negatively charged chloride ion. (b) Sodium chloride forms a
highly regular lattice of alternating sodium ions and chloride ions.
NaCl crystal
Cl
H11002
Cl
H11002
Cl
H11002
Cl
H11002
Cl
H11002
Na
H11001
Na
H11001
Na
H11001
Na
H11001
(b)
Ionic Bonds Form Crystals
A group of atoms held together by energy in a stable associ-
ation is called a molecule. When a molecule contains atoms
of more than one element, it is called a compound. The
atoms in a molecule are joined by chemical bonds; these
bonds can result when atoms with opposite charges attract
(ionic bonds), when two atoms share one or more pairs of
electrons (covalent bonds), or when atoms interact in other
ways. We will start by examining ionic bonds, which form
when atoms with opposite electrical charges (ions) attract.
A Closer Look at Table Salt
Common table salt, sodium chloride (NaCl), is a lattice of
ions in which the atoms are held together by ionic bonds
(figure 2.9). Sodium has 11 electrons: 2 in the inner energy
level, 8 in the next level, and 1 in the outer (valence) level.
The valence electron is unpaired (free) and has a strong ten-
dency to join with another electron. A stable configuration
can be achieved if the valence electron is lost to another
atom that also has an unpaired electron. The loss of this
electron results in the formation of a positively charged
sodium ion, Na
+
.
The chlorine atom has 17 electrons: 2 in the inner energy
level, 8 in the next level, and 7 in the outer level. Hence, one
of the orbitals in the outer energy level has an unpaired
electron. The addition of another electron to the outer level
fills that level and causes a negatively charged chloride ion,
Cl
–
, to form.
When placed together, metallic sodium and gaseous
chlorine react swiftly and explosively, as the sodium atoms
donate electrons to chlorine, forming Na
+
and Cl
–
ions. Be-
cause opposite charges attract, the Na
+
and Cl
–
remain asso-
ciated in an ionic compound, NaCl, which is electrically
neutral. However, the electrical attractive force holding
NaCl together is not directed specifically between particular
Na
+
and Cl
–
ions, and no discrete sodium chloride mole-
cules form. Instead, the force exists between any one ion and
all neighboring ions of the opposite charge, and the ions ag-
gregate in a crystal matrix with a precise geometry. Such ag-
gregations are what we know as salt crystals. If a salt such as
NaCl is placed in water, the electrical attraction of the water
molecules, for reasons we will point out later in this chapter,
disrupts the forces holding the ions in their crystal matrix,
causing the salt to dissolve into a roughly equal mixture of
free Na
+
and Cl
–
ions.
An ionic bond is an attraction between ions of opposite
charge in an ionic compound. Such bonds are not
formed between particular ions in the compound;
rather, they exist between an ion and all of the
oppositely charged ions in its immediate vicinity.
2.3 Chemical bonds hold molecules together.
Covalent Bonds Build
Stable Molecules
Covalent bonds form when two atoms
share one or more pairs of valence
electrons. Consider hydrogen (H) as an
example. Each hydrogen atom has an
unpaired electron and an unfilled outer
energy level; for these reasons the hy-
drogen atom is unstable. When two
hydrogen atoms are close to each
other, however, each atom’s electron
can orbit both nuclei. In effect, the nu-
clei are able to share their electrons.
The result is a diatomic (two-atom)
molecule of hydrogen gas (figure 2.10).
The molecule formed by the two hy-
drogen atoms is stable for three reasons:
1. It has no net charge. The di-
atomic molecule formed as a result
of this sharing of electrons is not
charged, because it still contains
two protons and two electrons.
2. The octet rule is satisfied.
Each of the two hydrogen atoms
can be considered to have two or-
biting electrons in its outer energy
level. This satisfies the octet rule,
because each shared electron orbits
both nuclei and is included in the
outer energy level of both atoms.
3. It has no free electrons. The
bonds between the two atoms
also pair the two free electrons.
Unlike ionic bonds, covalent bonds
are formed between two specific atoms, giving rise to true,
discrete molecules. While ionic bonds can form regular crys-
tals, the more specific associations made possible by covalent
bonds allow the formation of complex molecular structures.
Covalent Bonds Can Be Very Strong
The strength of a covalent bond depends on the number of
shared electrons. Thus double bonds, which satisfy the oc-
tet rule by allowing two atoms to share two pairs of elec-
trons, are stronger than single bonds, in which only one
electron pair is shared. This means more chemical energy is
required to break a double bond than a single bond. The
strongest covalent bonds are triple bonds, such as those
that link the two nitrogen atoms of nitrogen gas molecules.
Covalent bonds are represented in chemical formulations as
lines connecting atomic symbols, where each line between
two bonded atoms represents the sharing of one pair of
electrons. The structural formulas of hydrogen gas and
oxygen gas are H—H and OH33527O, respectively, while their
molecular formulas are H
2
and O
2
.
Molecules with Several Covalent
Bonds
Molecules often consist of more than
two atoms. One reason that larger mole-
cules may be formed is that a given atom
is able to share electrons with more than
one other atom. An atom that requires
two, three, or four additional electrons
to fill its outer energy level completely
may acquire them by sharing its elec-
trons with two or more other atoms.
For example, the carbon atom (C)
contains six electrons, four of which are
in its outer energy level. To satisfy the
octet rule, a carbon atom must gain ac-
cess to four additional electrons; that is,
it must form four covalent bonds. Be-
cause four covalent bonds may form in
many ways, carbon atoms are found in
many different kinds of molecules.
Chemical Reactions
The formation and breaking of chemi-
cal bonds, the essence of chemistry, is
called a chemical reaction. All chemi-
cal reactions involve the shifting of at-
oms from one molecule or ionic com-
pound to another, without any change
in the number or identity of the atoms.
For convenience, we refer to the origi-
nal molecules before the reaction starts
as reactants, and the molecules result-
ing from the chemical reaction as prod-
ucts. For example:
A — B + C — D → A — C + B + D
reactants products
The extent to which chemical reactions occur is influ-
enced by several important factors:
1. Temperature. Heating up the reactants increases
the rate of a reaction (as long as the temperature isn’t
so high as to destroy the molecules).
2. Concentration of reactants and products. Reac-
tions proceed more quickly when more reactants are
available. An accumulation of products typically
speeds reactions in the reverse direction.
3. Catalysts. A catalyst is a substance that increases the
rate of a reaction. It doesn’t alter the reaction’s equi-
librium between reactants and products, but it does
shorten the time needed to reach equilibrium, often
dramatically. In organisms, proteins called enzymes
catalyze almost every chemical reaction.
A covalent bond is a stable chemical bond formed when
two atoms share one or more pairs of electrons.
Chapter 2 The Nature of Molecules 27
FIGURE 2.10
Hydrogen gas. (a) Hydrogen gas is a
diatomic molecule composed of two
hydrogen atoms, each sharing its electron
with the other. (b) The flash of fire that
consumed the Hindenburg occurred when
the hydrogen gas that was used to inflate the
dirigible combined explosively with oxygen
gas in the air to form water.
H
2
(hydrogen gas)
Covalent bond
+ +
–
–
(a)
(b)
Chemistry of Water
Of all the molecules that are common on earth, only wa-
ter exists as a liquid at the relatively low temperatures that
prevail on the earth’s surface, three-fourths of which is
covered by liquid water (figure 2.11). When life was origi-
nating, water provided a medium in which other molecules
could move around and interact without being held in
place by strong covalent or ionic bonds. Life evolved as a
result of these interactions, and it is still inextricably tied
to water. Life began in water and evolved there for 3 bil-
lion years before spreading to land. About two-thirds of
any organism’s body is composed of water, and no organ-
ism can grow or reproduce in any but a water-rich envi-
ronment. It is no accident that tropical rain forests are
bursting with life, while dry deserts appear almost lifeless
except when water becomes temporarily plentiful, such as
after a rainstorm.
The Atomic Structure of Water
Water has a simple atomic structure. It consists of an oxy-
gen atom bound to two hydrogen atoms by two single cova-
lent bonds (figure 2.12a). The resulting molecule is stable: it
satisfies the octet rule, has no unpaired electrons, and carries
no net electrical charge.
The single most outstanding chemical property of wa-
ter is its ability to form weak chemical associations with
only 5 to 10% of the strength of covalent bonds. This
property, which derives directly from the structure of wa-
ter, is responsible for much of the organization of living
chemistry.
The chemistry of life is water chemistry. The way in
which life first evolved was determined in large part by
the chemical properties of the liquid water in which
that evolution occurred.
28 Part I The Origin of Living Things
FIGURE 2.11
Water takes many forms. As a liquid, water fills our rivers and runs down over the land to the sea. (a) The iceberg on which the penguins
are holding their meeting was formed in Antarctica from a huge block of ice that broke away into the ocean water. (b) When water cools
below 0°C, it forms beautiful crystals, familiar to us as snow and ice. However, water is not always plentiful. (c) At Badwater, in Death
Valley, California, there is no hint of water except for the broken patterns of dried mud.
(a) (b)
(c)
FIGURE 2.12
Water has a simple molecular structure. (a) Each molecule is
composed of one oxygen atom and two hydrogen atoms. The
oxygen atom shares one electron with each hydrogen atom. (b)
The greater electronegativity of the oxygen atom makes the water
molecule polar: water carries two partial negative charges (δ
–
) near
the oxygen atom and two partial positive charges (δ
+
), one on each
hydrogen atom.
δ
–
δ
–
δ
–
δ
–
δ
+
δ
+
δ
+
δ
+
104.5°
Oxygen
Hydrogen
Hydrogen
Bohr model Ball-and-stick model
H
H
8+
8n
+
+
O
(a) (b)
2.4 Water is the cradle of life.
Water Atoms Act Like Tiny
Magnets
Both the oxygen and the hydrogen atoms attract
the electrons they share in the covalent bonds of
a water molecule; this attraction is called elec-
tronegativity. However, the oxygen atom is
more electronegative than the hydrogen atoms,
so it attracts the electrons more strongly than do
the hydrogen atoms. As a result, the shared
electrons in a water molecule are far more
likely to be found near the oxygen nucleus than
near the hydrogen nuclei. This stronger attrac-
tion for electrons gives the oxygen atom two
partial negative charges (δ
–
), as though the elec-
tron cloud were denser near the oxygen atom
than around the hydrogen atoms. Because the
water molecule as a whole is electrically neu-
tral, each hydrogen atom carries a partial positive charge (δ
+
).
The Greek letter delta (δ) signifies a partial charge, much
weaker than the full unit charge of an ion.
What would you expect the shape of a water molecule to
be? Each of water’s two covalent bonds has a partial charge
at each end, δ
–
at the oxygen end and δ
+
at the hydrogen end.
The most stable arrangement of these charges is a tetrahe-
dron, in which the two negative and two positive charges are
approximately equidistant from one another (figure 2.12b).
The oxygen atom lies at the center of the tetrahedron, the
hydrogen atoms occupy two of the apexes, and the partial
negative charges occupy the other two apexes. This results
in a bond angle of 104.5° between the two covalent oxygen-
hydrogen bonds. (In a regular tetrahedron, the bond angles
would be 109.5°; in water, the partial negative charges occu-
py more space than the hydrogen atoms, and, therefore, they
compress the oxygen-hydrogen bond angle slightly.)
The water molecule, thus, has distinct “ends,” each with a
partial charge, like the two poles of a magnet. (These partial
charges are much less than the unit charges of ions, how-
ever.) Molecules that exhibit charge separation are called
polar molecules because of their magnet-like poles, and
water is one of the most polar molecules known. The polarity
of water underlies its chemistry and the chemistry of life.
Polar molecules interact with one another, as the δ
–
of
one molecule is attracted to the δ
+
of another. Because many
of these interactions involve hydrogen atoms, they are
called hydrogen bonds (figure 2.13). Each hydrogen bond
is individually very weak and transient, lasting on average
only
100,000
1
,000,000
second (10
–11
sec). However, the cumula-
tive effects of large numbers of these bonds can be enor-
mous. Water forms an abundance of hydrogen bonds,
which are responsible for many of its important physical
properties (table 2.2).
The water molecule is very polar, with ends that exhibit
partial positive and negative charges. Opposite charges
attract, forming weak linkages called hydrogen bonds.
Chapter 2 The Nature of Molecules 29
Hydrogen atom
Hydrogen bond
An organic molecule
Oxygen atom
δ
–
δ
+
FIGURE 2.13
Structure of a hydrogen bond.
Table 2.2 The Properties of Water
Property Explanation Example of Benefit to Life
Cohesion
High specific heat
High heat of
vaporization
Lower density
of ice
High polarity
Hydrogen bonds hold water molecules together
Hydrogen bonds absorb heat when they break, and release
heat when they form, minimizing temperature changes
Many hydrogen bonds must be broken for water to evapo-
rate
Water molecules in an ice crystal are spaced relatively far
apart because of hydrogen bonding
Polar water molecules are attracted to ions and polar com-
pounds, making them soluble
Leaves pull water upward from the
roots; seeds swell and germinate
Water stabilizes the temperature of
organisms and the environment
Evaporation of water cools body
surfaces
Because ice is less dense than water,
lakes do not freeze solid
Many kinds of molecules can move
freely in cells, permitting a diverse
array of chemical reactions
Water Clings to Polar Molecules
The polarity of water causes it to be attracted to other polar
molecules. When the other molecules are also water, the at-
traction is referred to as cohesion. When the other mole-
cules are of a different substance, the attraction is called ad-
hesion. It is because water is cohesive that it is a liquid, and
not a gas, at moderate temperatures.
The cohesion of liquid water is also responsible for its
surface tension. Small insects can walk on water (figure
2.14) because at the air-water interface all of the hydrogen
bonds in water face downward, causing the molecules of the
water surface to cling together. Water is adhesive to any
substance with which it can form hydrogen bonds. That is
why substances containing polar molecules get “wet” when
they are immersed in water, while those that are composed
of nonpolar molecules (such as oils) do not.
The attraction of water to substances like glass with sur-
face electrical charges is responsible for capillary action: if a
glass tube with a narrow diameter is lowered into a beaker
of water, water will rise in the tube above the level of the
water in the beaker, because the adhesion of water to the
glass surface, drawing it upward, is stronger than the force
of gravity, drawing it down. The narrower the tube, the
greater the electrostatic forces between the water and the
glass, and the higher the water rises (figure 2.15).
Water Stores Heat
Water moderates temperature through two properties: its
high specific heat and its high heat of vaporization. The
temperature of any substance is a measure of how rapidly
its individual molecules are moving. Because of the many
hydrogen bonds that water molecules form with one anoth-
er, a large input of thermal energy is required to break
these bonds before the individual water molecules can be-
gin moving about more freely and so have a higher temper-
ature. Therefore, water is said to have a high specific heat,
which is defined as the amount of heat that must be ab-
sorbed or lost by 1 gram of a substance to change its tem-
perature by 1 degree Celsius (°C). Specific heat measures
the extent to which a substance resists changing its temper-
ature when it absorbs or loses heat. Because polar substanc-
es tend to form hydrogen bonds, and energy is needed to
break these bonds, the more polar a substance is, the higher
is its specific heat. The specific heat of water (1 calo-
rie/gram/°C) is twice that of most carbon compounds and
nine times that of iron. Only ammonia, which is more
polar than water and forms very strong hydrogen bonds,
has a higher specific heat than water (1.23
calories/gram/°C). Still, only 20% of the hydrogen bonds
are broken as water heats from 0° to 100°C.
Because of its high specific heat, water heats up more
slowly than almost any other compound and holds its tem-
perature longer when heat is no longer applied. This char-
acteristic enables organisms, which have a high water con-
tent, to maintain a relatively constant internal temperature.
The heat generated by the chemical reactions inside cells
would destroy the cells, if it were not for the high specific
heat of the water within them.
A considerable amount of heat energy (586 calories) is re-
quired to change 1 gram of liquid water into a gas. Hence,
water also has a high heat of vaporization. Because the
transition of water from a liquid to a gas requires the input
of energy to break its many hydrogen bonds, the evapora-
tion of water from a surface causes cooling of that surface.
Many organisms dispose of excess body heat by evaporative
cooling; for example, humans and many other vertebrates
sweat.
At low temperatures, water molecules are locked into a
crystal-like lattice of hydrogen bonds, forming the solid we
call ice (figure 2.16). Interestingly, ice is less dense than liquid
water because the hydrogen bonds in ice space the water
molecules relatively far apart. This unusual feature enables
icebergs to float. Were it otherwise, ice would cover nearly all
bodies of water, with only shallow surface melting annually.
30 Part I The Origin of Living Things
FIGURE 2.14
Cohesion. Some insects, such as this water strider, literally walk
on water. In this photograph you can see the dimpling the insect’s
feet make on the water as its weight bears down on the surface.
Because the surface tension of the water is greater than the force
that one foot brings to bear, the strider glides atop the surface of
the water rather than sinking.
FIGURE 2.15
Capillary action. Capillary action
causes the water within a narrow tube
to rise above the surrounding water;
the adhesion of the water to the glass
surface, which draws water upward, is
stronger than the force of gravity,
which tends to draw it down. The
narrower the tube, the greater the
surface area available for adhesion for a
given volume of water, and the higher
the water rises in the tube.
Water Is a Powerful Solvent
Water is an effective solvent because of its ability to form
hydrogen bonds. Water molecules gather closely around
any substance that bears an electrical charge, whether that
substance carries a full charge (ion) or a charge separation
(polar molecule). For example, sucrose (table sugar) is
composed of molecules that contain slightly polar hydroxyl
(OH) groups. A sugar crystal dissolves rapidly in water be-
cause water molecules can form hydrogen bonds with indi-
vidual hydroxyl groups of the sucrose molecules. There-
fore, sucrose is said to be soluble in water. Every time a
sucrose molecule dissociates or breaks away from the crys-
tal, water molecules surround it in a cloud, forming a hy-
dration shell and preventing it from associating with oth-
er sucrose molecules. Hydration shells also form around
ions such as Na
+
and Cl
–
(figure 2.17).
Water Organizes Nonpolar Molecules
Water molecules always tend to form the maximum possi-
ble number of hydrogen bonds. When nonpolar molecules
such as oils, which do not form hydrogen bonds, are placed
in water, the water molecules act to exclude them. The
nonpolar molecules are forced into association with one an-
other, thus minimizing their disruption of the hydrogen
bonding of water. In effect, they shrink from contact with
water and for this reason they are referred to as hydropho-
bic (Greek hydros, “water” and phobos, “fearing”). In con-
trast, polar molecules, which readily form hydrogen bonds
with water, are said to be hydrophilic (“water-loving”).
The tendency of nonpolar molecules to aggregate in wa-
ter is known as hydrophobic exclusion. By forcing the hy-
drophobic portions of molecules together, water causes
these molecules to assume particular shapes. Different mo-
lecular shapes have evolved by alteration of the location
and strength of nonpolar regions. As you will see, much of
the evolution of life reflects changes in molecular shape
that can be induced in just this way.
Water molecules, which are very polar, cling to one
another, so that it takes considerable energy to separate
them. Water also clings to other polar molecules,
causing them to be soluble in water solution, but water
tends to exclude nonpolar molecules.
Chapter 2 The Nature of Molecules 31
Water
molecules
Stable
hydrogen bonds
Unstable hydrogen bonds
(a) Liquid water (b) Ice
FIGURE 2.16
The role of hydrogen bonds in an ice crystal. (a) In liquid
water, hydrogen bonds are not stable and constantly break and re-
form. (b) When water cools below 0°C, the hydrogen bonds are
more stable, and a regular crystalline structure forms in which the
four partial charges of one water molecule interact with the
opposite charges of other water molecules. Because water forms a
crystal latticework, ice is less dense than liquid water and floats. If
it did not, inland bodies of water far from the earth’s equator
might never fully thaw.
Hydration shells
Water
molecules
Salt
crystal
Na
H11001
Cl
H11002
Cl
H11002
Cl
H11002
Cl
H11002
Na
H11001
Na
H11001
Na
H11001
FIGURE 2.17
Why salt dissolves in water. When a crystal of table salt dissolves
in water, individual Na
+
and Cl
–
ions break away from the salt
lattice and become surrounded by water molecules. Water
molecules orient around Cl
–
ions so that their partial positive poles
face toward the negative Cl
–
ion; water molecules surrounding Na
+
ions orient in the opposite way, with their partial negative poles
facing the positive Na
+
ion. Surrounded by hydration shells, Na
+
and Cl
–
ions never reenter the salt lattice.
Water Ionizes
The covalent bonds within a water molecule sometimes
break spontaneously. In pure water at 25°C, only 1 out of
every 550 million water molecules undergoes this process.
When it happens, one of the protons (hydrogen atom nu-
clei) dissociates from the molecule. Because the dissociated
proton lacks the negatively charged electron it was sharing
in the covalent bond with oxygen, its own positive charge is
no longer counterbalanced, and it becomes a positively
charged ion, H
+
. The rest of the dissociated water molecule,
which has retained the shared electron from the covalent
bond, is negatively charged and forms a hydroxide ion
(OH
–
). This process of spontaneous ion formation is called
ionization:
H
2
O → OH
–
+H
+
water hydroxide ion hydrogenion (proton)
At 25°C, a liter of water contains
10,000
1
,000
(or
10
–7
) mole of H
+
ions. (A mole is defined as the weight in
grams that corresponds to the summed atomic masses of all
of the atoms in a molecule. In the case ofH
+
, the atomic
mass is 1, and a mole of H
+
ions would weigh 1 gram. One
mole of any substance always contains 6.02 × 10
23
mole-
cules of the substance.) Therefore, the molar concentra-
tion of hydrogen ions (represented as [H
+
]) in pure water is
10
–7
mole/liter. Actually, the hydrogen ion usually associ-
ates with another water molecule to form a hydronium
(H
3
O
+
) ion.
pH
A more convenient way to express the hydrogen ion concen-
tration of a solution is to use the pH scale (figure 2.18).
This scale defines pH as the negative logarithm of the hy-
drogen ion concentration in the solution:
pH = –log [H
+
]
Because the logarithm of the hydrogen ion concentration is
simply the exponent of the molar concentration of H
+
, the
pH equals the exponent times –1. Thus, pure water, with an
[H
+
] of 10
–7
mole/liter, has a pH of 7. Recall that for every
H
+
ion formed when water dissociates, an OH
–
ion is also
formed, meaning that the dissociation of water produces H
+
and OH
–
in equal amounts. Therefore, a pH value of 7 indi-
cates neutrality—a balance between H
+
and OH
–
—on the
pH scale.
Note that the pH scale is logarithmic, which means that a
difference of 1 on the scale represents a tenfold change in
hydrogen ion concentration. This means that a solution
with a pH of 4 has 10 times the concentration of H
+
than is
present in one with a pH of 5.
Acids. Any substance that dissociates in water to increase
the concentration of H
+
ions is called an acid. Acidic solu-
tions have pH values below 7. The stronger an acid is, the
more H
+
ions it produces and the lower its pH. For exam-
ple, hydrochloric acid (HCl), which is abundant in your
stomach, ionizes completely in water. This means that 10
–1
mole per liter of HCl will dissociate to form 10
–1
mole per
liter of H
+
ions, giving the solution a pH of 1. The pH of
champagne, which bubbles because of the carbonic acid
dissolved in it, is about 2.
Bases. A substance that combines with H
+
ions when dis-
solved in water is called a base. By combining with H
+
ions,
a base lowers the H
+
ion concentration in the solution.
Basic (or alkaline) solutions, therefore, have pH values
above 7. Very strong bases, such as sodium hydroxide
(NaOH), have pH values of 12 or more.
32 Part I The Origin of Living Things
10
H110021
H
+
Ion
Concentration
Examples of
Solutions
Hydrochloric acid
Stomach acid
1
pH Value
10
H110022
Lemon juice2
10
H110023
Vinegar, cola, beer3
10
H110024
Tomatoes4
10
H110025
Black coffee
Normal rainwater
5
10
H110026
Urine
Saliva
6
10
H110027
Pure water
Blood
7
10
H110028
Seawater8
10
H110029
Baking soda9
10
H1100210
Great Salt Lake10
10
H1100211
Household ammonia11
10
H1100212
Household bleach
12
10
H1100213
Oven cleaner
13
10
H1100214
Sodium hydroxide14
FIGURE 2.18
The pH scale. The pH value of a solution indicates its
concentration of hydrogen ions. Solutions with a pH less than 7
are acidic, while those with a pH greater than 7 are basic. The
scale is logarithmic, so that a pH change of 1 means a tenfold
change in the concentration of hydrogen ions. Thus, lemon juice
is 100 times more acidic than tomato juice, and seawater is 10
times more basic than pure water, which has a pH of 7.
Buffers
The pH inside almost all living cells, and in the fluid sur-
rounding cells in multicellular organisms, is fairly close to 7.
Most of the biological catalysts (enzymes) in living systems
are extremely sensitive to pH; often even a small change in
pH will alter their shape, thereby disrupting their activities
and rendering them useless. For this reason it is important
that a cell maintain a constant pH level.
Yet the chemical reactions of life constantly produce acids
and bases within cells. Furthermore, many animals eat sub-
stances that are acidic or basic; cola, for example, is a strong
(although dilute) acidic solution. Despite such variations in
the concentrations of H
+
and OH
–
, the pH of an organism is
kept at a relatively constant level by buffers (figure 2.19).
A buffer is a substance that acts as a reservoir for hy-
drogen ions, donating them to the solution when their con-
centration falls and taking them from the solution when
their concentration rises. What sort of substance will act in
this way? Within organisms, most buffers consist of pairs of
substances, one an acid and the other a base. The key buffer
in human blood is an acid-base pair consisting of carbonic
acid (acid) and bicarbonate (base). These two substances in-
teract in a pair of reversible reactions. First, carbon dioxide
(CO
2
) and H
2
O join to form carbonic acid (H
2
CO
3
), which
in a second reaction dissociates to yield bicarbonate ion
(HCO
3
–
) and H
+
(figure 2.20). If some acid or other sub-
stance adds H
+
ions to the blood, the HCO
3
–
ions act as a
base and remove the excess H
+
ions by forming H
2
CO
3
.
Similarly, if a basic substance removes H
+
ions from the
blood, H
2
CO
3
dissociates, releasing more H
+
ions into the
blood. The forward and reverse reactions that interconvert
H
2
CO
3
and HCO
3
–
thus stabilize the blood’s pH.
The reaction of carbon dioxide and water to form car-
bonic acid is important because it permits carbon, essential
to life, to enter water from the air. As we will discuss in
chapter 4, biologists believe that life first evolved in the
early oceans. These oceans were rich in carbon because of
the reaction of carbon dioxide with water.
In a condition called blood acidosis, human blood, which
normally has a pH of about 7.4, drops 0.2 to 0.4 points on the
pH scale. This condition is fatal if not treated immediately.
The reverse condition, blood alkalosis, involves an increase in
blood pH of a similar magnitude and is just as serious.
The pH of a solution is the negative logarithm of the H
+
ion concentration in the solution. Thus, low pH values
indicate high H
+
concentrations (acidic solutions), and
high pH values indicate low H
+
concentrations (basic
solutions). Even small changes in pH can be harmful to
life.
Chapter 2 The Nature of Molecules 33
10
0
1
2
3
4
5
6
7
8
9
3
Amount of base added
Buffering
range
pH
452
FIGURE 2.19
Buffers minimize changes in pH. Adding a base to a solution
neutralizes some of the acid present, and so raises the pH. Thus,
as the curve moves to the right, reflecting more and more base, it
also rises to higher pH values. What a buffer does is to make the
curve rise or fall very slowly over a portion of the pH scale, called
the “buffering range” of that buffer.
FIGURE 2.20
Buffer formation. Carbon dioxide and water combine chemically to form carbonic acid (H
2
CO
3
). The acid then dissociates in water,
freeing H
+
ions. This reaction makes carbonated beverages acidic, and produced the carbon-rich early oceans that cradled life.
H
2
O
Water
CO
2
Carbon dioxide
H
2
CO
3
Carbonic acid
HCO
H11002
3
Bicarbonate
ion
H
H11001
Hydrogen
ion
–
+
+
+
? The smallest stable particles of matter are protons, neutrons,
and electrons, which associate to form atoms.
? The core, or nucleus, of an atom consists of protons and
neutrons; the electrons orbit around the nucleus in a cloud.
The farther an electron is from the nucleus, the faster it
moves and the more energy it possesses.
? The chemical behavior of an atom is largely determined by
the distribution of its electrons and in particular by the num-
ber of electrons in its outermost (highest) energy level.
There is a strong tendency for atoms to have a completely
filled outer level; electrons are lost, gained, or shared until
this condition is reached.
2.2 The atoms of living things are among the smallest.
34 Part I The Origin of Living Things
Chapter 2
Summary Questions Media Resources
2.1 Atoms are nature’s building material.
1. An atom of nitrogen has 7
protons and 7 neutrons. What is
its atomic number? What is its
atomic mass? How many elec-
trons does it have?
2. How do the isotopes of a sin-
gle element differ from each
other?
3. The half-life of radium-226 is
1620 years. If a sample of mate-
rial contains 16 milligrams of ra-
dium-226, how much will it con-
tain in 1620 years? How much
will it contain in 3240 years?
How long will it take for the
sample to contain 1 milligram of
radium-226?
? More than 95% of the weight of an organism consists
of oxygen, hydrogen, carbon, and nitrogen, all of
which form strong covalent bonds with one another.
? Ionic bonds form when electrons transfer from one
atom to another, and the resulting oppositely charged
ions attract one another.
? Covalent bonds form when two atoms share elec-
trons. They are responsible for the formation of most
biologically important molecules.
4. What is the octet rule, and
how does it affect the chemical
behavior of atoms?
5. What is the difference be-
tween an ionic bond and a cova-
lent bond? Give an example of
each.
2.3 Chemical bonds hold molecules together.
2.4 Water is the cradle of life.
? The chemistry of life is the chemistry of water (H
2
O).
The central oxygen atom in water attracts the elec-
trons it shares with the two hydrogen atoms. This
charge separation makes water a polar molecule.
? A hydrogen bond is formed between the partial posi-
tive charge of a hydrogen atom in one molecule and
the partial negative charge of another atom, either in
another molecule or in a different portion of the same
molecule.
? Water is cohesive and adhesive, has a great capacity
for storing heat, is a good solvent for other polar
molecules, and tends to exclude nonpolar molecules.
? The H
+
concentration in a solution is expressed by
the pH scale, in which pH equals the negative loga-
rithm of the H
+
concentration.
6. What types of atoms partici-
pate in the formation of hydro-
gen bonds? How do hydrogen
bonds contribute to water’s high
specific heat?
7. What types of molecules are
hydrophobic? What types are
hydrophilic? Why do these two
types of molecules behave differ-
ently in water?
8. What is the pH of a solution
that has a hydrogen ion concen-
tration of 10
–3
mole/liter?
Would such a solution be acidic
or basic?
? Atomic Structure
? Basic Chemistry
? Atoms
? Bonds
? Ionic Bonds
? Bonds
? Water
? ph Scale
http://www.mhhe.com/raven6e http://www.biocourse.com
35
3
The Chemical Building
Blocks of Life
Concept Outline
3.1 Molecules are the building blocks of life.
The Chemistry of Carbon. Because individual carbon
atoms can form multiple covalent bonds, organic molecules
can be quite complex.
3.2 Proteins perform the chemistry of the cell.
The Many Functions of Proteins. Proteins can be cata-
lysts, transporters, supporters, and regulators.
Amino Acids Are the Building Blocks of Proteins. Proteins
are long chains of various combinations of amino acids.
A Protein’s Function Depends on the Shape of the
Molecule. A protein’s shape is determined by its amino acid
sequence.
How Proteins Fold Into Their Functional Shape. The
distribution of nonpolar amino acids along a protein chain
largely determines how the protein folds.
How Proteins Unfold. When conditions such as pH or
temperature fluctuate, proteins may denature or unfold.
3.3 Nucleic acids store and transfer genetic information.
Information Molecules. Nucleic acids store information in
cells. RNA is a single-chain polymer of nucleotides, while
DNA possesses two chains twisted around each other.
3.4 Lipids make membranes and store energy.
Phospholipids Form Membranes. The spontaneous ag-
gregation of phospholipids in water is responsible for the
formation of biological membranes.
Fats and Other Kinds of Lipids. Organisms utilize a wide
variety of water-insoluble molecules.
Fats as Food. Fats are very efficient energy storage
molecules because of their high proportion of C—H bonds.
3.5 Carbohydrates store energy and provide building
materials.
Simple Carbohydrates. Sugars are simple carbohydrates,
often consisting of six-carbon rings.
Linking Sugars Together. Sugars can be linked together to
form long polymers, or polysaccharides.
Structural Carbohydrates. Structural carbohydrates like
cellulose are chains of sugars linked in a way that enzymes
cannot easily attack.
M
olecules are extremely small compared with the fa-
miliar world we see about us. Imagine: there are
more water molecules in a cup than there are stars in the
sky. Many other molecules are gigantic, compared with wa-
ter, consisting of thousands of atoms. These atoms are or-
ganized into hundreds of smaller molecules that are linked
together into long chains (figure 3.1). These enormous
molecules, almost always synthesized by living things, are
called macromolecules. As we shall see, there are four gen-
eral types of macromolecules, the basic chemical building
blocks from which all organisms are assembled.
FIGURE 3.1
Computer-generated model of a macromolecule. Pictured is
an enzyme responsible for releasing energy from sugar. This
complex molecule consists of hundreds of different amino acids
linked into chains that form the characteristic coils and folds seen
here.
Biological Macromolecules
Some organic molecules in organisms are small and sim-
ple, containing only one or a few functional groups. Oth-
ers are large complex assemblies called macromolecules.
In many cases, these macromolecules are polymers, mole-
cules built by linking together a large number of small,
similar chemical subunits, like railroad cars coupled to
form a train. For example, complex carbohydrates like
starch are polymers of simple ring-shaped sugars, pro-
teins are polymers of amino acids, and nucleic acids
(DNA and RNA) are polymers of nucleotides. Biological
macromolecules are traditionally grouped into four major
categories: proteins, nucleic acids, lipids, and carbohy-
drates (table 3.1).
36 Part I The Origin of Living Things
The Chemistry of Carbon
In chapter 2 we discussed how atoms combine to form
molecules. In this chapter, we will focus on organic mole-
cules, those chemical compounds that contain carbon. The
frameworks of biological molecules consist predominantly
of carbon atoms bonded to other carbon atoms or to atoms
of oxygen, nitrogen, sulfur or hydrogen. Because carbon
atoms possess four valence electrons and so can form four
covalent bonds, molecules containing carbon can form
straight chains, branches, or even rings. As you can imag-
ine, all of these possibilities generate an immense range of
molecular structures and shapes.
Organic molecules consisting only of carbon and hydro-
gen are called hydrocarbons. Covalent bonds between car-
bon and hydrogen are energy-rich. We use hydrocarbons
from fossil fuels as a primary source of energy today.
Propane gas, for example, is a hydrocarbon consisting of a
chain of three carbon atoms, with eight hydrogen atoms
bound to it:
H H H
|||
H—C—C—C—H
| ||
H H H
Because carbon-hydrogen covalent bonds store consider-
able energy, hydrocarbons make good fuels. Gasoline, for
example, is rich in hydrocarbons.
Functional Groups
Carbon and hydrogen atoms both have very similar elec-
tronegativities, so electrons in C—C and C—H bonds are
evenly distributed, and there are no significant differences
in charge over the molecular surface. For this reason, hy-
drocarbons are nonpolar. Most organic molecules that are
produced by cells, however, also contain other atoms. Be-
cause these other atoms often have different electronegativ-
ities, molecules containing them exhibit regions of positive
or negative charge, and so are polar. These molecules can
be thought of as a C—H core to which specific groups of
atoms called functional groups are attached. For example,
a hydrogen atom bonded to an oxygen atom (—OH) is a
functional group called a hydroxyl group.
Functional groups have definite chemical properties that
they retain no matter where they occur. The hydroxyl
group, for example, is polar, because its oxygen atom, being
very electronegative, draws electrons toward itself (as we
saw in chapter 2). Figure 3.2 illustrates the hydroxyl group
and other biologically important functional groups. Most
chemical reactions that occur within organisms involve the
transfer of a functional group as an intact unit from one
molecule to another.
3.1 Molecules are the building blocks of life.
Hydroxyl
Carbonyl
Carboxyl
Amino
Sulfhydryl
Phosphate
Methyl
Carbohydrates,
alcohols
Amino acids,
vinegar
Group Structural
Formula
Ball-and-
Stick Model
Found In:
Formaldehyde
Ammonia
Proteins,
rubber
Phospholipids,
nucleic acids,
ATP
Methane
gas
HS
O
–
P
O
–
O
O
HC
H
H
OH
O
OH
C
H
H
N
C
O
H
H
O
O
–
PO
H
N
S
H
O
O
–
H
H
C
H
H
O
C
O
O
C
FIGURE 3.2
The primary functional chemical groups. These groups tend to
act as units during chemical reactions and confer specific chemical
properties on the molecules that possess them. Amino groups, for
example, make a molecule more basic, while carboxyl groups
make a molecule more acidic.
Building Macromolecules
Although the four categories of macromolecules contain dif-
ferent kinds of subunits, they are all assembled in the same
fundamental way: to form a covalent bond between two sub-
unit molecules, an —OH group is removed from one sub-
unit and a hydrogen atom (H) is removed from the other
(figure 3.3a). This condensation reaction is called a dehy-
dration synthesis, because the removal of the —OH group
and H during the synthesis of a new molecule in effect con-
stitutes the removal of a molecule of water (H
2
O). For every
subunit that is added to a macromolecule, one water mole-
cule is removed. Energy is required to break the chemical
bonds when water is extracted from the subunits, so cells
must supply energy to assemble macromolecules. These and
other biochemical reactions require that the reacting sub-
stances be held close together and that the correct chemical
bonds be stressed and broken. This process of positioning
and stressing, termed catalysis, is carried out in cells by a
special class of proteins known as enzymes.
Cells disassemble macromolecules into their constituent
subunits by performing reactions that are essentially the re-
verse of dehydration—a molecule of water is added instead
of removed (figure 3.3b). In this process, which is called
hydrolysis (Greek hydro, “water” + lyse, “break”), a hydro-
gen atom is attached to one subunit and a hydroxyl group
to the other, breaking a specific covalent bond in the
macromolecule. Hydrolytic reactions release the energy
that was stored in the bonds that were broken.
Polymers are large molecules consisting of long chains
of similar subunits joined by dehydration reactions. In a
dehydration reaction, a hydroxyl (—OH) group is
removed from one subunit and a hydrogen atom (H) is
removed from the other.
Chapter 3 The Chemical Building Blocks of Life 37
Table 3.1 Macromolecules
Macromolecule Subunit Function Example
PROTEINS
Globular
Structural
NUCLEIC ACIDS
DNA
RNA
LIPIDS
Fats
Phospholipids
Prostaglandins
Steroids
Terpenes
CARBOHYDRATES
Starch, glycogen
Cellulose
Chitin
Hemoglobin
Hair; silk
Chromosomes
Messenger RNA
Butter; corn oil; soap
Lecithin
Prostaglandin E (PGE)
Cholesterol; estrogen
Carotene; rubber
Potatoes
Paper; strings of celery
Crab shells
Amino acids
Amino acids
Nucleotides
Nucleotides
Glycerol and three fatty acids
Glycerol, two fatty acids,
phosphate, and polar R groups
Five-carbon rings with two
nonpolar tails
Four fused carbon rings
Long carbon chains
Glucose
Glucose
Modified glucose
Catalysis; transport
Support
Encodes genes
Needed for gene expression
Energy storage
Cell membranes
Chemical messengers
Membranes; hormones
Pigments; structural
Energy storage
Cell walls
Structural support
H
2
O
H
2
O
HO
HO H
HO H
HHHO
Energy
Dehydration synthesis
HO H HHO
Energy
Hydrolysis
(a)
(b)
FIGURE 3.3
Making and breaking
macromolecules.
(a) Biological
macromolecules are
polymers formed by
linking subunits
together. The
covalent bond
between the subunits
is formed by
dehydration synthesis,
an energy-requiring
process that creates a
water molecule for
every bond formed. (b)
Breaking the bond
between subunits
requires the returning
of a water molecule
with a subsequent
release of energy, a
process called
hydrolysis.
The Many Functions of Proteins
We will begin our discussion of macromolecules that make
up the bodies of organisms with proteins (see table 3.1). The
proteins within living organisms are immensely diverse in
structure and function (table 3.2 and figure 3.4).
1. Enzyme catalysis. We have already encountered
one class of proteins, enzymes, which are biological
catalysts that facilitate specific chemical reactions. Be-
cause of this property, the appearance of enzymes was
one of the most important events in the evolution of
life. Enzymes are globular proteins, with a three-
dimensional shape that fits snugly around the chemi-
cals they work on, facilitating chemical reactions by
stressing particular chemical bonds.
2. Defense. Other globular proteins use their shapes
to “recognize” foreign microbes and cancer cells.
These cell surface receptors form the core of the
body’s hormone and immune systems.
3. Transport. A variety of globular proteins transport
specific small molecules and ions. The transport pro-
tein hemoglobin, for example, transports oxygen in
the blood, and myoglobin, a similar protein, transports
oxygen in muscle. Iron is transported in blood by the
protein transferrin.
38 Part I The Origin of Living Things
3.2 Proteins perform the chemistry of the cell.
Table 3.2 The Many Functions of Proteins
Function Class of Protein Examples Use
Metabolism (Catalysis)
Defense
Cell recognition
Transport throughout body
Membrane transport
Structure/Support
Motion
Osmotic regulation
Regulation of gene action
Regulation of body functions
Storage
Enzymes
Immunoglobulins
Toxins
Cell surface antigens
Globins
Transporters
Fibers
Muscle
Albumin
Repressors
Hormones
Ion binding
Hydrolytic enzymes
Proteases
Polymerases
Kinases
Antibodies
Snake venom
MHC proteins
Hemoglobin
Myoglobin
Cytochromes
Sodium-potassium pump
Proton pump
Anion channels
Collagen
Keratin
Fibrin
Actin
Myosin
Serum albumin
lac repressor
Insulin
Vasopressin
Oxytocin
Ferritin
Casein
Calmodulin
Cleave polysaccharides
Break down proteins
Produce nucleic acids
Phosphorylate sugars and
proteins
Mark foreign proteins for
elimination
Block nerve function
“Self” recognition
Carries O
2
and CO
2
in blood
Carries O
2
and CO
2
in muscle
Electron transport
Excitable membranes
Chemiosmosis
Transport Cl– ions
Cartilage
Hair, nails
Blood clot
Contraction of muscle fibers
Contraction of muscle fibers
Maintains osmotic concentration
of blood
Regulates transcription
Controls blood glucose levels
Increases water retention by
kidneys
Regulates uterine contractions
and milk production
Stores iron, especially in spleen
Stores ions in milk
Binds calcium ions
4. Support. Fibrous, or threadlike, proteins play struc-
tural roles; these structural proteins (see figure 3.4) in-
clude keratin in hair, fibrin in blood clots, and col-
lagen, which forms the matrix of skin, ligaments,
tendons, and bones and is the most abundant protein
in a vertebrate body.
5. Motion. Muscles contract through the sliding mo-
tion of two kinds of protein filament: actin and myo-
sin. Contractile proteins also play key roles in the
cell’s cytoskeleton and in moving materials within
cells.
6. Regulation. Small proteins called hormones serve
as intercellular messengers in animals. Proteins also
play many regulatory roles within the cell, turning on
and shutting off genes during development, for exam-
ple. In addition, proteins also receive information, act-
ing as cell surface receptors.
Proteins carry out a diverse array of functions, including
catalysis, defense, transport of substances, motion, and
regulation of cell and body functions.
Chapter 3 The Chemical Building Blocks of Life 39
(a) (b)
(c) (d) (e)
FIGURE 3.4
Some of the more common structural proteins. (a) Collagen: strings of a tennis racket from gut tissue; (b) fibrin: scanning electron
micrograph of a blood clot (3000×); (c) keratin: a peacock feather; (d) silk: a spider’s web; (e) keratin: human hair.
Amino Acids Are the Building Blocks
of Proteins
Although proteins are complex and versatile molecules, they
are all polymers of only 20 amino acids, in a specific order.
Many scientists believe amino acids were among the first
molecules formed in the early earth. It seems highly likely
that the oceans that existed early in the history of the earth
contained a wide variety of amino acids.
Amino Acid Structure
An amino acid is a molecule containing an amino group
(—NH
2
), a carboxyl group (—COOH), and a hydrogen
atom, all bonded to a central carbon atom:
R
|
H
2
N—C—COOH
|
H
Each amino acid has unique chemical properties deter-
mined by the nature of the side group (indicated by R) cova-
lently bonded to the central carbon atom. For example,
when the side group is —CH
2
OH, the amino acid (serine) is
polar, but when the side group is —CH
3
, the amino acid
(alanine) is nonpolar. The 20 common amino acids are
grouped into five chemical classes, based on their side
groups:
1. Nonpolar amino acids, such as leucine, often have R
groups that contain —CH
2
or —CH
3
.
2. Polar uncharged amino acids, such as threonine, have
R groups that contain oxygen (or only —H).
3. Ionizable amino acids, such as glutamic acid, have R
groups that contain acids or bases.
4. Aromatic amino acids, such as phenylalanine, have R
groups that contain an organic (carbon) ring with al-
ternating single and double bonds.
5. Special-function amino acids have unique individual
properties; methionine often is the first amino acid in
a chain of amino acids, proline causes kinks in chains,
and cysteine links chains together.
Each amino acid affects the shape of a protein differently
depending on the chemical nature of its side group. Portions
of a protein chain with numerous nonpolar amino acids, for
example, tend to fold into the interior of the protein by hy-
drophobic exclusion.
Proteins Are Polymers of Amino Acids
In addition to its R group, each amino acid, when ionized,
has a positive amino (NH
3
+
) group at one end and a nega-
tive carboxyl (COO
–
) group at the other end. The amino
and carboxyl groups on a pair of amino acids can undergo a
condensation reaction, losing a molecule of water and
forming a covalent bond. A covalent bond that links two
amino acids is called a peptide bond (figure 3.5). The two
amino acids linked by such a bond are not free to rotate
around the N—C linkage because the peptide bond has a
partial double-bond character, unlike the N—C and C—C
bonds to the central carbon of the amino acid. The stiffness
of the peptide bond is one factor that makes it possible for
chains of amino acids to form coils and other regular
shapes.
A protein is composed of one or more long chains, or
polypeptides, composed of amino acids linked by peptide
bonds. It was not until the pioneering work of Frederick
Sanger in the early 1950s that it became clear that each
kind of protein had a specific amino acid sequence. Sanger
succeeded in determining the amino acid sequence of insu-
lin and in so doing demonstrated clearly that this protein
had a defined sequence, the same for all insulin molecules
in the solution. Although many different amino acids occur
in nature, only 20 commonly occur in proteins. Figure 3.6
illustrates these 20 “common” amino acids and their side
groups.
A protein is a polymer containing a combination of up
to 20 different kinds of amino acids. The amino acids
fall into five chemical classes, each with different
properties. These properties determine the nature of
the resulting protein.
40 Part I The Origin of Living Things
H
—
H — N — C — OH
O
— C
—
H
—
H
H
2
O
—
H — N — C — OH
O
— C
—
H
—
Amino acidAmino acid
H
—
H — N — C —
O
— —
— —
— C
—
H
—
H
—
N — C — OH
O
— —
— —
— C
—
H
—
Polypeptide chain
RR
RR
FIGURE 3.5
The peptide bond. A peptide bond forms when the —NH
2
end
of one amino acid joins to the —COOH end of another. Because
of the partial double-bond nature of peptide bonds, the resulting
peptide chain cannot rotate freely around these bonds.
Chapter 3 The Chemical Building Blocks of Life 41
Proline
(Pro)
CH
2
CH — C — OH
—
—
H
N
— —
—
—
CH
2
CH
2
—
O
Methionine
(Met)
CH
2
H
2
N — C — C — OH
—
—
HO
— —
—
CH
2
—
S
—
CH
3
Cysteine
(Cys)
CH
2
H
2
N — C — C — OH
—
—
HO
— —
—
SH
SPECIAL STRUCTURAL PROPERTY
CH
3
H
2
N — C — C — OH
—
—
HO
— —
CH
H
2
N — C — C — OH
—
—
HO
— —
—
—
CH
3
CH
3
CH
2
H
2
N — C — C — OH
—
—
HO
— —
—
—
CH
3
CH
3
—
CH CH
2
H
2
N — C — C — OH
—
—
HO
— —
—
CH
3
H — C — CH
3
—
CH
2
H
2
N — C — C — OH
—
—
HO
— —
—
CH
2
H
2
N — C — C — OH
—
—
HO
NH
—
C
— ——
—
— —
—
Alanine
(Ala)
Leucine
(Leu)
Isoleucine
(Ile)
Phenylalanine
(Phe)
Tryptophan
(Trp)
CH
2
H
2
N — C — C — OH
—
—
HO
— —
OH
—
H — C — OH
H
2
N — C — C — OH
—
—
HO
— —
CH
3
—
CH
2
H
2
N — C — C — OH
—
—
HO
— —
—
CH
2
—
C
—
NH
2
—
—
O
CH
2
H
2
N — C — C — OH
—
—
HO
OH
—
— ——
—
C
H
2
N — C — C — OH
—
—
HO
—
NH
2
—
— —
—
—
CH
2
O
H
H
2
N — C — C — OH
—
—
HO
— —
Tyrosine
(Tyr)
Glutamine
(Gln)
Asparagine
(Asn)
Threonine
(Thr)
Serine
(Ser)
Glycine
(Gly)
CH
2
H
2
N — C — C — OH
—
—
HO
— —
—
CH
2
—
C
—
—
—
O O
–
CH
2
H
2
N — C — C — OH
—
—
HO
CH
2
—
NH
—
CH
2
—
—
— —
—
—
CNH
2
+
—
NH
2
CH
2
H
2
N — C — C — OH
—
—
HO
— —
—
CH
2
— NH
3
+
—
CH
2
CH
2
—
CH
2
H
2
N — C — C — OH
—
—
HO
—
C — N
— —
— —
HC — N
—
CH
H
H
2
N — C — C — OH
—
—
HO
— —
CH
2
—
C
—
—
—
O O
–
Glutamic
acid (Glu)
Aspartic
acid (Asp)
Histidine
(His)
Lysine
(Lys)
Arginine
(Arg)
Ionizable (charged)
Polar uncharged
Nonpolar
NONAROMATIC AROMATIC
H
+
—
—
Valine
(Val)
FIGURE 3.6
The 20 common amino acids. Each amino acid has the
same chemical backbone, but differs in the side, or R, group
it possesses. Six of the amino acids are nonpolar because
they have —CH
2
or —CH
3
in their R groups. Two of the
six are bulkier because they contain ring structures, which
classifies them also as aromatic. Another six are polar
because they have oxygen or just hydrogen in their R
groups; these amino acids, which are uncharged, differ from
one another in how polar they are. Five other amino acids
are polar and, because they have a terminal acid or base in
their R group, are capable of ionizing to a charged form.
The remaining three have special chemical properties that
allow them to help form links between protein chains or
kinks in proteins.
A Protein’s Function Depends
on the Shape of the Molecule
The shape of a protein is very important because it
determines the protein’s function. If we picture a
polypeptide as a long strand similar to a reed, a pro-
tein might be the basket woven from it.
Overview of Protein Structure
Proteins consist of long amino acid chains folded
into complex shapes. What do we know about the
shape of these proteins? One way to study the
shape of something as small as a protein is to look
at it with very short wavelength energy—with X
rays. X-ray diffraction is a painstaking procedure
that allows the investigator to build up a three-
dimensional picture of the position of each atom.
The first protein to be analyzed in this way was
myoglobin, soon followed by the related protein
hemoglobin. As more and more proteins were add-
ed to the list, a general principle became evident:
in every protein studied, essentially all the internal
amino acids are nonpolar ones, amino acids such as
leucine, valine, and phenylalanine. Water’s tenden-
cy to hydrophobically exclude nonpolar molecules
literally shoves the nonpolar portions of the amino
acid chain into the protein’s interior. This posi-
tions the nonpolar amino acids in close contact
with one another, leaving little empty space inside.
Polar and charged amino acids are restricted to the
surface of the protein except for the few that play
key functional roles.
Levels of Protein Structure
The structure of proteins is traditionally discussed
in terms of four levels of structure, as primary, sec-
ondary, tertiary, and quaternary (figure 3.7). Because
of progress in our knowledge of protein structure,
two additional levels of structure are increasingly
distinguished by molecular biologists: motifs and do-
mains. Because these latter two elements play im-
portant roles in coming chapters, we introduce
them here.
Primary Structure. The specific amino acid se-
quence of a protein is its primary structure. This
sequence is determined by the nucleotide se-
quence of the gene that encodes the protein. Be-
cause the R groups that distinguish the various
amino acids play no role in the peptide backbone
of proteins, a protein can consist of any sequence
of amino acids. Thus, a protein containing 100
amino acids could form any of 20
100
different ami-
no acid sequences (that’s the same as 10
130
, or 1
42 Part I The Origin of Living Things
N
N
N
N
N
H
H
H
H
H
C
C
C
C
O
O
O
C
C
C
C
C
C
O
C
O
N
N
H
H
C
O
C
C
O
C
H
N
N
H
O
O
C
C
C
β-pleated
sheet
α helix
C
OHO
N
HH
Tertiary
structure
Secondary
structure
Primary
structure
Quaternary
structure
(c)
(b)
(a)
(d)
H
H
H
H
H
H
H
H
H
HO
O
O
O
O
O
O
OO
O
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
CN
N
N
N
N
N
N
N
N
N
FIGURE 3.7
Levels of protein structure. The amino acid sequence of a protein is called
its (a) primary structure. Hydrogen bonds form between nearby amino acids,
producing (b) fold-backs called beta-pleated sheets and coils called alpha
helices. These fold-backs and coils constitute the protein’s secondary
structure. A globular protein folds up on itself further to assume a three-
dimensional (c) tertiary structure. Many proteins aggregate with other
polypeptide chains in clusters; this clustering is called the (d) quaternary
structure of the protein.
followed by 130 zeros—more than the number of atoms
known in the universe). This is an important property of
proteins because it permits such great diversity.
Secondary Structure. The amino acid side groups are
not the only portions of proteins that form hydrogen
bonds. The —COOH and —NH
2
groups of the main
chain also form quite good hydrogen bonds—so good that
their interactions with water might be expected to offset
the tendency of nonpolar sidegroups to be forced into the
protein interior. Inspection of the protein structures deter-
mined by X-ray diffraction reveals why they don’t—the
polar groups of the main chain form hydrogen bonds with
each other! Two patterns of H bonding occur. In one, hy-
drogen bonds form along a single chain, linking one amino
acid to another farther down the chain. This tends to pull
the chain into a coil called an alpha (α) helix. In the other
pattern, hydrogen bonds occur across two chains, linking
the amino acids in one chain to those in the other. Often,
many parallel chains are linked, forming a pleated, sheet-
like structure called a β-pleated sheet. The folding of the
amino acid chain by hydrogen bonding into these charac-
teristic coils and pleats is called a protein’s secondary
structure.
Motifs. The elements of secondary structure can combine
in proteins in characteristic ways called motifs, or some-
times “supersecondary structure.” One very common motif
is the β α β motif, which creates a fold or crease; the so-
called “Rossmann fold” at the core of nucleotide binding
sites in a wide variety of proteins is a β α β α β motif. A sec-
ond motif that occurs in many proteins is the β barrel, a β
sheet folded around to form a tube. A third type of motif,
the α turn α motif, is important because many proteins use
it to bind the DNA double helix.
Tertiary Structure. The final folded shape of a globular
protein, which positions the various motifs and folds non-
polar side groups into the interior, is called a protein’s ter-
tiary structure. A protein is driven into its tertiary structure
by hydrophobic interactions with water. The final folding
of a protein is determined by its primary structure—by the
chemical nature of its side groups. Many proteins can be
fully unfolded (“denatured”) and will spontaneously refold
back into their characteristic shape.
The stability of a protein, once it has folded into its 3-D
shape, is strongly influenced by how well its interior fits
together. When two nonpolar chains in the interior are in
very close proximity, they experience a form of molecular
attraction called van der Waal’s forces. Individually quite
weak, these forces can add up to a strong attraction when
many of them come into play, like the combined strength
of hundreds of hooks and loops on a strip of Velcro. They
are effective forces only over short distances, however;
there are no “holes” or cavities in the interior of proteins.
That is why there are so many different nonpolar amino
acids (alanine, valine, leucine, isoleucine). Each has a dif-
ferent-sized R group, allowing very precise fitting of non-
polar chains within the protein interior. Now you can un-
derstand why a mutation that converts one nonpolar ami-
no acid within the protein interior (alanine) into another
(leucine) very often disrupts the protein’s stability; leucine
is a lot bigger than alanine and disrupts the precise way the
chains fit together within the protein interior. A change in
even a single amino acid can have profound effects on pro-
tein shape and can result in loss or altered function of the
protein.
Domains. Many proteins in your body are encoded within
your genes in functional sections called exons (exons will be
discussed in detail in chapter 15). Each exon-encoded sec-
tion of a protein, typically 100 to 200 amino acids long,
folds into a structurally independent functional unit called a
domain. As the polypeptide chain folds, the domains fold
into their proper shape, each more-or-less independent of
the others. This can be demonstrated experimentally by ar-
tificially producing the fragment of polypeptide that forms
the domain in the intact protein, and showing that the frag-
ment folds to form the same structure as it does in the intact
protein.
A single polypeptide chain connects the domains of a
protein, like a rope tied into several adjacent knots. Often
the domains of a protein have quite separate functions—one
domain of an enzyme might bind a cofactor, for example,
and another the enzyme’s substrate.
Quaternary Structure. When two or more polypeptide
chains associate to form a functional protein, the individ-
ual chains are referred to as subunits of the protein. The
subunits need not be the same. Hemoglobin, for example,
is a protein composed of two α-chain subunits and two β-
chain subunits. A protein’s subunit arrangement is called
its quaternary structure. In proteins composed of sub-
units, the interfaces where the subunits contact one an-
other are often nonpolar, and play a key role in transmit-
ting information between the subunits about individual
subunit activities.
A change in the identity of one of these amino acids can
have profound effects. Sickle cell hemoglobin is a mutation
that alters the identity of a single amino acid at the corner
of the β subunit from polar glutamate to nonpolar valine.
Putting a nonpolar amino acid on the surface creates a
“sticky patch” that causes one hemoglobin molecule to
stick to another, forming long nonfunctional chains and
leading to the cell sickling characteristic of this hereditary
disorder.
Protein structure can be viewed at six levels: 1. the
amino acid sequence, or primary structure; 2. coils and
sheets, called secondary structure; 3. folds or creases,
called motifs; 4. the three-dimensional shape, called
tertiary structure; 5. functional units, called domains;
and 6. individual polypeptide subunits associated in a
quaternary structure.
Chapter 3 The Chemical Building Blocks of Life 43
How Proteins Fold Into
Their Functional Shape
How does a protein fold into a specific
shape? Nonpolar amino acids play a key
role. Until recently, investigators
thought that newly made proteins fold
spontaneously as hydrophobic interac-
tions with water shove nonpolar amino
acids into the protein interior. We now
know this is too simple a view. Protein
chains can fold in so many different
ways that trial and error would simply
take too long. In addition, as the open
chain folds its way toward its final form,
nonpolar “sticky” interior portions are
exposed during intermediate stages. If
these intermediate forms are placed in a
test tube in the same protein environ-
ment that occurs in a cell, they stick to
other unwanted protein partners, form-
ing a gluey mess.
Chaperonins
How do cells avoid this? A vital clue
came in studies of unusual mutations
that prevented viruses from replicating
in E. coli bacterial cells—it turned out
the virus proteins could not fold prop-
erly! Further study revealed that nor-
mal cells contain special proteins called
chaperonins that help new proteins
fold correctly (figure 3.8). When the E.
coli gene encoding its chaperone pro-
tein is disabled by mutation, the bacte-
ria die, clogged with lumps of incor-
rectly folded proteins. Fully 30% of the
bacteria’s proteins fail to fold to the
right shape.
Molecular biologists have now identified more than 17
kinds of proteins that act as molecular chaperones. Many
are heat shock proteins, produced in greatly elevated
amounts if a cell is exposed to elevated temperature; high
temperatures cause proteins to unfold, and heat shock
chaperonins help the cell’s proteins refold.
There is considerable controversy about how chaper-
onins work. It was first thought that they provided a pro-
tected environment within which folding could take place
unhindered by other proteins, but it now seems more like-
ly that chaperonins rescue proteins that are caught in a
wrongly folded state, giving them another chance to fold
correctly. When investigators “fed” a deliberately mis-
folded protein called malate dehydrogenase to chaper-
onins, the protein was rescued, refolding to its active
shape.
Protein Folding and Disease
There are tantalizing suggestions that chaperone protein
deficiencies may play a role in certain diseases by failing to
facilitate the intricate folding of key proteins. Cystic fibrosis
is a hereditary disorder in which a mutation disables a pro-
tein that plays a vital part in moving ions across cell mem-
branes. In at least some cases, the vital membrane protein
appears to have the correct amino acid sequence, but fails to
fold to its final form. It has also been speculated that chaper-
one deficiency may be a cause of the protein clumping in
brain cells that produces the amyloid plaques characteristic
of Alzheimer’s disease.
Proteins called chaperones aid newly produced proteins
to fold properly.
44 Part I The Origin of Living Things
Correctly
folded protein
Unfolded
protein
Chaperone proteins present
Chaperone proteins absent
Unfolded
Incorrectly folded
Enzyme
degradation
of protein
FIGURE 3.8
A current model of protein folding. A newly synthesized protein rapidly folds into
characteristic motifs composed of H9251 helices and H9252 sheets, but these elements of structure are
only roughly aligned in an open conformation. Subsequent folding occurs more slowly, by
trial and error. This process is aided by chaperone proteins, which appear to recognize
improperly folded proteins and unfold them, giving them another chance to fold properly.
Eventually, if proper folding is not achieved, the misfolded protein is degraded by
proteolytic enzymes.
How Proteins Unfold
If a protein’s environment is altered, the protein
may change its shape or even unfold. This pro-
cess is called denaturation. Proteins can be de-
natured when the pH, temperature, or ionic
concentration of the surrounding solution is
changed. When proteins are denatured, they are
usually rendered biologically inactive. This is
particularly significant in the case of enzymes.
Because practically every chemical reaction in a
living organism is catalyzed by a specific en-
zyme, it is vital that a cell’s enzymes remain
functional. That is the rationale behind tradi-
tional methods of salt-curing and pickling: prior
to the ready availability of refrigerators and
freezers, the only practical way to keep microor-
ganisms from growing in food was to keep the
food in a solution containing high salt or vine-
gar concentrations, which denatured the en-
zymes of microorganisms and kept them from
growing on the food.
Most enzymes function within a very narrow
range of physical parameters. Blood-borne en-
zymes that course through a human body at a
pH of about 7.4 would rapidly become dena-
tured in the highly acidic environment of the
stomach. On the other hand, the protein-
degrading enzymes that function at a pH of 2 or
less in the stomach would be denatured in the
basic pH of the blood. Similarly, organisms that live near
oceanic hydrothermal vents have enzymes that work well at
the temperature of this extreme environment (over 100°C).
They cannot survive in cooler waters, because their en-
zymes would denature at lower temperatures. Any given
organism usually has a “tolerance range” of pH, tempera-
ture, and salt concentration. Within that range, its enzymes
maintain the proper shape to carry out their biological
functions.
When a protein’s normal environment is reestablished
after denaturation, a small protein may spontaneously re-
fold into its natural shape, driven by the interactions be-
tween its nonpolar amino acids and water (figure 3.9).
Larger proteins can rarely refold spontaneously because
of the complex nature of their final shape. It is important
to distinguish denaturation from dissociation. The four
subunits of hemoglobin (figure 3.10) may dissociate into
four individual molecules (two α-globin and two β-
globin) without denaturation of the folded globin pro-
teins, and will readily reassume their four-subunit quater-
nary structure.
Every globular protein has a narrow range of conditions
in which it folds properly; outside that range, proteins
tend to unfold.
Chapter 3 The Chemical Building Blocks of Life 45
Reducin
agent
—N—C—
H H O
CH
2
— —
—
— —
Disulfide
—C—C—
—
—
——
O H H
CH
2
—
—
S
—
S
—
Cooling or
removal of
urea
Heating or
addition of
urea
Unfolded
ribonuclease
Reduced
ribonuclease
Native
ribonuclease
FIGURE 3.9
Primary structure determines tertiary structure. When the protein
ribonuclease is treated with reducing agents to break the covalent disulfide bonds
that cross-link its chains, and then placed in urea or heated, the protein denatures
(unfolds) and loses its enzymatic activity. Upon cooling or removal of urea, it
refolds and regains its enzymatic activity. This demonstrates that no information
but the amino acid sequence of the protein is required for proper folding: the
primary structure of the protein determines its tertiary structure.
Beta chains
Heme group
Alpha chains
FIGURE 3.10
The four subunits of hemoglobin. The hemoglobin molecule is
made of four globin protein subunits, informally referred to as
polypeptide chains. The two lower α chains, identical α-globin
proteins, are shaded pink; the two upper β chains, identical β-
globin proteins, are shaded blue.
Information Molecules
The biochemical activity of a cell depends on production of
a large number of proteins, each with a specific sequence.
The ability to produce the correct proteins is passed be-
tween generations of organisms, even though the protein
molecules themselves are not.
Nucleic acids are the information storage devices of
cells, just as disks or tapes store the information that com-
puters use, blueprints store the information that builders
use, and road maps store the information that tourists use.
There are two varieties of nucleic acids: deoxyribonucleic
acid (DNA; figure 3.11) and ribonucleic acid (RNA). The
way in which DNA encodes the information used to as-
semble proteins is similar to the way in which the letters
on a page encode information (see chapter 14). Unique
among macromolecules, nucleic acids are able to serve as
templates to produce precise copies of themselves, so that
the information that specifies what an organism is can be
copied and passed down to its descendants. For this rea-
son, DNA is often referred to as the hereditary material.
Cells use the alternative form of nucleic acid, RNA, to
read the cell’s DNA-encoded information and direct the
synthesis of proteins. RNA is similar to DNA in structure
and is made as a transcripted copy of portions of the
DNA. This transcript passes out into the rest of the cell,
where it serves as a blueprint specifying a protein’s amino
acid sequence. This process will be described in detail in
chapter 15.
“Seeing” DNA
DNA molecules cannot be seen with an optical micro-
scope, which is incapable of resolving anything smaller
than 1000 atoms across. An electron microscope can
image structures as small as a few dozen atoms across, but
still cannot resolve the individual atoms of a DNA strand.
This limitation was finally overcome in the last decade
with the introduction of the scanning-tunneling micro-
scope (figure 3.12).
How do these microscopes work? Imagine you are in a
dark room with a chair. To determine the shape of the
chair, you could shine a flashlight on it, so that the light
bounces off the chair and forms an image on your eye.
That’s what optical and electron microscopes do; in the lat-
ter, the “flashlight” emits a beam of electrons instead of
light. You could, however, also reach out and feel the
chair’s surface with your hand. In effect, you would be put-
ting a probe (your hand) near the chair and measuring how
far away the surface is. In a scanning-tunneling microscope,
computers advance a probe over the surface of a molecule
in steps smaller than the diameter of an atom.
46 Part I The Origin of Living Things
3.3 Nucleic acids store and transfer genetic information.
FIGURE 3.11
The first photograph of a DNA molecule. This micrograph,
with sketch below, shows a section of DNA magnified a million
times! The molecule is so slender that it would take 50,000 of
them to equal the diameter of a human hair.
FIGURE 3.12
A scanning tunneling micrograph of DNA (false color;
2,000,000×). The micrograph shows approximately three turns of
the DNA double helix (see figure 3.15).
The Structure of Nucleic Acids
Nucleic acids are long polymers of repeating subunits called
nucleotides. Each nucleotide consists of three components:
a five-carbon sugar (ribose in RNA and deoxyribose in
DNA); a phosphate (—PO
4
) group; and an organic nitrogen-
containing base (figure 3.13). When a nucleic acid polymer
forms, the phosphate group of one nucleotide binds to the
hydroxyl group of another, releasing water and forming a
phosphodiester bond. A nucleic acid, then, is simply a
chain of five-carbon sugars linked together by phosphodi-
ester bonds with an organic base protruding from each
sugar (figure 3.14).
Two types of organic bases occur in nucleotides. The
first type, purines, are large, double-ring molecules found in
both DNA and RNA; they are adenine (A) and guanine
(G). The second type, pyrimidines, are smaller, single-ring
molecules; they include cytosine (C, in both DNA and
RNA), thymine (T, in DNA only), and uracil (U, in RNA
only).
Chapter 3 The Chemical Building Blocks of Life 47
Phosphate group
Sugar
Nitrogenous base
N
O
4
5
1
32
P CH
2
O
–
O
O
–
OH R
OH in RNA
H in DNA
O
FIGURE 3.13
Structure of a nucleotide. The nucleotide subunits of DNA and
RNA are made up of three elements: a five-carbon sugar, an
organic nitrogenous base, and a phosphate group.
5H11032
3H11032
P
P
P
P
OH
5-carbon sugar
Nitrogenous base
Phosphate group
Phosphodiester
bond
HC C
NC
H
N
C
NH
2
Adenine
H
2
NC C
N
N
N
C
H
N
C
CH
O
O
O
O
O
Guanine
H
N
N
CH
OC C
NC
H
N
C
NH
2
Cytosine
(both DNA
and RNA)
Thymine
(DNA only)
Uracil
(RNA only)
O CC
NC
H
N
C
O
H
H
H
CH
3
H
O CC
NC
H
N
C
O
H H
H
P
U
R
I
N
E
S
P
Y
R
I
M
I
D
I
N
E
S
(a)
(b)
FIGURE 3.14
The structure of a nucleic acid and the organic nitrogen-containing bases. (a) In a nucleic acid, nucleotides are linked to one another
via phosphodiester bonds, with organic bases protruding from the chain. (b) The organic nitrogenous bases can be either purines or
pyrimidines. In DNA, thymine replaces the uracil found in RNA.
DNA
Organisms encode the information
specifying the amino acid sequences
of their proteins as sequences of nu-
cleotides in the DNA. This method
of encoding information is very sim-
ilar to that by which the sequences
of letters encode information in a
sentence. While a sentence written
in English consists of a combination
of the 26 different letters of the al-
phabet in a specific order, the code
of a DNA molecule consists of dif-
ferent combinations of the four
types of nucleotides in specific se-
quences such as CGCTTACG. The
information encoded in DNA is used
in the everyday metabolism of the
organism and is passed on to the or-
ganism’s descendants.
DNA molecules in organisms ex-
ist not as single chains folded into
complex shapes, like proteins, but
rather as double chains. Two DNA
polymers wind around each other
like the outside and inside rails of a
circular staircase. Such a winding
shape is called a helix, and a helix
composed of two chains winding
about one another, as in DNA, is
called a double helix. Each step of
DNA’s helical staircase is a base-
pair, consisting of a base in one
chain attracted by hydrogen bonds
to a base opposite it on the other
chain. These hydrogen bonds hold
the two chains together as a duplex
(figure 3.15). The base-pairing rules
are rigid: adenine can pair only with
thymine (in DNA) or with uracil (in
RNA), and cytosine can pair only
with guanine. The bases that partici-
pate in base-pairing are said to be
complementary to each other. Addi-
tional details of the structure of
DNA and how it interacts with RNA
in the production of proteins are
presented in chapters 14 and 15.
RNA
RNA is similar to DNA, but with two major chemical
differences. First, RNA molecules contain ribose sugars
in which the number 2 carbon is bonded to a hydroxyl
group. In DNA, this hydroxyl group is replaced by a hy-
drogen atom. Second, RNA molecules utilize uracil in
48 Part I The Origin of Living Things
O
OH
3H11032 end
5H11032 end
O
O
O
G
C
P
O
O
O
O
O
O
O
P
P
P
P
P
P
P
P
P
C
C
G
G
A
A
T
T
Sugar-phosphate "backbone"
Hydrogen bonds between
nitrogenous bases
Phosphodiester
bond
FIGURE 3.15
The structure of DNA. Hydrogen bond formation (dashed lines) between the organic
bases, called base-pairing, causes the two chains of a DNA duplex to bind to each other
and form a double helix.
place of thymine. Uracil has the same structure as thy-
mine, except that one of its carbons lacks a methyl (—
CH
3
) group.
Transcribing the DNA message into a chemically differ-
ent molecule such as RNA allows the cell to tell which is
the original information storage molecule and which is the
transcript. DNA molecules are always double-stranded (ex-
cept for a few single-stranded DNA viruses that will be dis-
cussed in chapter 33), while the RNA molecules tran-
scribed from DNA are typically single-stranded (figure
3.16). Although there is no chemical reason why RNA can-
not form double helices as DNA does, cells do not possess
the enzymes necessary to assemble double strands of RNA,
as they do for DNA. Using two different molecules, one
single-stranded and the other double-stranded, separates
the role of DNA in storing hereditary information from the
role of RNA in using this information to specify protein
structure.
Which Came First, DNA or RNA?
The information necessary for the synthesis of proteins is
stored in the cell’s double-stranded DNA base sequences.
The cell uses this information by first making an RNA
transcript of it: RNA nucleotides pair with complementary
DNA nucleotides. By storing the information in DNA while
using a complementary RNA sequence to actually direct
protein synthesis, the cell does not expose the information-
encoding DNA chain to the dangers of single-strand cleav-
age every time the information is used. Therefore, DNA is
thought to have evolved from RNA as a means of preserv-
ing the genetic information, protecting it from the ongo-
ing wear and tear associated with cellular activity. This ge-
netic system has come down to us from the very
beginnings of life.
The cell uses the single-stranded, short-lived RNA tran-
script to direct the synthesis of a protein with a specific se-
quence of amino acids. Thus, the information flows from
DNA to RNA to protein, a process that has been termed the
“central dogma” of molecular biology.
ATP
In addition to serving as subunits of DNA and RNA, nu-
cleotide bases play other critical roles in the life of a cell.
For example, adenine is a key component of the molecule
adenosine triphosphate (ATP; figure 3.17), the energy cur-
rency of the cell. It also occurs in the molecules nicotina-
mide adenine dinucleotide (NAD
+
) and flavin adenine dinu-
cleotide (FAD
+
), which carry electrons whose energy is used
to make ATP.
A nucleic acid is a long chain of five-carbon sugars with
an organic base protruding from each sugar. DNA is a
double-stranded helix that stores hereditary
information as a specific sequence of nucleotide bases.
RNA is a single-stranded molecule that transcribes this
information to direct protein synthesis.
Chapter 3 The Chemical Building Blocks of Life 49
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
DNA
Deoxyribose-phosphate
backbone
Bases
Hydrogen bonding
occurs between base-pairs
RNA
Ribose-phosphate
backbone
Bases
G
C
G
G
G
C
C
T
A
A
A
A
A
A
T
TT
G
C
U
U
FIGURE 3.16
DNA versus RNA. DNA forms a double helix, uses deoxyribose as
the sugar in its sugar-phosphate backbone, and utilizes thymine
among its nitrogenous bases. RNA, on the other hand, is usually
single-stranded, uses ribose as the sugar in its sugar-phosphate
backbone, and utilizes uracil in place of thymine.
Triphosphate group
5-carbon sugar
Nitrogenous base
(adenine)
O
4
5
1
32
P CH
2
O
O
O
–
P
O
O
O
–
P
O
–
O
O
–
OH OH
O
NH
2
N
N
N
N
FIGURE 3.17
ATP. Adenosine triphosphate (ATP) contains adenine, a five-
carbon sugar, and three phosphate groups. This molecule serves
to transfer energy rather than store genetic information.
Lipids are a loosely defined group of molecules with one
main characteristic: they are insoluble in water. The most
familiar lipids are fats and oils. Lipids have a very high pro-
portion of nonpolar carbon-hydrogen (C—H) bonds, and so
long-chain lipids cannot fold up like a protein to sequester
their nonpolar portions away from the surrounding aqueous
environment. Instead, when placed in water many lipid mol-
ecules will spontaneously cluster together and expose what
polar groups they have to the surrounding water while se-
questering the nonpolar parts of the molecules together
within the cluster. This spontaneous assembly of lipids is of
paramount importance to cells, as it underlies the structure
of cellular membranes.
Phospholipids Form Membranes
Phospholipids are among the most important molecules of
the cell, as they form the core of all biological membranes.
An individual phospholipid is a composite molecule, made
up of three kinds of subunits:
1. Glycerol, a three-carbon alcohol, with each carbon
bearing a hydroxyl group. Glycerol forms the back-
bone of the phospholipid molecule.
2. Fatty acids, long chains of C—H bonds (hydrocarbon
chains) ending in a carboxyl (—COOH) group. Two
fatty acids are attached to the glycerol backbone in a
phospholipid molecule.
3. Phosphate group, attached to one end of the glycerol.
The charged phosphate group usually has a charged
organic molecule linked to it, such as choline, etha-
nolamine, or the amino acid serine.
The phospholipid molecule can be thought of as having a
polar “head” at one end (the phosphate group) and two
long, very nonpolar “tails” at the other. In water, the non-
polar tails of nearby phospholipids aggregate away from the
water, forming two layers of tails pointed toward each oth-
er—a lipid bilayer (figure 3.18). Lipid bilayers are the basic
framework of biological membranes, discussed in detail in
chapter 6.
H
|
H—C—Fatty acid
|
H—C—Fatty acid
|
H—C—Phosphate group
|
H
Because the C—H bonds in lipids are very nonpolar,
they are not water-soluble, and aggregate together in
water. This kind of aggregation by phospholipids forms
biological membranes.
50 Part I The Origin of Living Things
3.4 Lipids make membranes and store energy.
Hydrophobic
"tails"
Hydrophilic
"heads"
Hydrophobic
region
Hydrophilic
region
Hydrophilic
region
Oil
Water
Water
Water
(a)
(b)
FIGURE 3.18
Phospholipids. (a) At an oil-water interface, phospholipid molecules will orient so that their polar (hydrophilic) heads are in the polar
medium, water, and their nonpolar (hydrophobic) tails are in the nonpolar medium, oil. (b) When surrounded by water, phospholipid
molecules arrange themselves into two layers with their heads extending outward and their tails inward.
Fats and Other Kinds of Lipids
Fats are another kind of lipid, but unlike phospholipids, fat
molecules do not have a polar end. Fats consist of a glycerol
molecule to which is attached three fatty acids, one to each
carbon of the glycerol backbone. Because it contains three
fatty acids, a fat molecule is called a triglyceride, or, more
properly, a triacylglycerol (figure 3.19). The three fatty
acids of a triglyceride need not be identical, and often they
differ markedly from one another. Organisms store the en-
ergy of certain molecules for long periods in the many
C—H bonds of fats.
Because triglyceride molecules lack a polar end, they are
not soluble in water. Placed in water, they spontaneously
clump together, forming fat globules that are very large rel-
ative to the size of the individual molecules. Because fats are
insoluble, they can be deposited at specific locations within
an organism.
Storage fats are one kind of lipid. Oils such as olive oil,
corn oil, and coconut oil are also lipids, as are waxes such as
beeswax and earwax (see table 3.1). The hydrocarbon chains
of fatty acids vary in length; the most common are even-
numbered chains of 14 to 20 carbons. If all of the internal
carbon atoms in the fatty acid chains are bonded to at least
two hydrogen atoms, the fatty acid is said to be saturated,
because it contains the maximum possible number of hydro-
gen atoms (figure 3.20). If a fatty acid has double bonds be-
tween one or more pairs of successive carbon atoms, the
fatty acid is said to be unsaturated. If a given fatty acid has
more than one double bond, it is said to be polyunsaturat-
ed. Fats made from polyunsaturated fatty acids have low
melting points because their fatty acid chains bend at the
double bonds, preventing the fat molecules from aligning
closely with one another. Consequently, a polyunsaturated
fat such as corn oil is usually liquid at room temperature and
is called an oil. In contrast, most saturated fats such as those
in butter are solid at room temperature.
Organisms contain many other kinds of lipids besides
fats (see figure 3.19). Terpenes are long-chain lipids that are
components of many biologically important pigments, such
as chlorophyll and the visual pigment retinal. Rubber is
also a terpene. Steroids, another type of lipid found in mem-
branes, are composed of four carbon rings. Most animal
cell membranes contain the steroid cholesterol. Other ste-
roids, such as testosterone and estrogen, function in multi-
cellular organisms as hormones. Prostaglandins are a group
of about 20 lipids that are modified fatty acids, with two
nonpolar “tails” attached to a five-carbon ring. Prostaglan-
dins act as local chemical messengers in many vertebrate
tissues.
Cells contain many kinds of molecules in addition to
membrane phospholipids that are not soluble in water.
Chapter 3 The Chemical Building Blocks of Life 51
(a) Phospholipid (phosphatidyl choline)
HO
(c) Terpene (citronellol)
(d) Steroid (cholesterol)
OP
O
–
O
O
C
CH
3
CH
3
OH
CH
3
CH
3
CH
CH
2
CH
CH
2
CH
2
(CH
2
)
14
CH
3
CH
3
CO
(CH
2
)
7
(CH
2
)
7
CH CHCO
CH
2
CH
3
CH
3
CH
3
N
+
H
3
CCH
2
CH
2
CH
2
CH
CH
3
CH
3
CH
2
CH
2
CH
2
CH
2
CH
CH
2
CH
2
O
O
H
H
CHO
CHO
CHO C
O
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
H
C
O
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
H
C
O
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
H
(b) Triacylglycerol molecule
FIGURE 3.19
Lipids. These structures represent four major classes of biologically important lipids: (a) phospholipids, (b) triacylglycerols
(triglycerides), (c) terpenes, and (d) steroids.
Fats as Food
Most fats contain over 40 carbon atoms. The ratio of energy-
storing C—H bonds to carbon atoms in fats is more than
twice that of carbohydrates (see next section), making fats
much more efficient molecules for storing chemical energy.
On the average, fats yield about 9 kilocalories (kcal) of
chemical energy per gram, as compared with somewhat less
than 4 kcal per gram for carbohydrates.
All fats produced by animals are saturated (except some
fish oils), while most plant fats are unsaturated. The excep-
tions are the tropical oils (palm oil and coconut oil), which
are saturated despite their fluidity at room temperature. It
is possible to convert an oil into a solid fat by adding hy-
drogen. Peanut butter sold in stores is usually artificially
hydrogenated to make the peanut fats solidify, preventing
them from separating out as oils while the jar sits on the
store shelf. However, artificially hydrogenating unsaturat-
ed fats seems to eliminate the health advantage they have
over saturated fats, as it makes both equally rich in C—H
bonds. Therefore, it now appears that margarine made
from hydrogenated corn oil is no better for your health
than butter.
When an organism consumes excess carbohydrate, it is
converted into starch, glycogen, or fats and reserved for fu-
ture use. The reason that many humans gain weight as they
grow older is that the amount of energy they need decreas-
es with age, while their intake of food does not. Thus, an
increasing proportion of the carbohydrate they ingest is
available to be converted into fat.
A diet rich in fats is one of several factors that are
thought to contribute to heart disease, particularly to ath-
erosclerosis, a condition in which deposits of fatty tissue
called plaque adhere to the lining of blood vessels, blocking
the flow of blood. Fragments of plaque, breaking off from a
deposit, are a major cause of strokes.
Fats are efficient energy-storage molecules because of
their high concentration of C—H bonds.
52 Part I The Origin of Living Things
O
C
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
HHO
H
C
H
C
H
H
C
H
H
O
C
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
H
C
H
C
H
H
C
H
C
H
C
H
H
C
H
C
H
C
H
HO
No double bonds between carbon atoms; fatty acid
chains fit close together
Double bonds present between carbon atoms; fatty acid
chains do not fit close together
(a) Saturated fat
(b) Unsaturated fat
FIGURE 3.20
Saturated and unsaturated fats. (a) Palmitic acid, with no double bonds and, thus, a maximum number of hydrogen atoms bonded to the
carbon chain, is a saturated fatty acid. Many animal triacylglycerols (fats) are saturated. Because their fatty acid chains can fit closely
together, these triacylglycerols form immobile arrays called hard fat. (b) Linoleic acid, with three double bonds and, thus, fewer than the
maximum number of hydrogen atoms bonded to the carbon chain, is an unsaturated fatty acid. Plant fats are typically unsaturated. The
many kinks the double bonds introduce into the fatty acid chains prevent the triacylglycerols from closely aligning and produce oils, which
are liquid at room temperature.
Simple Carbohydrates
Carbohydrates function as energy-storage molecules as well
as structural elements. Some are small, simple molecules,
while others form long polymers.
Sugars Are Simple Carbohydrates
The carbohydrates are a loosely defined group of mole-
cules that contain carbon, hydrogen, and oxygen in the
molar ratio 1:2:1. Their empirical formula (which lists the
atoms in the molecule with subscripts to indicate how many
there are of each) is (CH
2
O)
n
, where n is the number of car-
bon atoms. Because they contain many carbon-hydrogen
(C—H) bonds, which release energy when they are broken,
carbohydrates are well suited for energy storage.
Monosaccharides. The simplest of the carbohydrates
are the simple sugars, or monosaccharides (Greek mono,
“single” + Latin saccharum, “sugar”). Simple sugars may
contain as few as three carbon atoms, but those that play
the central role in energy storage have six (figure 3.21).
The empirical formula of six-carbon sugars is:
C
6
H
12
O
6
, or (CH
2
O)
6
Six-carbon sugars can exist in a straight-chain form, but in
an aqueous environment they almost always form rings.
The most important of these for energy storage is glucose
(figure 3.22), a six-carbon sugar which has seven energy-
storing C—H bonds.
Chapter 3 The Chemical Building Blocks of Life 53
3.5 Carbohydrates store energy and provide building materials.
4
5
1
3 2
H H
H
HH
H
OH OH
OH
O
4
4
4
4
5
5
5
5
6
6
6
1
1
1
1
3
3
2
23
3
2
2
OH H
OH
O
CH
2
OH
CH
2
OH
CH
2
OH
CH
2
OH
OH
HO
O
OH
OH
HO
CH
2
OH
H
H
H
HO
H H
H
H
OH
H
OH
OH
OH
O
H
H
HO
CH
2
OH
H
H
H
OH
O
GalactoseFructoseGlucose
RiboseGlyceraldehyde
3-carbon
sugar
5-carbon sugars
6-carbon sugars
Deoxyribose
HH
1
3
2
H
H
H
H
OH
OH
O
C
C
C
FIGURE 3.21
Monosaccharides. Monosaccharides, or simple sugars, can
contain as few as three carbon atoms and are often used as
building blocks to form larger molecules. The five-carbon sugars
ribose and deoxyribose are components of nucleic acids (see figure
3.15). The six-carbon sugar glucose is a component of large
energy-storage molecules.
CH
2
OH
CH
2
OH
HO
H
H
CCC
324
O
H
OHH
H
H
H
OH
H
H
H
H
H
OH
O
H
O
H
O
H
O
H
O
C
H
1
O
HC
56
H
C
H
H
H
HO
O
H
HO
OH
C
H
CCC
23
4
6
5
3
2 1
4
5
6
1
C
H
HO
H
CCO
C
H
HH
OH
H
O
H
OH
C
O
H
H
23
4
5
6
1
H
O
H
O O
H
O
C
H O
H
O
H
HC
H
341256
6
5
4
1
2
3
H
CH
H
O
FIGURE 3.22
Structure of the glucose molecule. Glucose is a linear six-
carbon molecule that forms a ring shape in solution. The structure
of the ring can be represented in many ways; the ones shown here
are the most common, with the carbons conventionally numbered
(in green) so that the forms can be compared easily. The bold,
darker lines represent portions of the molecule that are projecting
out of the page toward you—remember, these are three-
dimensional molecules!
Disaccharides. Many familiar sugars like sucrose are
“double sugars,” two monosaccharides joined by a covalent
bond (figure 3.23). Called disaccharides, they often play a
role in the transport of sugars, as we will discuss shortly.
Polysaccharides. Polysaccharides are macromolecules
made up of monosaccharide subunits. Starch is a polysac-
charide used by plants to store energy. It consists entirely of
glucose molecules, linked one after another in long chains.
Cellulose is a polysaccharide that serves as a structural
building material in plants. It too consists entirely of glucose
molecules linked together into chains, and special enzymes
are required to break the links.
Sugar Isomers
Glucose is not the only sugar with the formula C
6
H
12
O
6
.
Other common six-carbon sugars such as fructose and ga-
lactose also have this same empirical formula (figure 3.24).
These sugars are isomers, or alternative forms, of glucose.
Even though isomers have the same empirical formula,
their atoms are arranged in different ways; that is, their
three-dimensional structures are different. These structural
differences often account for substantial functional differ-
ences between the isomers. Glucose and fructose, for exam-
ple, are structural isomers. In fructose, the double-bonded
oxygen is attached to an internal carbon rather than to a
terminal one. Your taste buds can tell the difference, as
fructose tastes much sweeter than glucose, despite the fact
that both sugars have the same chemical composition. This
structural difference also has an important chemical conse-
quence: the two sugars form different polymers.
Unlike fructose, galactose has the same bond structure as
glucose; the only difference between galactose and glucose
is the orientation of one hydroxyl group. Because the hy-
droxyl group positions are mirror images of each other,
galactose and glucose are called stereoisomers. Again, this
seemingly slight difference has important consequences, as
this hydroxyl group is often involved in creating polymers
with distinct functions, such as starch (energy storage) and
cellulose (structural support).
Sugars are among the most important energy-storage
molecules in organisms, containing many energy-storing
C—H bonds. The structural differences among sugar
isomers can confer substantial functional differences
upon the molecules.
54 Part I The Origin of Living Things
Sucrose
CH
2
OH
CH
2
OH
OH
HO
O
CH
2
OH
HO
OH
OH
O
O
Lactose
OH
OH
HO
CH
2
OH
H
H H
H
H
H
H
H
H
H
H
H
H
OH
OH
OH
O
H
H
O
CH
2
OH
H
H
H
OH
O
Galactose
Glucose
Glucose Fructose
FIGURE 3.23
Disaccharides. Sugars like sucrose and lactose are disaccharides,
composed of two monosaccharides linked by a covalent bond.
C
—
— —
OC
—
— —
O
—
——
H — C — OH
OH
—
H — C — OH
—
H
HO — C — H
C
—
—
O
—
H — C — OH
—
H
Fructose
H — C — OH
—
——
H — C — OH
—
H — C — OH
—
H
HO — C — H
—
H — C — OH
—
H
Glucose
HO — C — H
—
——
H — C — OH
—
H — C — OH
—
H
HO — C — H
—
H — C — OH
—
H
Galactose
Structural
isomer Stereoisomer
H — C —
FIGURE 3.24
Isomers and stereoisomers. Glucose,
fructose, and galactose are isomers with the
empirical formula C
6
H
12
O
6
. A structural
isomer of glucose, such as fructose, has
identical chemical groups bonded to different
carbon atoms, while a stereoisomer of glucose,
such as galactose, has identical chemical groups
bonded to the same carbon atoms but in
different orientations.
Linking Sugars Together
Transport Disaccharides
Most organisms transport sugars within their bodies. In
humans, the glucose that circulates in the blood does so as
a simple monosaccharide. In plants and many other or-
ganisms, however, glucose is converted into a transport
form before it is moved from place to place within the or-
ganism. In such a form it is less readily metabolized (used
for energy) during transport. Transport forms of sugars
are commonly made by linking two monosaccharides to-
gether to form a disaccharide (Greek di, “two”). Disaccha-
rides serve as effective reservoirs of glucose because the
normal glucose-utilizing enzymes of the organism cannot
break the bond linking the two monosaccharide subunits.
Enzymes that can do so are typically present only in the
tissue where the glucose is to be used.
Transport forms differ depending on which monosac-
charides link to form the disaccharide. Glucose forms
transport disaccharides with itself and many other mono-
saccharides, including fructose and galactose. When glu-
cose forms a disaccharide with its structural isomer, fruc-
tose, the resulting disaccharide is sucrose, or table sugar
(figure 3.25a). Sucrose is the form in which most plants
transport glucose and the sugar that most humans (and
other animals) eat. Sugarcane is rich in sucrose, and so are
sugar beets.
When glucose is linked to its stereoisomer, galactose,
the resulting disaccharide is lactose, or milk sugar. Many
mammals supply energy to their young in the form of lac-
tose. Adults have greatly reduced levels of lactase, the en-
zyme required to cleave lactose into its two monosaccha-
ride components, and thus cannot metabolize lactose as
efficiently. Most of the energy that is channeled into lac-
tose production is therefore reserved for their offspring.
Storage Polysaccharides
Organisms store the metabolic energy contained in
monosaccharides by converting them into disaccharides,
such as maltose (figure 3.25b), which are then linked togeth-
er into insoluble forms that are deposited in specific storage
areas in their bodies. These insoluble polysaccharides are
long polymers of monosaccharides formed by dehydration
synthesis. Plant polysaccharides formed from glucose are
called starches. Plants store starch as granules within chlo-
roplasts and other organelles. Because glucose is a key met-
abolic fuel, the stored starch provides a reservoir of energy
available for future needs. Energy for cellular work can be
retrieved by hydrolyzing the links that bind the glucose
subunits together.
The starch with the simplest structure is amylose, which
is composed of many hundreds of glucose molecules linked
together in long, unbranched chains. Each linkage occurs
between the number 1 carbon of one glucose molecule and
the number 4 carbon of another, so that amylose is, in ef-
fect, a longer form of maltose. The long chains of amylose
tend to coil up in water (figure 3.26a), a property that ren-
ders amylose insoluble. Potato starch is about 20% amy-
lose. When amylose is digested by a sprouting potato plant
(or by an animal that eats a potato), enzymes first break it
into fragments of random length, which are more soluble
because they are shorter. Baking or boiling potatoes has the
same effect, breaking the chains into fragments. Another
enzyme then cuts these fragments into molecules of mal-
tose. Finally, the maltose is cleaved into two glucose mole-
cules, which cells are able to metabolize.
Most plant starch, including the remaining 80% of po-
tato starch, is a somewhat more complicated variant of
amylose called amylopectin (figure 3.26b). Pectins are
branched polysaccharides. Amylopectin has short, linear
amylose branches consisting of 20 to 30 glucose subunits.
Chapter 3 The Chemical Building Blocks of Life 55
CH
2
OH
Glucose
HO
OH
OH
O
CH
2
OH
H
2
O
+
Maltose
Glucose
HO
OH
OH
OH
O
CH
2
OH
HO
OH
OH
O
CH
2
OH
O
OH
OH
OH
O
CH
2
OH
Glucose
HO
OH
OH
OH
OH
O
CH
2
OH
CH
2
OH
H
2
O
+
Sucrose
Fructose
HO
OH
HO
O
CH
2
OH
CH
2
OH
OH
HO
O
CH
2
OH
HO
OH
OH
O
O
(a)
(b)
FIGURE 3.25
How disaccharides form.
Some disaccharides are used
to transport glucose from one
part of an organism’s body to
another; one example is
sucrose (a), which is found in
sugarcane. Other
disaccharides, such as maltose
in grain (b), are used for
storage.
In some plants these chains are cross-
linked. The cross-links create an insol-
uble mesh of glucose, which can be de-
graded only by another kind of
enzyme. The size of the mesh differs
from plant to plant; in rice about 100
amylose chains, each with one or two
cross-links, forms the mesh.
The animal version of starch is gly-
cogen. Like amylopectin, glycogen is an
insoluble polysaccharide containing
branched amylose chains. In glycogen,
the average chain length is much
greater and there are more branches
than in plant starch (figure 3.26c). Hu-
mans and other vertebrates store ex-
cess food energy as glycogen in the liv-
er and in muscle cells; when the
demand for energy in a tissue increas-
es, glycogen is hydrolyzed to release
glucose.
Nonfattening Sweets
Imagine a kind of table sugar that
looks, tastes, and cooks like the real
thing, but has no calories or harmful
side effects. You could eat mountains
of candy made from such sweeteners
without gaining weight. As Louis Pas-
teur discovered in the late 1800s, most
sugars are “right-handed” molecules,
in that the hydroxyl group that binds a
critical carbon atom is on the right
side. However, “left-handed” sugars,
in which the hydroxyl group is on the
left side, can be made readily in the
laboratory. These synthetic sugars are
mirror-image chemical twins of the
natural form, but the enzymes that
break down sugars in the human di-
gestive system can tell the difference.
To digest a sugar molecule, an enzyme
must first grasp it, much like a shoe
fitting onto a foot, and all of the
body’s enzymes are right-handed! A
left-handed sugar doesn’t fit, any more
than a shoe for the right foot fits onto
a left foot.
The Latin word for “left” is levo, and left-handed sugars
are called levo-, or 1-sugars. They do not occur in nature
except for trace amounts in red algae, snail eggs, and sea-
weed. Because they pass through the body without being
used, they can let diet-conscious sweet-lovers have their
cake and eat it, too. Nor will they contribute to tooth
decay because bacteria cannot metabolize them, either.
Starches are glucose polymers. Most starches are
branched and some are cross-linked. The branching
and cross-linking render the polymer insoluble and
protect it from degradation.
56 Part I The Origin of Living Things
(a) Amylose (b) Amylopectin
(c) Glycogen
FIGURE 3.26
Storage polysaccharides. Starches are long glucose polymers that store energy in plants.
(a) The simplest starches are long chains of maltose called amylose, which tend to coil up
in water. (b) Most plants contain more complex starches called amylopectins, which are
branched. (c) Animals store glucose in glycogen, which is more extensively branched than
amylopectin and contains longer chains of amylose.
Structural
Carbohydrates
While some chains of sugars store en-
ergy, others serve as structural material
for cells.
Cellulose
For two glucose molecules to link to-
gether, the glucose subunits must be
the same form. Glucose can form a
ring in two ways, with the hydroxyl
group attached to the carbon where the
ring closes being locked into place ei-
ther below or above the plane of the
ring. If below, it is called the alpha
form, and if above, the beta form. All
of the glucose subunits of the starch
chain are alpha-glucose. When a chain
of glucose molecules consists of all
beta-glucose subunits, a polysaccharide
with very different properties results.
This structural polysaccharide is cel-
lulose, the chief component of plant cell
walls (figure 3.27). Cellulose is chemi-
cally similar to amylose, with one im-
portant difference: the starch-degrading
enzymes that occur in most organisms
cannot break the bond between two
beta-glucose sugars. This is not because
the bond is stronger, but rather be-
cause its cleavage requires an enzyme
most organisms lack. Because cellulose
cannot be broken down readily, it works well as a biological
structural material and occurs widely in this role in plants.
Those few animals able to break down cellulose find it a
rich source of energy. Certain vertebrates, such as cows, can
digest cellulose by means of bacteria and protists they har-
bor in their intestines which provide the necessary enzymes.
Chitin
The structural building material in insects, many fungi, and
certain other organisms is called chitin (figure 3.28). Chitin
is a modified form of cellulose with a nitrogen group added
to the glucose units. When cross-linked by proteins, it
forms a tough, resistant surface material that serves as the
hard exoskeleton of arthropods such as insects and crusta-
ceans (see chapter 46). Few organisms are able to digest
chitin.
Structural carbohydrates are chains of sugars that are
not easily digested. They include cellulose in plants and
chitin in arthropods and fungi.
Chapter 3 The Chemical Building Blocks of Life 57
FIGURE 3.27
A journey into wood. This jumble of cellulose
fibers (a) is from a yellow pine (Pinus ponderosa)
(20×). (b) While starch chains consist of alpha-
glucose subunits, (c) cellulose chains consist of
beta-glucose subunits. Cellulose fibers can be
very strong and are quite resistant to metabolic
breakdown, which is one reason why wood is such
a good building material.
OH
CH
2
OH
α form of glucose Starch: chain of α-glucose subunits
HO
OH
OH H
H
HH
H
OH
O
OO
O O
41
1
1
O
O O
4
4
CH
2
OH
(b)
(c)
β form of glucose Cellulose: chain of β-glucose subunits
HO
OH
OH H
H
H
H
H
O
O
O
O
O
O
O
O
41
(a)
FIGURE 3.28
Chitin. Chitin, which might be considered to be a modified
form of cellulose, is the principal structural element in the
external skeletons of many invertebrates, such
as this lobster.
Chapter 3
Summary Questions Media Resources
3.1 Molecules are the building blocks of life.
? The chemistry of living systems is the chemistry of
carbon-containing compounds.
? Carbon’s unique chemical properties allow it to poly-
merize into chains by dehydration synthesis, forming
the four key biological macromolecules: carbohy-
drates, lipids, proteins, and nucleic acids.
58 Part I The Origin of Living Things
1. What types of molecules are
formed by dehydration reac-
tions? What types of molecules
are formed by hydrolysis?
? Organic Chemistry
? Explorations: How
Proteins Function
? Proteins
? Student Research: A
new Protein in Inserts
? Nucleic Acids
? Lipids
? Experiment:
Anfinsen: Amino Acid
Sequence Determines
Protein Shape
? Carbohydrates
2. How are amino acids linked to
form proteins?
3. Explain what is meant by the
primary, secondary, tertiary, and
quaternary structure of a protein.
? Proteins are polymers of amino acids.
? Because the 20 amino acids that occur in proteins
have side groups with different chemical properties,
the function and shape of a protein are critically af-
fected by its particular sequence of amino acids.
3.2 Proteins perform the chemistry of the cell.
4. What are the three compo-
nents of a nucleotide? How are
nucleotides linked to form nucle-
ic acids?
5. Which of the purines and py-
rimidines are capable of forming
base-pairs with each other?
? Hereditary information is stored as a sequence of nu-
cleotides in a linear polymer called DNA, which ex-
ists in cells as a double helix.
? Cells use the information in DNA by producing a
complementary single strand of RNA which directs
the synthesis of a protein whose amino acid sequence
corresponds to the nucleotide sequence of the DNA
from which the RNA was transcribed.
3.3 Nucleic acids store and transfer genetic information.
6. What are the two kinds of
subunits that make up a fat mole-
cule, and how are they arranged
in the molecule?
7. Describe the differences be-
tween a saturated and an unsat-
urated fat.
? Fats are one type of water-insoluble molecules called
lipids.
? Fats are molecules that contain many energy-rich
C—H bonds and, thus, provide an efficient form of
long-term energy storage.
? Types of lipids include phospholipids, fats, terpenes,
steroids, and prostaglandins.
3.4 Lipids make membranes and store energy.
3.5 Carbohydrates store energy and provide building materials.
? Carbohydrates store considerable energy in their
carbon-hydrogen (C—H) bonds.
? The most metabolically important carbohydrate is
glucose, a six-carbon sugar.
? Excess energy resources may be stored in complex
sugar polymers called starches (in plants) and glyco-
gen (in animals and fungi).
8. What does it mean to say that
glucose, fructose, and galactose
are isomers? Which two are
structural isomers, and how do
they differ from each other?
Which two are stereoisomers,
and how do they differ from each
other?
http://www.mhhe.com/raven6e http://www.biocourse.com
59
4
The Origin and Early
History of Life
Concept Outline
4.1 All living things share key characteristics.
What Is Life? All known organisms share certain general
properties, and to a large degree these properties define what
we mean by life.
4.2 There are many ideas about the origin of life.
Theories about the Origin of Life. There are both
religious and scientific views about the origin of life. This
text treats only the latter—only the scientifically testable.
Scientists Disagree about Where Life Started. The
atmosphere of the early earth was rich in hydrogen,
providing a ready supply of energetic electrons with which to
build organic molecules.
The Miller-Urey Experiment. Experiments attempting to
duplicate the conditions of early earth produce many of the
key molecules of living organisms.
4.3 The first cells had little internal structure.
Theories about the Origin of Cells. The first cells are
thought to have arisen spontaneously, but there is little
agreement as to the mechanism.
The Earliest Cells. The earliest fossils are of bacteria too
small to see with the unaided eye.
4.4 The first eukaryotic cells were larger and more
complex than bacteria.
The First Eukaryotic Cells. Fossils of the first eukaryotic
cells do not appear in rocks until 1.5 billion years ago, over 2
billion years after bacteria. Multicellular life is restricted to
the four eukaryotic kingdoms of life.
Has Life Evolved Elsewhere? It seems probable that life
has evolved on other worlds besides our own. The possible
presence of life in the warm waters beneath the surface of
Europa, a moon of Jupiter, is a source of current speculation.
T
here are a great many scientists with intriguing ideas
that explain how life may have originated on earth, but
there is very little that we know for sure. New hypotheses
are being proposed constantly, and old ones reevaluated. By
the time this text is published, some of the ideas presented
here about the origin of life will surely be obsolete. Thus,
the contesting ideas are presented in this chapter in an
open-ended format, attempting to make clear that there is
as yet no one answer to the question of how life originated
on earth. Although recent photographs taken by the Hubble
Space Telescope have revived controversy about the age of
the universe, it seems clear the earth itself was formed about
4.6 billion years ago. The oldest clear evidence of life—mi-
crofossils in ancient rock—are 3.5 billion years old. The ori-
gin of life seems to have taken just the right combination of
physical events and chemical processes (figure 4.1).
FIGURE 4.1
The origin of life. The fortuitous mix of physical events and
chemical elements at the right place and time created the first
living cells on earth.
Movement. One of the first things the astronauts
might do is observe the blob to see if it moves. Most
animals move about (figure 4.2), but movement from
one place to another in itself is not diagnostic of life.
Most plants and even some animals do not move about,
while numerous nonliving objects, such as clouds, do
move. The criterion of movement is thus neither neces-
sary (possessed by all life) nor sufficient (possessed only
by life).
Sensitivity. The astronauts might prod the blob to
see if it responds. Almost all living things respond to
stimuli (figure 4.3). Plants grow toward light, and
animals retreat from fire. Not all stimuli produce
responses, however. Imagine kicking a redwood tree or
singing to a hibernating bear. This criterion, although
superior to the first, is still inadequate to define life.
Death. The astronauts might attempt to kill the blob.
All living things die, while inanimate objects do not.
Death is not easily distinguished from disorder, how-
ever; a car that breaks down has not died because it was
never alive. Death is simply the loss of life, so this is a
circular definition at best. Unless one can detect life,
death is a meaningless concept, and hence a very inade-
quate criterion for defining life.
Complexity. Finally, the astronauts might cut up
the blob, to see if it is complexly organized. All living
things are complex. Even the simplest bacteria
60 Part I The Origin of Living Things
The earth formed as a hot mass of molten rock about 4.6
billion years ago. As the earth cooled, much of the water
vapor present in its atmosphere condensed into liquid water,
which accumulated on the surface in chemically rich oceans.
One scenario for the origin of life is that it originated in this
dilute, hot smelly soup of ammonia, formaldehyde, formic
acid, cyanide, methane, hydrogen sulfide, and organic hy-
drocarbons. Whether at the oceans’ edge, in hydrothermal
deep-sea vents, or elsewhere, the consensus among re-
searchers is that life arose spontaneously from these early
waters less than 4 billion years ago. While the way in which
this happened remains a puzzle, one cannot escape a certain
curiosity about the earliest steps that eventually led to the
origin of all living things on earth, including ourselves. How
did organisms evolve from the complex molecules that
swirled in the early oceans?
What Is Life?
Before we can address this question, we must first consider
what qualifies something as “living.” What is life? This is a
difficult question to answer, largely because life itself is not a
simple concept. If you try to write a definition of “life,” you
will find that it is not an easy task, because of the loose man-
ner in which the term is used.
Imagine a situation in which two astronauts encounter a
large, amorphous blob on the surface of a planet. How
would they determine whether it is alive?
4.1 All living things share key characteristics.
FIGURE 4.2
Movement. Animals have evolved
mechanisms that allow them to move about
in their environment. While some animals,
like this giraffe, move on land, others move
through water or air.
FIGURE 4.3
Sensitivity. This father lion is responding to a stimulus: he has just been bitten on the
rump by his cub. As far as we know, all organisms respond to stimuli, although not always
to the same ones or in the same way. Had the cub bitten a tree instead of its father, the
response would not have been as dramatic.
contain a bewildering array of molecules, organized
into many complex structures. However a computer is
also complex, but not alive. Complexity is a necessary
criterion of life, but it is not sufficient in itself to iden-
tify living things because many complex things are not
alive.
To determine whether the blob is alive, the astronauts
would have to learn more about it. Probably the best thing
they could do would be to examine it more carefully and de-
termine whether it resembles the organisms we are familiar
with, and if so, how.
Fundamental Properties of Life
As we discussed in chapter 1, all known organisms share cer-
tain general properties. To a large degree, these properties
define what we mean by life. The following fundamental
properties are shared by all organisms on earth.
Cellular organization. All organisms consist of one
or more cells—complex, organized assemblages of mol-
ecules enclosed within membranes (figure 4.4).
Sensitivity. All organisms respond to stimuli—
though not always to the same stimuli in the same
ways.
Growth. All living things assimilate energy and use it
to grow, a process called metabolism. Plants, algae, and
some bacteria use sunlight to create covalent carbon-
carbon bonds from CO
2
and H
2
O through photosyn-
thesis. This transfer of the energy in covalent bonds is
essential to all life on earth.
Development. Multicellular organisms undergo
systematic gene-directed changes as they grow and
mature.
Reproduction. All living things reproduce, passing
on traits from one generation to the next. Although
some organisms live for a very long time, no organism
lives forever, as far as we know. Because all organisms
die, ongoing life is impossible without reproduction.
Regulation. All organisms have regulatory mecha-
nisms that coordinate internal processes.
Homeostasis. All living things maintain relatively
constant internal conditions, different from their envi-
ronment.
The Key Role of Heredity
Are these properties adequate to define life? Is a
membrane-enclosed entity that grows and reproduces
alive? Not necessarily. Soap bubbles and proteinoid
microspheres spontaneously form hollow bubbles that
enclose a small volume of water. These spheres can
enclose energy-processing molecules, and they may also
grow and subdivide. Despite these features, they are
certainly not alive. Therefore, the criteria just listed,
although necessary for life, are not sufficient to define life.
One ingredient is missing—a mechanism for the preserva-
tion of improvement.
Heredity. All organisms on earth possess a genetic
system that is based on the replication of a long, complex
molecule called DNA. This mechanism allows for
adaptation and evolution over time, also distinguishing
characteristics of living things.
To understand the role of heredity in our definition of
life, let us return for a moment to proteinoid microspheres.
When we examine an individual microsphere, we see it at
that precise moment in time but learn nothing of its prede-
cessors. It is likewise impossible to guess what future
droplets will be like. The droplets are the passive prisoners
of a changing environment, and it is in this sense that they
are not alive. The essence of being alive is the ability to en-
compass change and to reproduce the results of change
permanently. Heredity, therefore, provides the basis for the
great division between the living and the nonliving.
Change does not become evolution unless it is passed on to
a new generation. A genetic system is the sufficient condi-
tion of life. Some changes are preserved because they
increase the chances of survival in a hostile world, while
others are lost. Not only did life evolve—evolution is the
very essence of life.
All living things on earth are characterized by cellular
organization, heredity, and a handful of other
characteristics that serve to define the term life.
Chapter 4 The Origin and Early History of Life 61
FIGURE 4.4
Cellular organization (150×). These Paramecia, which are
complex, single-celled organisms called protists, have just
ingested several yeast cells. The yeasts, stained red in this
photograph, are enclosed within membrane-bounded sacs
called digestive vacuoles. A variety of other organelles are also
visible.
Theories about the
Origin of Life
The question of how life originated
is not easy to answer because it is
impossible to go back in time and
observe life’s beginnings; nor are
there any witnesses. There is testi-
mony in the rocks of the earth, but it
is not easily read, and often it is
silent on issues crying out for an-
swers. There are, in principle, at
least three possibilities:
1. Special creation. Life-forms
may have been put on earth by
supernatural or divine forces.
2. Extraterrestrial origin. Life
may not have originated on
earth at all; instead, life may
have infected earth from some
other planet.
3. Spontaneous origin. Life may
have evolved from inanimate
matter, as associations among
molecules became more and
more complex.
Special Creation. The theory of
special creation, that a divine God
created life is at the core of most
major religions. The oldest hypothesis
about life’s origins, it is also the most
widely accepted. Far more Americans,
for example, believe that God created
life on earth than believe in the other
two hypotheses. Many take a more ex-
treme position, accepting the biblical
account of life’s creation as factually
correct. This viewpoint forms the basis for the very
unscientific “scientific creationism” viewpoint discussed
in chapter 21.
Extraterrestrial Origin. The theory of panspermia
proposes that meteors or cosmic dust may have carried
significant amounts of complex organic molecules to
earth, kicking off the evolution of life. Hundreds of thou-
sands of meteorites and comets are known to have
slammed into the early earth, and recent findings suggest
that at least some may have carried organic materials.
Nor is life on other planets ruled out. For example, the
discovery of liquid water under the surface of Jupiter’s
ice-shrouded moon Europa and suggestions of fossils in
rocks from Mars lend some credence
to this idea. The hypothesis that an
early source of carbonaceous material
is extraterrestrial is testable, although
it has not yet been proven. Indeed,
NASA is planning to land on Europa,
drill through the surface, and send a
probe down to see if there is life.
Spontaneous Origin. Most scien-
tists tentatively accept the theory of
spontaneous origin, that life evolved
from inanimate matter. In this view,
the force leading to life was selec-
tion. As changes in molecules in-
creased their stability and caused
them to persist longer, these mole-
cules could initiate more and more
complex associations, culminating in
the evolution of cells.
Taking a Scientific Viewpoint
In this book we will focus on the sec-
ond and third possibilities, attempting
to understand whether natural forces
could have led to the origin of life
and, if so, how the process might have
occurred. This is not to say that the
first possibility is definitely not the
correct one. Any one of the three pos-
sibilities might be true. Nor do the
second and third possibilities preclude
religion (a divine agency might have
acted via evolution, for example).
However, we are limiting the scope of
our inquiry to scientific matters, and
only the second and third possibilities
permit testable hypotheses to be
constructed—that is, explanations
that can be tested and potentially disproved.
In our search for understanding, we must look back to
the early times. There are fossils of simple living things,
bacteria, in rocks 3.5 billion years old. They tell us that
life originated during the first billion years of the history
of our planet. As we attempt to determine how this
process took place, we will first focus on how organic
molecules may have originated (figure 4.5), and then we
will consider how those molecules might have become
organized into living cells.
Panspermia and spontaneous origin are the only testable
hypotheses of life’s origin currently available.
62 Part I The Origin of Living Things
4.2 There are many ideas about the origin of life.
FIGURE 4.5
Lightning. Before life evolved, the simple
molecules in the earth’s atmosphere
combined to form more complex molecules.
The energy that drove these chemical
reactions may have come from lightning
and forms of geothermal energy.
Scientists Disagree about Where
Life Started
While most researchers agree that life first appeared as the
primitive earth cooled and its rocky crust formed, there is
little agreement as to just where this occurred.
Did Life Originate at the Ocean’s Edge?
The more we learn about earth’s early history, the more
likely it seems that earth’s first organisms emerged and lived
at very high temperatures. Rubble from the forming solar
system slammed into early earth from 4.6 to 3.8 billion years
ago, keeping the surface molten hot. As the bombardment
slowed down, temperatures dropped. By about 3.8 billion
years ago, ocean temperatures are thought to have dropped
to a hot 49° to 88°C (120° to 190°F). Between 3.8 and 3.5
billion years ago, life first appeared, promptly after the earth
was inhabitable. Thus, as intolerable as early earth’s infernal
temperatures seem to us today, they gave birth to life.
Very few geochemists agree on the exact composition of
the early atmosphere. One popular view is that it contained
principally carbon dioxide (CO
2
) and nitrogen gas (N
2
),
along with significant amounts of water vapor (H
2
O). It is
possible that the early atmosphere also contained hydrogen
gas (H
2
) and compounds in which hydrogen atoms were
bonded to the other light elements (sulfur, nitrogen, and
carbon), producing hydrogen sulfide (H
2
S), ammonia
(NH
3
), and methane (CH
4
).
We refer to such an atmosphere as a reducing atmosphere
because of the ample availability of hydrogen atoms and
their electrons. In such a reducing atmosphere it would not
take as much energy as it would today to form the carbon-
rich molecules from which life evolved.
The key to this reducing atmosphere hypothesis is the
assumption that there was very little oxygen around. In an
atmosphere with oxygen, amino acids and sugars react
spontaneously with the oxygen to form carbon dioxide and
water. Therefore, the building blocks of life, the amino
acids, would not last long and the spontaneous formation of
complex carbon molecules could not occur. Our atmosphere
changed once organisms began to carry out photosynthesis,
harnessing the energy in sunlight to split water molecules
and form complex carbon molecules, giving off gaseous oxy-
gen molecules in the process. The earth’s atmosphere is
now approximately 21% oxygen.
Critics of the reducing atmosphere hypothesis point out
that no carbonates have been found in rocks dating back to
the early earth. This suggests that at that time carbon diox-
ide was locked up in the atmosphere, and if that was the
case, then the prebiotic atmosphere would not have been re-
ducing.
Another problem for the reducing atmosphere hypothe-
sis is that because a prebiotic reducing atmosphere would
have been oxygen free, there would have been no ozone.
Without the protective ozone layer, any organic compounds
that might have formed would have been broken down
quickly by ultraviolet radiation.
Other Suggestions
If life did not originate at the ocean’s edge under the blanket
of a reducing atmosphere, where did it originate?
Under frozen oceans. One hypothesis proposes that
life originated under a frozen ocean, not unlike the one
that covers Jupiter’s moon Europa today. All evidence
suggests, however, that the early earth was quite warm
and frozen oceans quite unlikely.
Deep in the earth’s crust. Another hypothesis is that
life originated deep in the earth’s crust. In 1988 Gunter
Wachtershauser proposed that life might have formed as
a by-product of volcanic activity, with iron and nickel
sulfide minerals acting as chemical catalysts to recom-
bine gases spewing from eruptions into the building
blocks of life. In later work he and coworkers were able
to use this unusual chemistry to build precursors for
amino acids (although they did not actually succeed in
making amino acids), and to link amino acids together to
form peptides. Critics of this hypothesis point out that
the concentration of chemicals used in their experiments
greatly exceed what is found in nature.
Within clay. Other researchers have proposed the un-
usual hypothesis that life is the result of silicate surface
chemistry. The surface of clays have positive charges to
attract organic molecules, and exclude water, providing a
potential catalytic surface on which life’s early chemistry
might have occurred. While interesting conceptually,
there is little evidence that this sort of process could ac-
tually occur.
At deep-sea vents. Becoming more popular is the hy-
pothesis that life originated at deep-sea hydrothermal
vents, with the necessary prebiotic molecules being syn-
thesized on metal sulfides in the vents. The positive
charge of the sulfides would have acted as a magnet for
negatively charged organic molecules. In part, the cur-
rent popularity of this hypothesis comes from the new
science of genomics, which suggests that the ancestors of
today’s prokaryotes are most closely related to the bacte-
ria that live on the deep-sea vents.
No one is sure whether life originated at the ocean’s
edge, under frozen ocean, deep in the earth’s crust, within
clay, or at deep-sea vents. Perhaps one of these hypotheses
will be proven correct. Perhaps the correct theory has not
yet been proposed.
When life first appeared on earth, the environment was
very hot. All of the spontaneous origin hypotheses
assume that the organic chemicals that were the
building blocks of life arose spontaneously at that time.
How is a matter of considerable disagreement.
Chapter 4 The Origin and Early History of Life 63
The Miller-Urey Experiment
An early attempt to see what kinds of organic molecules
might have been produced on the early earth was carried
out in 1953 by Stanley L. Miller and Harold C. Urey. In
what has become a classic experiment, they attempted to
reproduce the conditions at ocean’s edge under a reducing
atmosphere. Even if this assumption proves incorrect—
the jury is still out on this—their experiment is critically
important, as it ushered in the whole new field of prebi-
otic chemistry.
To carry out their experiment, they (1) assembled a re-
ducing atmosphere rich in hydrogen and excluding gaseous
oxygen; (2) placed this atmosphere over liquid water, which
would have been present at ocean’s edge; (3) maintained this
mixture at a temperature somewhat below 100°C; and
(4) simulated lightning by bombarding it with energy in the
form of sparks (figure 4.6).
They found that within a week, 15% of the carbon origi-
nally present as methane gas (CH
4
) had converted into
other simple carbon compounds. Among these compounds
were formaldehyde (CH
2
O) and hydrogen cyanide (HCN;
figure 4.7). These compounds then combined to form sim-
ple molecules, such as formic acid (HCOOH) and urea
(NH
2
CONH
2
), and more complex molecules containing
carbon-carbon bonds, including the amino acids glycine and
alanine.
As we saw in chapter 3, amino acids are the basic build-
ing blocks of proteins, and proteins are one of the major
kinds of molecules of which organisms are composed. In
similar experiments performed later by other scientists,
more than 30 different carbon compounds were identified,
including the amino acids glycine, alanine, glutamic acid,
valine, proline, and aspartic acid. Other biologically impor-
tant molecules were also formed in these experiments. For
example, hydrogen cyanide contributed to the production
of a complex ring-shaped molecule called adenine—one of
the bases found in DNA and RNA. Thus, the key mole-
cules of life could have formed in the atmosphere of the
early earth.
The Path of Chemical Evolution
A raging debate among biologists who study the origin of
life concerns which organic molecules came first, RNA or
proteins. Scientists are divided into three camps, those that
focus on RNA, protein, or a combination of the two. All
three arguments have their strong points. Like the hypothe-
ses that try to account for where life originated, these com-
peting hypotheses are diverse and speculative.
An RNA World. The “RNA world” group feels that with-
out a hereditary molecule, other molecules could not have
formed consistently. The “RNA world” argument earned
support when Thomas Cech at the University of Colorado
discovered ribozymes, RNA molecules that can behave as
enzymes, catalyzing their own assembly. Recent work has
shown that the RNA contained in ribosomes (discussed in
chapter 5) catalyzes the chemical reaction that links amino
acids to form proteins. Therefore, the RNA in ribosomes
also functions as an enzyme. If RNA has the ability to pass
on inherited information and the capacity to act like an
enzyme, were proteins really needed?
A Protein World. The “protein-first” group argues that
without enzymes (which are proteins), nothing could
replicate at all, heritable or not. The “protein-first” pro-
ponents argue that nucleotides, the individual units of
nucleic acids such as RNA, are too complex to have
formed spontaneously, and certainly too complex to form
spontaneously again and again. While there is no doubt
that simple proteins are easier to synthesize from abiotic
components than nucleotides, both can form in the labo-
ratory under the right conditions. Deciding which came
first is a chicken-and-egg paradox. In an effort to shed
light on this problem, Julius Rebek and a number of
64 Part I The Origin of Living Things
Water
vapor
Condensed liquid
with complex
molecules
Stopcocks
for testing
samples
Condensor
Mixture
of gases
("primitive
atmosphere")
Heated water
("ocean")
Electrodes
discharge sparks
(lightning
simulation)
Water
FIGURE 4.6
The Miller-Urey experiment. The apparatus consisted of a
closed tube connecting two chambers. The upper chamber
contained a mixture of gases thought to resemble the primitive
earth’s atmosphere. Electrodes discharged sparks through this
mixture, simulating lightning. Condensers then cooled the gases,
causing water droplets to form, which passed into the second
heated chamber, the “ocean.” Any complex molecules formed in
the atmosphere chamber would be dissolved in these droplets and
carried to the ocean chamber, from which samples were
withdrawn for analysis.
other chemists have created synthetic nucleotide-like
molecules in the laboratory that are able to replicate.
Moving even further, Rebek and his colleagues have cre-
ated synthetic molecules that could replicate and “make
mistakes.” This simulates mutation, a necessary ingredi-
ent for the process of evolution.
A Peptide-Nucleic Acid World. Another important and
popular theory about the first organic molecules assumes
key roles for both peptides and nucleic acids. Because RNA
is so complex and unstable, this theory assumes there must
have been a pre-RNA world where the peptide-nucleic acid
(PNA) was the basis for life. PNA is stable and simple
enough to have formed spontaneously, and is also a self-
replicator.
Molecules that are the building blocks of living
organisms form spontaneously under conditions
designed to simulate those of the primitive earth.
Chapter 4 The Origin and Early History of Life 65
Water
Aldehydes
Proprionic acid
Lactic
acid
Glycolic
acid
Succinic
acid
Glycine
Alanine β-Alanine
β-Aminobutyric acid α-Aminobutyric acidN-Methylalanine
Valine
Proline
Aspartic acid
Glutamic acid
Raw
materials
First group of
intermediate products
Second group of
intermediate products
Examples of final products (isomers are boxed)
Immunoacetic propionic acid
Iminodiacetic acid
Norvaline Isovaline
Sarcosine
Acetic acid
Formic acid
N-Methylurea
Urea
Energy Energy
Hydrogen
cyanide
Nitrogen
Ammonia
Carbon
dioxide
Carbon
monoxide
Methane
Hydrogen
gas
Energy
FIGURE 4.7
Results of the Miller-Urey experiment. Seven simple molecules, all gases, were included in the original mixture. Note that oxygen was
not among them; instead, the atmosphere was rich in hydrogen. At each stage of the experiment, more complex molecules were formed:
first aldehydes, then simple acids, then more complex acids. Among the final products, the molecules that are structural isomers of one
another are grouped together in boxes. In most cases only one isomer of a compound is found in living systems today, although many may
have been produced in the Miller-Urey experiment.
Theories about the Origin of Cells
The evolution of cells required early organic molecules
to assemble into a functional, interdependent unit. Cells,
discussed in the next chapter, are essentially little bags of
fluid. What the fluid contains depends on the individual
cell, but every cell’s contents differ from the environ-
ment outside the cell. Therefore, an early cell may have
floated along in a dilute “primordial soup,” but its inte-
rior would have had a higher concentration of specific
organic molecules.
Cell Origins: The Importance of Bubbles
How did these “bags of fluid” evolve from simple organic
molecules? As you can imagine, the answer to this ques-
tion is a matter for debate. Scientists favoring an “ocean’s
edge” scenario for the origin of life have proposed that
bubbles may have played a key role in this evolutionary
step. A bubble, such as those produced by soap solutions,
is a hollow spherical structure. Certain molecules, partic-
ularly those with hydrophobic regions, will spontaneously
form bubbles in water. The structure of the bubble shields
the hydrophobic regions of the molecules from contact
with water. If you have ever watched the ocean surge
upon the shore, you may have noticed the foamy froth
created by the agitated water. The edges of the primitive
oceans were more than likely very frothy places bom-
barded by ultraviolet and other ionizing radiation, and ex-
posed to an atmosphere that may have contained methane
and other simple organic molecules.
Oparin’s Bubble Theory
The first bubble theory is attributed to Alexander
Oparin, a Russian chemist with extraordinary insight. In
the mid-1930s, Oparin suggested that the present-day at-
mosphere was incompatible with the creation of life. He
proposed that life must have arisen from nonliving matter
under a set of very different environmental circumstances
some time in the distant history of the earth. His was the
theory of primary abiogenesis (primary because all liv-
ing cells are now known to come from previously living
cells, except in that first case). At the same time, J. B. S.
Haldane, a British geneticist, was also independently es-
pousing the same views. Oparin decided that in order for
cells to evolve, they must have had some means of devel-
oping chemical complexity, separating their contents
from their environment by means of a cell membrane,
and concentrating materials within themselves. He
termed these early, chemical-concentrating bubblelike
structures protobionts.
Oparin’s theories were published in English in 1938,
and for awhile most scientists ignored them. However,
Harold Urey, an astronomer at the University of
Chicago, was quite taken with Oparin’s ideas. He con-
vinced one of his graduate students, Stanley Miller, to
follow Oparin’s rationale and see if he could “create” life.
The Urey-Miller experiment has proven to be one of the
most significant experiments in the history of science. As
a result Oparin’s ideas became better known and more
widely accepted.
A Host of Bubble Theories
Different versions of “bubble theories” have been cham-
pioned by numerous scientists since Oparin. The bubbles
they propose go by a variety of names; they may be called
microspheres, protocells, protobionts, micelles, liposomes, or
coacervates, depending on the composition of the bubbles
(lipid or protein) and how they form. In all cases, the
bubbles are hollow spheres, and they exhibit a variety of
cell-like properties. For example, the lipid bubbles called
coacervates form an outer boundary with two layers that
resembles a biological membrane. They grow by accumu-
lating more subunit lipid molecules from the surrounding
medium, and they can form budlike projections and
divide by pinching in two, like bacteria. They also can
contain amino acids and use them to facilitate various
acid-base reactions, including the decomposition of
glucose. Although they are not alive, they obviously have
many of the characteristics of cells.
A Bubble Scenario
It is not difficult to imagine that a process of chemical evo-
lution involving bubbles or microdrops preceded the origin
of life (figure 4.8). The early oceans must have contained
untold numbers of these microdrops, billions in a spoonful,
each one forming spontaneously, persisting for a while, and
then dispersing. Some would, by chance, have contained
amino acids with side groups able to catalyze growth-
promoting reactions. Those microdrops would have
survived longer than ones that lacked those amino acids,
because the persistence of both proteinoid microspheres
and lipid coacervates is greatly increased when they carry
out metabolic reactions such as glucose degradation and
when they are actively growing.
Over millions of years, then, the complex bubbles that
were better able to incorporate molecules and energy from
the lifeless oceans of the early earth would have tended to
persist longer than the others. Also favored would have been
the microdrops that could use these molecules to expand in
size, growing large enough to divide into “daughter”
66 Part I The Origin of Living Things
4.3 The first cells had little internal structure.
microdrops with features similar to those of their “parent”
microdrop. The daughter microdrops have the same favor-
able combination of characteristics as their parent, and
would have grown and divided, too. When a way to facilitate
the reliable transfer of new ability from parent to offspring
developed, heredity—and life—began.
Current Thinking
Whether the early bubbles that gave rise to cells were lipid
or protein remains an unresolved argument. While it is
true that lipid microspheres (coacervates) will form readily
in water, there appears to be no mechanism for their heri-
table replication. On the other hand, one can imagine a
heritable mechanism for protein microspheres. Although
protein microspheres do not form readily in water, Sidney
Fox and his colleagues at the University of Miami have
shown that they can form under dry conditions.
The discovery that RNA can act as an enzyme to assem-
ble new RNA molecules on an RNA template has raised
the interesting possibility that neither coacervates nor pro-
tein microspheres were the first step in the evolution of
life. Perhaps the first components were RNA molecules,
and the initial steps on the evolutionary journey led to
increasingly complex and stable RNA molecules. Later,
stability might have improved further when a lipid (or
possibly protein) microsphere surrounded the RNA. At
present, those studying this problem have not arrived at a
consensus about whether RNA evolved before or after a
bubblelike structure that likely preceded cells.
Eventually, DNA took the place of RNA as the replicator
in the cell and the storage molecule for genetic information.
DNA, because it is a double helix, stores information in a
more stable fashion than RNA, which is single-stranded.
Little is known about how the first cells originated.
Current hypotheses involve chemical evolution within
bubbles, but there is no general agreement about their
composition, or about how the process occurred.
Chapter 4 The Origin and Early History of Life 67
1. Volcanoes erupted under
the sea, releasing gases
enclosed in bubbles.
2. The gases, concentrated
inside the bubbles, reacted
to produce simple organic
molecules.
3. When the bubbles persisted
long enough to rise to the surface,
they popped, releasing their
contents to the air.
4. Bombarded by the sun's ultraviolet
radiation, lightning, and other energy
sources, the simple organic molecules
released from the bubbles reacted to
form more complex organic molecules.
5. The more complex organic
molecules fell back into the
sea in raindrops. There, they
could again be enclosed in
bubbles and begin the
process again.
FIGURE 4.8
A current bubble hypothesis. In 1986 geophysicist Louis Lerman proposed that the chemical processes leading to the evolution of life
took place within bubbles on the ocean’s surface.
The Earliest Cells
What do we know about the earliest life-forms? The fossils
found in ancient rocks show an obvious progression from
simple to complex organisms, beginning about 3.5 billion
years ago. Life may have been present earlier, but rocks of
such great antiquity are rare, and fossils have not yet been
found in them.
Microfossils
The earliest evidence of life appears in microfossils, fos-
silized forms of microscopic life (figure 4.9). Microfossils
were small (1 to 2 micrometers in diameter) and single-
celled, lacked external appendages, and had little evidence of
internal structure. Thus, they physically resemble present-
day bacteria (figure 4.10), although some ancient forms
cannot be matched exactly. We call organisms with this sim-
ple body plan prokaryotes, from the Greek words meaning
“before” and “kernel,” or “nucleus.” The name reflects their
lack of a nucleus, a spherical organelle characteristic of the
more complex cells of eukaryotes.
Judging from the fossil record, eukaryotes did not appear
until about 1.5 billion years ago. Therefore, for at least 2
billion years—nearly a half of the age of the earth—bacteria
were the only organisms that existed.
Ancient Bacteria: Archaebacteria
Most organisms living today are adapted to the relatively
mild conditions of present-day earth. However, if we
look in unusual environments, we encounter organisms
that are quite remarkable, differing in form and metabo-
lism from other living things. Sheltered from evolution-
ary alteration in unchanging habitats that resemble
earth’s early environment, these living relics are the sur-
viving representatives of the first ages of life on earth. In
places such as the oxygen-free depths of the Black Sea or
the boiling waters of hot springs and deep-sea vents, we
can find bacteria living at very high temperatures without
oxygen.
These unusual bacteria are called archaebacteria, from
the Greek word for “ancient ones.” Among the first to be
studied in detail have been the methanogens, or
methane-producing bacteria, among the most primitive
bacteria that exist today. These organisms are typically
simple in form and are able to grow only in an oxygen-free
environment; in fact, oxygen poisons them. For this reason
they are said to grow “without air,” or anaerobically
(Greek an, “without” + aer, “air” + bios, “life”). The
methane-producing bacteria convert CO
2
and H
2
into
methane gas (CH
4
). Although primitive, they resemble all
other bacteria in having DNA, a lipid cell membrane, an
exterior cell wall, and a metabolism based on an energy-
carrying molecule called ATP.
68 Part I The Origin of Living Things
FIGURE 4.9
Cross-sections of fossil bacteria. These microfossils from the
Bitter Springs formation of Australia are of ancient cyanobacteria,
far too small to be seen with the unaided eye. In this electron
micrograph, the cell walls are clearly evident.
FIGURE 4.10
The oldest microfossil. This ancient bacterial fossil, discovered
by J. William Schopf of UCLA in 3.5-billion-year-old rocks in
western Australia, is similar to present-day cyanobacteria, as you
can see by comparing it to figure 4.11.
Unusual Cell Structures
When the details of cell wall and membrane structure of
the methane-producing bacteria were examined, they
proved to be different from those of all other bacteria. Ar-
chaebacteria are characterized by a conspicuous lack of a
protein cross-linked carbohydrate material called peptido-
glycan in their cell walls, a key compound in the cell walls
of most modern bacteria. Archaebacteria also have unusual
lipids in their cell membranes that are not found in any
other group of organisms. There are also major differences
in some of the fundamental biochemical processes of me-
tabolism, different from those of all other bacteria. The
methane-producing bacteria are survivors from an earlier
time when oxygen gas was absent.
Earth’s First Organisms?
Other archaebacteria that fall into this classification are
some of those that live in very salty environments like the
Dead Sea (extreme halophiles—“salt lovers”) or very hot
environments like hydrothermal volcanic vents under the
ocean (extreme thermophiles—“heat lovers”). Ther-
mophiles have been found living comfortably in boiling
water. Indeed, many kinds of thermophilic archaebacteria
thrive at temperatures of 110°C (230°F). Because these
thermophiles live at high temperatures similar to those that
may have existed when life first evolved, microbiologists
speculate that thermophilic archaebacteria may be relics of
earth’s first organisms.
Just how different are extreme thermophiles from other
organisms? A methane-producing archaebacteria called
Methanococcus isolated from deep-sea vents provides a star-
tling picture. These bacteria thrive at temperatures of 88°C
(185°F) and crushing pressures 245 times greater than at
sea level. In 1996 molecular biologists announced that they
had succeeded in determining the full nucleotide sequence
of Methanococcus. This was possible because archaebacterial
DNA is relatively small—it has only 1700 genes, coded in a
DNA molecule only 1,739,933 nucleotides long (a human
cell has 2000 times more!). The thermophile nucleotide se-
quence proved to be astonishingly different from the DNA
sequence of any other organism ever studied; fully two-
thirds of its genes are unlike any ever known to science be-
fore! Clearly these archaebacteria separated from other life
on earth a long time ago. Preliminary comparisons to the
gene sequences of other bacteria suggest that archaebacte-
ria split from other types of bacteria over 3 billion years
ago, soon after life began.
Eubacteria
The second major group of bacteria, the eubacteria, have
very strong cell walls and a simpler gene architecture. Most
bacteria living today are eubacteria. Included in this group
are bacteria that have evolved the ability to capture the en-
ergy of light and transform it into the energy of chemical
bonds within cells. These organisms are photosynthetic, as
are plants and algae.
One type of photosynthetic eubacteria that has been im-
portant in the history of life on earth is the cyanobacteria,
sometimes called “blue-green algae” (figure 4.11). They
have the same kind of chlorophyll pigment that is most
abundant in plants and algae, as well as other pigments
that are blue or red. Cyanobacteria produce oxygen as a
result of their photosynthetic activities, and when they
appeared at least 3 billion years ago, they played a decisive
role in increasing the concentration of free oxygen in the
earth’s atmosphere from below 1% to the current level of
21%. As the concentration of oxygen increased, so did the
amount of ozone in the upper layers of the atmosphere.
The thickening ozone layer afforded protection from
most of the ultraviolet radiation from the sun, radiation
that is highly destructive to proteins and nucleic acids.
Certain cyanobacteria are also responsible for the accu-
mulation of massive limestone deposits.
All bacteria now living are members of either
Archaebacteria or Eubacteria.
Chapter 4 The Origin and Early History of Life 69
FIGURE 4.11
Living cyanobacteria. Although not multicellular, these bacteria
often aggregate into chains such as those seen here.
All fossils more than 1.5 billion years old are generally sim-
ilar to one another structurally. They are small, simple
cells; most measure 0.5 to 2 micrometers in diameter, and
none are more than about 6 micrometers in diameter.
These simple cells eventually evolved into larger, more
complex forms—the first eukaryotic cells.
The First Eukaryotic Cells
In rocks about 1.5 billion years old, we begin to see the first
microfossils that are noticeably different in appearance
from the earlier, simpler forms (figure 4.12). These cells
are much larger than bacteria and have internal membranes
and thicker walls. Cells more than 10 micrometers in diam-
eter rapidly increased in abundance. Some fossilized cells
1.4 billion years old are as much as 60 micrometers in di-
ameter; others, 1.5 billion years old, contain what appear to
be small, membrane-bound structures. Indirect chemical
traces hint that eukaryotes may go as far back as 2.7 billion
years, although no fossils as yet support such an early ap-
pearance of eukaryotes.
These early fossils mark a major event in the evolution
of life: a new kind of organism had appeared (figure 4.13).
These new cells are called eukaryotes, from the Greek
words for “true” and “nucleus,” because they possess an in-
ternal structure called a nucleus. All organisms other than
the bacteria are eukaryotes.
Origin of the Nucleus and ER
Many bacteria have infoldings of their outer membranes
extending into the cytoplasm and serving as passageways
to the surface. The network of internal membranes in
70 Part I The Origin of Living Things
4.4 The first eukaryotic cells were larger and more complex than bacteria.
100 μm
FIGURE 4.12
Microfossil of a primitive eukaryote. This multicellular alga is
between 900 million and 1 billion years old.
Geological evidence Life forms
PRECAMBRIAN
CAMBRIAN
PHANEROZOIC
PROTEROZOIC
ARCHEAN
Appearance of first multicellular organisms
Appearance of first eukaryotes
Appearance of aerobic (oxygen-using) respiration
Appearance of oxygen-forming photosynthesis
(cyanobacteria)
Appearance of chemoautotrophs (sulfate respiration)
Appearance of life (prokaryotes): anaerobic
(methane-producing) bacteria and anaerobic (hydrogen
sulfide–forming) photosynthesis
Formation of the earth
Oldest multicellular fossils
Oldest compartmentalized fossil cells
Disappearance of iron from oceans and
formation of iron oxides
Oldest definite fossils
Oldest dated rocks
Millions
of years
ago
570
600
1500
2500
3500
4500
FIGURE 4.13
The geological timescale. The periods refer to different stages in the evolution of life on earth. The timescale is calibrated by examining
rocks containing particular kinds of fossils; the fossils are dated by determining the degree of spontaneous decay of radioactive isotopes
locked within rock when it was formed.
eukaryotes called endoplasmic reticulum (ER) is thought
to have evolved from such infoldings, as is the nuclear en-
velope, an extension of the ER network that isolates and
protects the nucleus (figure 4.14).
Origin of Mitochondria and Chloroplasts
Bacteria that live within other cells and perform specific
functions for their host cells are called endosymbiotic bacte-
ria. Their widespread presence in nature led Lynn
Margulis to champion the endosymbiotic theory in the
early 1970s. This theory, now widely accepted, suggests
that a critical stage in the evolution of eukaryotic cells in-
volved endosymbiotic relationships with prokaryotic or-
ganisms. According to this theory, energy-producing
bacteria may have come to reside within larger bacteria,
eventually evolving into what we now know as mitochon-
dria. Similarly, photosynthetic bacteria may have come to
live within other larger bacteria, leading to the evolution
of chloroplasts, the photosynthetic organelles of plants
and algae. Bacteria with flagella, long whiplike cellular
appendages used for propulsion, may have become sym-
biotically involved with nonflagellated bacteria to pro-
duce larger, motile cells. The fact that we now witness so
many symbiotic relationships lends general support to
this theory. Even stronger support comes from the obser-
vation that present-day organelles such as mitochondria,
chloroplasts, and centrioles contain their own DNA,
which is remarkably similar to the DNA of bacteria in
size and character.
Sexual Reproduction
Eukaryotic cells also possess the ability to reproduce sexu-
ally, something prokaryotes cannot do effectively. Sexual
reproduction is the process of producing offspring, with
two copies of each chromosome, by fertilization, the union
of two cells that each have one copy of each chromosome.
The great advantage of sexual reproduction is that it allows
for frequent genetic recombination, which generates the
variation that is the raw material for evolution. Not all eu-
karyotes reproduce sexually, but most have the capacity to
do so. The evolution of meiosis and sexual reproduction
(discussed in chapter 12) led to the tremendous explosion
of diversity among the eukaryotes.
Multicellularity
Diversity was also promoted by the development of multi-
cellularity. Some single eukaryotic cells began living in as-
sociation with others, in colonies. Eventually individual
members of the colony began to assume different duties,
and the colony began to take on the characteristics of a sin-
gle individual. Multicellularity has arisen many times
among the eukaryotes. Practically every organism big
enough to see with the unaided eye is multicellular, includ-
ing all animals and plants. The great advantage of multicel-
lularity is that it fosters specialization; some cells devote all
of their energies to one task, other cells to another. Few in-
novations have had as great an impact on the history of life
as the specialization made possible by multicellularity.
Chapter 4 The Origin and Early History of Life 71
Infolding of the
plasma membrane
DNA
Cell wall
Bacterial cell
Prokaryotic ancestor
of eukaryotic cells
Eukaryotic cell
Endoplasmic
reticulum (ER)
Nuclear envelope
Nucleus
FIGURE 4.14
Origin of the nucleus and endoplasmic reticulum. Many bacteria today have infoldings of the plasma membrane (see also figure 34.7).
The eukaryotic internal membrane system called the endoplasmic reticulum (ER) and the nuclear envelope may have evolved from such
infoldings of the plasma membrane encasing prokaryotic cells that gave rise to eukaryotic cells.
The Kingdoms of Life
Confronted with the great diversity of life on earth today,
biologists have attempted to categorize similar organisms
in order to better understand them, giving rise to the sci-
ence of taxonomy. In later chapters, we will discuss tax-
onomy and classification in detail, but for now we can
generalize that all living things fall into one of three
domains which include six kingdoms (figure 4.15):
Kingdom Archaebacteria: Prokaryotes that lack a
peptidoglycan cell wall, including the methanogens and
extreme halophiles and thermophiles.
Kingdom Eubacteria: Prokaryotic organisms with a
peptidoglycan cell wall, including cyanobacteria, soil
bacteria, nitrogen-fixing bacteria, and pathogenic
(disease-causing) bacteria.
Kingdom Protista: Eukaryotic, primarily unicellu-
lar (although algae are multicellular), photosynthetic
or heterotrophic organisms, such as amoebas and
paramecia.
Kingdom Fungi: Eukaryotic, mostly multicellular (al-
though yeasts are unicellular), heterotrophic, usually
nonmotile organisms, with cell walls of chitin, such as
mushrooms.
Kingdom Plantae: Eukaryotic, multicellular, non-
motile, usually terrestrial, photosynthetic organisms,
such as trees, grasses, and mosses.
Kingdom Animalia: Eukaryotic, multicellular, motile,
heterotrophic organisms, such as sponges, spiders,
newts, penguins, and humans.
As more is learned about living things, particularly from
the newer evidence that DNA studies provide, scientists will
continue to reevaluate the relationships among the king-
doms of life.
For at least the first 1 billion years of life on earth, all
organisms were bacteria. About 1.5 billion years ago,
the first eukaryotes appeared. Biologists place living
organisms into six general categories called kingdoms.
72 Part I The Origin of Living Things
Microsporidia
Animals
Plants
Ciliates
Slime molds
S. cerevisiae
Euglena
Diplomonads
(Lamblia)
EUKARYA
E. coli
B. subtilus
Thermotoga
Synechocystis sp.
Flavobacteria
Green sulfur
bacteria
Borrelia
burgdorferi
Aquifex
BACTERIA
Methano-
pyrus
Methanococcus
jannaschii
Halobacterium
Halococcus
Archaeoglobus
Methanobacterium
Thermococcus
Sulfolobus
ARCHAEA
(b)
(a)
FIGURE 4.15
The three domains of life. The kingdoms Archaebacteria and Eubacteria are as different from each other as from eukaryotes, so
biologists have assigned them a higher category, a “domain.” (a) A three-domain tree of life based on ribosomal RNA consists of the
Eukarya, Bacteria, and Archaea. (b) New analyses of complete genome sequences contradict the rRNA tree, and suggest other
arrangements such as this one, which splits the Archaea. Apparently genes hopped from branch to branch as early organisms either stole
genes from their food or swapped DNA with their neighbors, even distantly related ones.
Has Life Evolved Elsewhere?
We should not overlook the possibility that life processes
might have evolved in different ways on other planets. A
functional genetic system, capable of accumulating and
replicating changes and thus of adaptation and evolution,
could theoretically evolve from molecules other than car-
bon, hydrogen, nitrogen, and oxygen in a different environ-
ment. Silicon, like carbon, needs four electrons to fill its
outer energy level, and ammonia is even more polar than
water. Perhaps under radically different temperatures and
pressures, these elements might form molecules as diverse
and flexible as those carbon has formed on earth.
The universe has 10
20
(100,000,000,000,000,000,000)
stars similar to our sun. We don’t know how many of
these stars have planets, but it seems increasingly likely
that many do. Since 1996, astronomers have been detect-
ing planets orbiting distant stars. At least 10% of stars are
thought to have planetary systems. If only 1 in 10,000 of
these planets is the right size and at the right distance
from its star to duplicate the conditions in which life orig-
inated on earth, the “life experiment” will have been re-
peated 10
15
times (that is, a million billion times). It does
not seem likely that we are alone.
Ancient Bacteria on Mars?
A dull gray chunk of rock collected in 1984 in Antarctica
ignited an uproar about ancient life on Mars with the report
that the rock contains evidence of possible life. Analysis of
gases trapped within small pockets of the rock indicate it is a
meteorite from Mars. It is, in fact, the oldest rock known to
science—fully 4.5 billion years old. Back then, when this
rock formed on Mars, that cold, arid planet was much
warmer, flowed with water, and had a carbon dioxide
atmosphere—conditions not too different from those that
spawned life on earth.
When examined with powerful electron microscopes,
carbonate patches within the meteorite exhibit what look
like microfossils, some 20 to 100 nanometers in length. One
hundred times smaller than any known bacteria, it is not
clear they actually are fossils, but the resemblance to bacte-
ria is striking.
Viewed as a whole, the evidence of bacterial life associ-
ated with the Mars meteorite is not compelling. Clearly,
more painstaking research remains to be done before the
discovery can claim a scientific consensus. However, while
there is no conclusive evidence of bacterial life associated
with this meteorite, it seems very possible that life has
evolved on other worlds in addition to our own.
Deep-Sea Vents
The possibility that life on earth actually originated in the
vicinity of deep-sea hydrothermal vents is gaining popular-
ity. At the bottom of the ocean, where these vents spewed
out a rich froth of molecules, the geological turbulence and
radioactive energy battering the land was absent, and things
were comparatively calm. The thermophilic archaebacteria
found near these vents today are the most ancient group of
organisms living on earth. Perhaps the gentler environment
of the ocean depths was the actual cradle of life.
Other Planets
Has life evolved on other worlds within our solar system?
There are planets other than ancient Mars with conditions
not unlike those on earth. Europa, a large moon of Jupiter, is
a promising candidate (figure 4.16). Europa is covered with
ice, and photos taken in close orbit in the winter of 1998 re-
veal seas of liquid water beneath a thin skin of ice. Additional
satellite photos taken in 1999 suggest that a few miles under
the ice lies a liquid ocean of water larger than earth’s, warmed
by the push and pull of the gravitational attraction of Jupiter’s
many large satellite moons. The conditions on Europa now
are far less hostile to life than the conditions that existed in
the oceans of the primitive earth. In coming decades satellite
missions are scheduled to explore this ocean for life.
There are so many stars that life may have evolved
many times. Although evidence for life on Mars is not
compelling, the seas of Europa offer a promising
candidate which scientists are eager to investigate.
Chapter 4 The Origin and Early History of Life 73
FIGURE 4.16
Is there life elsewhere? Currently the most likely candidate for
life elsewhere within the solar system is Europa, one of the many
moons of the large planet Jupiter.
Chapter 4
Summary Questions Media Resources
4.1 All living things share key characteristics.
74 Part I The Origin of Living Things
?All living things are characterized by cellular
organization, growth, reproduction, and heredity.
?Other properties commonly exhibited by living
things include movement and sensitivity to stimuli.
1. What characteristics of living
things are necessary
characteristics (possessed by all
living things), and which are
sufficient characteristics
(possessed only by living things)?
?Of the many explanations of how life might have
originated, only the theories of spontaneous and
extraterrestrial origins provide scientifically testable
explanations.
?Experiments recreating the atmosphere of primitive
earth, with the energy sources and temperatures
thought to be prevalent at that time, have led to the
spontaneous formation of amino acids and other
biologically significant molecules.
2. What molecules are thought
to have been present in the
atmosphere of the early earth?
Which molecule that was
notably absent then is now a
major component of the
atmosphere?
4.2 There are many ideas about the origin of life.
? The first cells are thought to have arisen from
aggregations of molecules that were more stable and,
therefore, persisted longer.
? It has been suggested that RNA may have arisen
before cells did, and subsequently became packaged
within a membrane.
? Bacteria were the only life-forms on earth for about 1
billion years. At least three kinds of bacteria were
present in ancient times: methane utilizers, anaerobic
photosynthesizers, and eventually O
2
-forming
photosynthesizers.
3. What evidence supports the
argument that RNA evolved first
on the early earth? What
evidence supports the argument
that proteins evolved first?
4. What are coacervates, and
what characteristics do they have
in common with organisms? Are
they alive? Why or why not?
5. What were the earliest known
organisms like, and when did
they appear? What present-day
organisms do they resemble?
4.3 The first cells had little internal structure.
? The first eukaryotes can be seen in the fossil record
about 1.5 billion years ago. All organisms other than
bacteria are their descendants.
? Biologists group all living organisms into six
“kingdoms,” each profoundly different from the
others.
? The two most ancient kingdoms contain prokaryotes
(bacteria); the other four contain eukaryotes.
?There are approximately 10
20
stars in the universe
similar to our sun. It is almost certain that life has
evolved on planets circling some of them.
6. When did the first eukaryotes
appear? By what mechanism are
they thought to have evolved
from the earlier prokaryotes?
7. What sorts of organisms are
contained in each of the six
kingdoms of life recognized by
biologists?
4.4 The first eukaryotic cells were larger and more complex than bacteria.
? Origin of Life
? Art Quizzes:
-Miller-Urey
Experiment
-Miller-Urey
Experiment Results
? Key Events in Earth’s
History
BIOLOGY
RAVEN
JOHNSON
SIX TH
EDITION
www.mhhe.com/raven6/resources4.mhtml
? Art Quiz:
Current Bubble
Hypothesis
75
How Do the Cells of a Growing
Plant Know in Which Direction
to Elongate?
Sometimes questions that seem simple can be devilishly dif-
ficult to answer. Imagine, for example, that you are holding
a green blade of grass in your hand. The grass blade has
been actively growing, its cells dividing and then stretching
and elongating as the blade lengthens. Did you ever wonder
how the individual cells within the blade of grass know in
what direction to grow?
To answer this deceptively simple question, we will first
need to provide answers to several others. Like Sherlock
Holmes following a trail of clues, we must approach the an-
swer we seek in stages.
Question One. First, we need to ask how a blade of grass
is able to grow at all. Plant cells are very different from ani-
mal cells in one key respect: every plant cell is encased
within a tough cell wall made of cellulose and other tough
building materials. This wall provides structural strength
and protection to the plant cell, just as armor plate does for
a battle tank. But battle tanks can’t stretch into longer
shapes! How is a plant cell able to elongate?
It works like this. A growing cell first performs a little
chemistry to make its wall slightly acidic. The acidity acti-
vates enzymes that attack the cell wall from the inside, rear-
ranging cellulose cross-links until the wall loses its rigidity.
The cell wall is now able to stretch. The cell then sucks in
water, creating pressure. Like blowing up a long balloon,
the now-stretchable cell elongates.
Question Two. In a growing plant organ, like the blade
of grass, each growing cell balloons out lengthwise. Stating
this more formally, a botanist would say the cell elongates
parallel to the axis along which the blade of grass is extend-
ing. This observation leads to the second question we must
answer: How does an individual plant cell control the di-
rection in which it elongates?
It works like this. Before the stretchable cell balloons out,
tiny microfibrils of cellulose are laid down along its inside sur-
face. On a per weight basis, these tiny fibrils have the tensile
strength of steel! Arrays of these cellulose microfibrils are
organized in bands perpendicular to the axis of elongation,
like steel belts. These tough bands reinforce the plant cell wall
laterally, so that when the cell sucks in water, there is only one
way for the cell to expand—lengthwise, along the axis.
Question Three. Now we’re getting somewhere. How
are the newly made cellulose microfibrils laid down so that
they are oriented correctly, perpendicular to the axis of
elongation?
It works like this. The complicated enzymic machine that
makes the cellulose microfibrils is guided by special
guiderails that run like railroad tracks along the interior sur-
face. The enzyme complex travels along these guiderails,
laying down microfibrils as it goes. The guiderails are con-
structed of chainlike protein molecules called microtubules,
assembled into overlapping arrays. Botanists call these ar-
rays of microtubules associated with the interior of the cell
surface “cortical microtubules.”
Question Four. But we have only traded one puzzle for
another. How are the cortical microtubules positioned cor-
rectly, perpendicular to the axis of elongation?
It works like this. In newly made cells, the microtubule
assemblies are already present, but are not organized. They
simply lie about in random disarray. As the cell prepares to
elongate by lessening the rigidity of its cell wall, the micro-
tubule assemblies become organized into the orderly trans-
verse arrays we call cortical microtubules.
Question Five. Finally, we arrive at the question we had
initially set out to answer. How are microtubule assemblies
aligned properly? What sort of signal directs them to ori-
ent perpendicular to the axis of elongation? THAT is the
question we need to answer.
Part
II
Biology of the Cell
Seeing cortical microtubules. Cortical microtubules in
epidermal cells of a fava bean are tagged with a flourescent
protein so that their ordered array can be seen.
Real People Doing Real Science
The Experiment
This question has been addressed experimentally in a sim-
ple and direct way in the laboratory of Richard Cyr at
Pennsylvania State University. Rigid plant cells conduct
mechanical force well from one cell to another, and Carol
Wymer (then a graduate student in the Cyr lab) suspected
some sort of mechanical force is the signal guiding cortical
microtubule alignment
Wymer set out to test this hypothesis using centrifuga-
tion. If cortical microtubules are obtaining their positional
information from an applied force, then their alignment
should be affected by centrifugal force, and should be im-
possible if the integrity of the cell wall (which is supposedly
transmitting the mechanical force) is perturbed with chem-
icals that prevent cell wall formation.
Wymer, along with others in the Cyr lab, started out with
cells that were not elongated. She isolated protoplasts (cells
without walls) from the tobacco plant, Nicotiana tabacum, by
exposing the plant cells to enzymes that break down the cell
wall, creating a spherical plant cell. If allowed to grow in cul-
ture, these protoplasts will eventually re-form their cell walls.
In order to examine the effects of directional force on the
elongation patterns of plant cells, Wymer and co-workers
exposed the tobacco protoplasts to a directional force gener-
ated by a centrifuge. Prior experiments had determined
that centrifugation at the low speeds used in these experi-
ments does not disrupt the integrity or shape of the proto-
plasts. The protoplasts were immobilized for centrifugation
by embedding them in an agar medium supported in a
mold. The embedded protoplasts were spun in a centrifuge
at 450 rpm for 15 minutes. Following centrifugation, the
embedded cells were cultured for 72 hours, allowing for cell
elongation to occur.
Following centrifugation, fluorescently tagged micro-
tubule antibody was applied to the protoplasts, which were
then examined with immunofluorescence microscopy for
microtubule orientation.
To confirm the involvement of microtubules as sensors
of directional force in cell elongation, some protoplasts
were incubated prior to centrifugation with a chemical her-
bicide, APM, which disrupts microtubules.
The Results
The biophysical force of centrifugation had significant ef-
fects on the pattern of elongation in the protoplasts follow-
ing the 72-hour culturing period. The microtubules were
randomly arranged in protoplasts that were not centrifuged
but were more ordered in protoplasts that had been cen-
trifuged. The microtubules in these cells were oriented
parallel to the direction of the force, in a direction approxi-
mately perpendicular to the axis of elongation (graph a
above). These results support the hypothesis that plant cell
growth responds to an external biophysical force.
It is true that plant cells are not usually exposed to the
type of mechanical force generated by centrifugation but
this manipulation demonstrates how a physical force can af-
fect cell growth, assumably by influencing the orientation of
cortical microtubules. These could be small, transient bio-
physical forces acting at the subcellular level.
In preparations exposed to the microtubule disrupting
chemical amiprophos-methyl (APM), directed elongation
was blocked (graph b above). This suggests that reorienta-
tion of microtubules is indeed necessary to direct the elon-
gation axis of the plant
Taken together, these experiments support the hypothesis
that the microtubule reorientation that directs cell elongation
may be oriented by a mechanical force. Just what the natural
force might be is an open question, providing an opportunity
for lots of interesting future experiments that are being pur-
sued in the Cyr lab.
Axis of elongation relative to centrifugal force (degrees)
6
9
Number of cells (percentage)
12
0306090
3
0
Axis of elongation relative to centrifugal force (degrees)
6
9
12
0306090
3
0
Not centrifuged
Centrifuged
Not centrifuged
Centrifuged
(b)(a)
APM treated
Effects of centrifugation on cell elongation. (a) Protoplasts (plant cells without cell walls) that were centrifuged showed preferential
elongation in a direction approximately perpendicular to the direction of the force. (b) Protoplasts exposed to APM, a microtubule dis-
rupting chemical, exhibited random cell elongation with or without centrifugation.
To explore this experiment further,
go to the Virtual Lab at
www.mhhe.com/raven6/vlab2.mhtml
77
5
Cell Structure
Concept Outline
5.1 All organisms are composed of cells.
Cells. A cell is a membrane-bounded unit that contains
DNA and cytoplasm. All organisms are cells or aggregates of
cells, descendants of the first cells.
Cells Are Small. The greater relative surface area of small
cells enables more rapid communication between the cell
interior and the environment.
5.2 Eukaryotic cells are far more complex than
bacterial cells.
Bacteria Are Simple Cells. Bacterial cells are small and
lack membrane-bounded organelles.
Eukaryotic Cells Have Complex Interiors. Eukaryotic
cells are compartmentalized by membranes.
5.3 Take a tour of a eukaryotic cell.
The Nucleus: Information Center for the Cell. The
nucleus of a eukaryotic cell isolates the cell’s DNA.
The Endoplasmic Reticulum: Compartmentalizing the
Cell. An extensive system of membranes subdivides the cell
interior.
The Golgi Apparatus: Delivery System of the Cell. A
system of membrane channels collects, modifies, packages,
and distributes molecules within the cell.
Vesicles: Enzyme Storehouses. Sacs that contain enzymes
digest or modify particles in the cell, while other vesicles
transport substances in and out of cells.
Ribosomes: Sites of Protein Synthesis. An RNA-protein
complex directs the production of proteins.
Organelles That Contain DNA. Some organelles with
very different functions contain their own DNA.
The Cytoskeleton: Interior Framework of the Cell. A
network of protein fibers supports the shape of the cell and
anchors organelles.
Cell Movement. Eukaryotic cell movement utilizes
cytoskeletal elements.
Special Things about Plant Cells. Plant cells have a large
central vacuole and strong, multilayered cell walls.
5.4 Symbiosis played a key role in the origin of some
eukaryotic organelles.
Endosymbiosis. Mitochondria and chloroplasts may have
arisen from prokaryotes engulfed by other prokaryotes.
A
ll organisms are composed of cells. The gossamer
wing of a butterfly is a thin sheet of cells, and so is the
glistening outer layer of your eyes. The hamburger or
tomato you eat is composed of cells, and its contents soon
become part of your cells. Some organisms consist of a sin-
gle cell too small to see with the unaided eye (figure 5.1),
while others, like us, are composed of many cells. Cells are
so much a part of life as we know it that we cannot imagine
an organism that is not cellular in nature. In this chapter
we will take a close look at the internal structure of cells. In
the following chapters, we will focus on cells in action—on
how they communicate with their environment, grow, and
reproduce.
FIGURE. 5.1
The single-celled organism Dileptus. The hairlike projections
that cover its surface are cilia, which it undulates to propel itself
through the water (1000×).
78 Part II Biology of the Cell
5.1 All organisms are composed of cells.
2 X 10
-4
mm2 X 10
-2
mm
2 X 10
1
mm
2 X 10
0
mm
2 X 10
-1
mm
2 X 10
-3
mm
Cells
What is a typical cell like, and what would we find inside it?
The general plan of cellular organization varies in the cells
of different organisms, but despite these modifications, all
cells resemble each other in certain fundamental ways. Be-
fore we begin our detailed examination of cell structure,
let’s first summarize three major features all cells have in
common: a plasma membrane, a nucleoid or nucleus, and
cytoplasm.
The Plasma Membrane Surrounds the Cell
The plasma membrane encloses a cell and separates its
contents from its surroundings. The plasma membrane is a
phospholipid bilayer about 5 to 10 nanometers (5 to 10 bil-
lionths of a meter) thick with proteins embedded in it.
Viewed in cross-section with the electron microscope, such
membranes appear as two dark lines separated by a lighter
area. This distinctive appearance arises from the tail-to-tail
packing of the phospholipid molecules that make up the
membrane (see figure 3.18). The proteins of a membrane
may have large hydrophobic domains, which associate with
and become embedded in the phospholipid bilayer.
The proteins of the plasma membrane are in large
part responsible for a cell’s ability to interact with its en-
vironment. Transport proteins help molecules and ions
move across the plasma membrane, either from the envi-
ronment to the interior of the cell or vice versa. Receptor
proteins induce changes within the cell when they come
in contact with specific molecules in the environment,
such as hormones. Markers identify the cell as a particu-
lar type. This is especially important in multicellular
FIGURE 5.2
The size of cells and their contents. This diagram shows the size of human skin cells, organelles, and molecules. In general, the diameter
of a human skin cell is 20 micrometers (μm) or 2 × 10
-2
mm, of a mitochondrion is 2 μm or 2 × 10
-3
mm, of a ribosome is 20 nanometers
(nm) or 2 × 10
-5
mm, of a protein molecule is 2 nm or 2 × 10
-6
mm, and of an atom is 0.2 nm or 2 × 10
-7
mm.
Chapter 5 Cell Structure 79
2 X 10
-7
mm
2 X 10
-5
mm 2 X 10
-6
mm
organisms, whose cells must be able to recognize each
other as they form tissues.
We’ll examine the structure and function of cell mem-
branes more thoroughly in chapter 6.
The Central Portion of the Cell Contains
the Genetic Material
Every cell contains DNA, the hereditary molecule. In
prokaryotes (bacteria), most of the genetic material lies in
a single circular molecule of DNA. It typically resides near
the center of the cell in an area called the nucleoid, but
this area is not segregated from the rest of the cell’s interior
by membranes. By contrast, the DNA of eukaryotes is
contained in the nucleus, which is surrounded by two
membranes. In both types of organisms, the DNA contains
the genes that code for the proteins synthesized by the cell.
The Cytoplasm Comprises the Rest of the Cell’s
Interior
A semifluid matrix called the cytoplasm fills the interior of
the cell, exclusive of the nucleus (nucleoid in prokaryotes)
lying within it. The cytoplasm contains the chemical wealth
of the cell: the sugars, amino acids, and proteins the cell uses
to carry out its everyday activities. In eukaryotic cells, the
cytoplasm also contains specialized membrane-bounded
compartments called organelles.
The Cell Theory
A general characteristic of cells is their microscopic size.
While there are a few exceptions—the marine alga Acetabu-
laria can be up to 5 centimeters long—a typical eukaryotic
cell is 10 to 100 micrometers (10 to 100 millionths of a
meter) in diameter (figure 5.2); most bacterial cells are only
1 to 10 micrometers in diameter.
Because cells are so small, no one observed them until
microscopes were invented in the mid-seventeenth century.
Robert Hooke first described cells in 1665, when he used a
microscope he had built to examine a thin slice of cork, a
nonliving tissue found in the bark of certain trees. Hooke
observed a honeycomb of tiny, empty (because the cells
were dead) compartments. He called the compartments in
the cork cellulae (Latin, “small rooms”), and the term has
come down to us as cells. The first living cells were observed
a few years later by the Dutch naturalist Antonie van
Leeuwenhoek, who called the tiny organisms that he ob-
served “animalcules,” meaning little animals. For another
century and a half, however, biologists failed to recognize
the importance of cells. In 1838, botanist Matthias Schlei-
den made a careful study of plant tissues and developed the
first statement of the cell theory. He stated that all plants
“are aggregates of fully individualized, independent, sepa-
rate beings, namely the cells themselves.” In 1839,
Theodor Schwann reported that all animal tissues also con-
sist of individual cells.
The cell theory, in its modern form, includes the fol-
lowing three principles:
1. All organisms are composed of one or more cells, and
the life processes of metabolism and heredity occur
within these cells.
2. Cells are the smallest living things, the basic units of
organization of all organisms.
3. Cells arise only by division of a previously existing
cell. Although life likely evolved spontaneously in the
environment of the early earth, biologists have con-
cluded that no additional cells are originating sponta-
neously at present. Rather, life on earth represents a
continuous line of descent from those early cells.
A cell is a membrane-bounded unit that contains the
DNA hereditary machinery and cytoplasm. All
organisms are cells or aggregates of cells.
Cells Are Small
How many cells are big enough to see
with the unaided eye? Other than egg
cells, not many. Most are less than 50
micrometers in diameter, far smaller
than the period at the end of this sen-
tence.
The Resolution Problem
How do we study cells if they are too
small to see? The key is to understand
why we can’t see them. The reason we
can’t see such small objects is the
limited resolution of the human eye.
Resolution is defined as the minimum
distance two points can be apart and
still be distinguished as two separated
points. When two objects are closer
together than about 100 micrometers,
the light reflected from each strikes the
same “detector” cell at the rear of the
eye. Only when the objects are farther
than 100 micrometers apart will the
light from each strike different cells,
allowing your eye to resolve them as
two objects rather than one.
Microscopes
One way to increase resolution is to increase magnification,
so that small objects appear larger. Robert Hooke and
Antonie van Leeuwenhoek were able to see small cells by
magnifying their size, so that the cells appeared larger than
the 100-micrometer limit imposed by the human eye.
Hooke and van Leeuwenhoek accomplished this feat with
microscopes that magnified images of cells by bending
light through a glass lens. The size of the image that falls
on the sheet of detector cells lining the back of your eye
depends on how close the object is to your eye—the closer
the object, the bigger the image. Your eye, however, is
incapable of focusing comfortably on an object closer than
about 25 centimeters, because the eye is limited by the size
and thickness of its lens. Hooke and van Leeuwenhoek
assisted the eye by interposing a glass lens between object
and eye. The glass lens adds additional focusing power.
Because the glass lens makes the object appear closer, the
image on the back of the eye is bigger than it would be
without the lens.
Modern light microscopes use two magnifying lenses (and
a variety of correcting lenses) that act like back-to-back eyes.
The first lens focuses the image of the object on the second
lens, which magnifies it again and focuses it on the back of
the eye. Microscopes that magnify in stages using several
lenses are called compound microscopes. They can resolve
structures that are separated by more than 200 nm. An image
from a compound microscope is shown in figure 5.3a.
Increasing Resolution
Light microscopes, even compound ones, are not powerful
enough to resolve many structures within cells. For exam-
ple, a membrane is only 5 nanometers thick. Why not just
add another magnifying stage to the microscope and so in-
crease its resolving power? Because when two objects are
closer than a few hundred nanometers, the light beams re-
flecting from the two images start to overlap. The only way
two light beams can get closer together and still be resolved
is if their “wavelengths” are shorter.
One way to avoid overlap is by using a beam of electrons
rather than a beam of light. Electrons have a much shorter
wavelength, and a microscope employing electron beams
has 1000 times the resolving power of a light microscope.
Transmission electron microscopes, so called because
the electrons used to visualize the specimens are transmitted
through the material, are capable of resolving objects only
0.2 nanometer apart—just twice the diameter of a hydrogen
atom! Figure 5.3b shows a transmission electron micro-
graph.
A second kind of electron microscope, the scanning
electron microscope, beams the electrons onto the surface
of the specimen from a fine probe that passes rapidly back
and forth. The electrons reflected back from the surface of
the specimen, together with other electrons that the speci-
men itself emits as a result of the bombardment, are ampli-
fied and transmitted to a television screen, where the image
can be viewed and photographed. Scanning electron mi-
croscopy yields striking three-dimensional images and has
improved our understanding of many biological and physi-
cal phenomena (figure 5.3c).
80 Part II Biology of the Cell
(a) (b)
(c)
FIGURE 5.3
Human sperm cells viewed with three
different microscopes. (a) Image of sperm
taken with a light microscope.
(b)Transmission electron micrograph of a
sperm cell. (c) Scanning electron
micrograph of sperm cells.
Why Aren’t Cells Larger?
Most cells are not large for practical reasons. The most im-
portant of these is communication. The different regions of
a cell need to communicate with one another in order for
the cell as a whole to function effectively. Proteins and or-
ganelles are being synthesized, and materials are continually
entering and leaving the cell. All of these processes involve
the diffusion of substances at some point, and the larger a
cell is, the longer it takes for substances to diffuse from the
plasma membrane to the center of the cell. For this reason,
an organism made up of many relatively small cells has an
advantage over one composed of fewer, larger cells.
The advantage of small cell size is readily visualized in
terms of the surface area-to-volume ratio. As a cell’s
size increases, its volume increases much more rapidly
than its surface area. For a spherical cell, the increase in
surface area is equal to the square of the increase in di-
ameter, while the increase in volume is equal to the cube
of the increase in diameter. Thus, if two cells differ by a
factor of 10 cm in diameter, the larger cell will have 10
2
,
or 100 times, the surface area, but 10
3
, or 1000 times,
the volume, of the smaller cell (figure 5.4). A cell’s sur-
face provides its only opportunity for interaction with
the environment, as all substances enter and exit a cell
via the plasma membrane. This membrane plays a key
role in controlling cell function, and because small cells
have more surface area per unit of volume than large
ones, the control is more effective when cells are rela-
tively small.
Although most cells are small, some cells are nonetheless
quite large and have apparently overcome the surface area-
to-volume problem by one or more adaptive mechanisms.
For example, some cells have more than one nucleus, allow-
ing genetic information to be spread around a large cell.
Also, some large cells actively move material around their
cytoplasm so that diffusion is not a limiting factor. Lastly,
some large cells, like your own neurons, are long and skinny
so that any given point in the cytoplasm is close to the
plasma membrane, and thus diffusion between the inside
and outside of the cell can still be rapid.
Multicellular organisms usually consist of many small
cells rather than a few large ones because small cells
function more efficiently. They have a greater relative
surface area, enabling more rapid communication
between the center of the cell and the environment.
Chapter 5 Cell Structure 81
Cell radius
(r)
Surface area
(4H9266r
2
)
Volume
(
4
–
3
H9266r
3
)
1 cm
12.57 cm
2
4.189 cm
3
10 cm
1257 cm
2
4189 cm
3
FIGURE 5.4
Surface area-to-volume ratio. As a cell gets larger, its volume increases at a faster rate than its surface area. If the cell radius increases by
10 times, the surface area increases by 100 times, but the volume increases by 1000 times. A cell’s surface area must be large enough to
meet the needs of its volume.
Bacteria Are Simple Cells
Prokaryotes, the bacteria, are the sim-
plest organisms. Prokaryotic cells are
small, consisting of cytoplasm sur-
rounded by a plasma membrane and en-
cased within a rigid cell wall, with no
distinct interior compartments (figure
5.5). A prokaryotic cell is like a one-
room cabin in which eating, sleeping,
and watching TV all occur in the same
room. Bacteria are very important in
the economy of living organisms. They
harvest light in photosynthesis, break
down dead organisms and recycle their
components, cause disease, and are in-
volved in many important industrial
processes. Bacteria are the subject of
chapter 34.
Strong Cell Walls
Most bacteria are encased by a strong cell wall composed
of peptidoglycan, which consists of a carbohydrate matrix
(polymers of sugars) that is cross-linked by short
polypeptide units. No eukaryotes possess cell walls with
this type of chemical composition. With a few exceptions
like TB and leprosy-causing bacteria, all bacteria may be
classified into two types based on differences in their cell
walls detected by the Gram staining procedure. The
name refers to the Danish microbiologist Hans Christian
Gram, who developed the procedure to detect the pres-
ence of certain disease-causing bacteria. Gram-positive
bacteria have a thick, single-layered cell wall that retains
a violet dye from the Gram stain procedure, causing the
stained cells to appear purple under a microscope. More
complex cell walls have evolved in other groups of bacte-
ria. In them, the wall is multilayered and does not retain
the purple dye after Gram staining; such bacteria exhibit
the background red dye and are characterized as gram-
negative.
The susceptibility of bacteria to antibiotics often depends
on the structure of their cell walls. Penicillin and van-
comycin, for example, interfere with the ability of bacteria
to cross-link the peptide units that hold the carbohydrate
chains of the wall together. Like removing all the nails
from a wooden house, this destroys the integrity of the ma-
trix, which can no longer prevent water from rushing in,
swelling the cell to bursting.
Cell walls protect the cell, maintain its shape, and prevent
excessive uptake of water. Plants, fungi, and most protists
also have cell walls of a different chemical structure, which
we will discuss in later chapters.
Long chains of sugars called polysaccharides cover the
cell walls of many bacteria. They enable a bacterium to ad-
here to teeth, skin, food—practically any surface that will
support their growth. Many disease-causing bacteria secrete
a jellylike protective capsule of polysaccharide around the
cell.
Rotating Flagella
Some bacteria use a flagellum (plural, flagella) to move.
Flagella are long, threadlike structures protruding from the
surface of a cell that are used in locomotion and feeding.
Bacterial flagella are protein fibers that extend out from a
bacterial cell. There may be one or more per cell, or none,
depending on the species. Bacteria can swim at speeds up to
20 cell diameters per second by rotating their flagella like
screws (figure 5.6). A “motor” unique to bacteria that is em-
bedded within their cell walls and membranes powers the
rotation. Only a few eukaryotic cells have structures that
truly rotate.
Simple Interior Organization
If you were to look at an electron micrograph of a bacterial
cell, you would be struck by the cell’s simple organiza-
tion. There are few, if any, internal compartments, and
while they contain simple structures like ribosomes, most
have no membrane-bounded organelles, the kinds so
characteristic of eukaryotic cells. Nor do bacteria have a
true nucleus. The entire cytoplasm of a bacterial cell is
one unit with no internal support structure. Consequently,
82 Part II Biology of the Cell
5.2 Eukaryotic cells are far more complex than bacterial cells.
Flagellum
Cell wall
Plasma
membrane
Capsule
Ribosomes
DNAPili
FIGURE 5.5
Structure of a bacterial cell. Generalized cell organization of a bacterium. Some
bacteria have hairlike growths on the outside of the cell called pili.
the strength of the cell comes primarily from its rigid wall
(see figure 5.5).
The plasma membrane of a bacterial cell carries out some
of the functions organelles perform in eukaryotic cells. When
a bacterial cell divides, for example, the bacterial chromosome,
a simple circle of DNA, replicates before the cell divides. The
two DNA molecules that result from the replication attach to
the plasma membrane at different points, ensuring that each
daughter cell will contain one of the identical units of DNA.
Moreover, some photosynthetic bacteria, such as cyanobacte-
ria and Prochloron (figure 5.7), have an extensively folded
plasma membrane, with the folds extending into the cell’s
interior. These membrane folds contain the bacterial
pigments connected with photosynthesis.
Because a bacterial cell contains no membrane-bounded
organelles, the DNA, enzymes, and other cytoplasmic con-
stituents have access to all parts of the cell. Reactions are not
compartmentalized as they are in eukaryotic cells, and the
whole bacterium operates as a single unit.
Bacteria are small cells that lack interior organization.
They are encased by an exterior wall composed of
carbohydrates cross-linked by short polypeptides, and
some are propelled by rotating flagella.
Chapter 5 Cell Structure 83
Bacterial cell wall
Flagellin
Rotary
motor
(b)
Sheath
(a)
(c)
FIGURE 5.6
Bacteria swim by rotating their flagella. (a) The photograph is of Vibrio cholerae, the microbe that causes the serious disease cholera. The
unsheathed core visible at the top of the photograph is composed of a single crystal of the protein flagellin. (b) In intact flagella, the core is
surrounded by a flexible sheath. Imagine that you are standing inside the Vibrio cell, turning the flagellum like a crank. (c) You would
create a spiral wave that travels down the flagellum, just as if you were turning a wire within a flexible tube. The bacterium creates this
kind of rotary motion when it swims.
FIGURE 5.7
Electron micrograph of a photosynthetic bacterial cell.
Extensive folded photosynthetic membranes are visible in this
Prochloron cell (14,500×). The single, circular DNA molecule is
located in the clear area in the central region of the cell.
Eukaryotic Cells Have
Complex Interiors
Eukaryotic cells (figures 5.8 and 5.9) are
far more complex than prokaryotic cells.
The hallmark of the eukaryotic cell is
compartmentalization. The interiors of
eukaryotic cells contain numerous
organelles, membrane-bounded struc-
tures that close off compartments within
which multiple biochemical processes can
proceed simultaneously and indepen-
dently. Plant cells often have a large
membrane-bounded sac called a central
vacuole, which stores proteins, pigments,
and waste materials. Both plant and ani-
mal cells contain vesicles, smaller sacs
that store and transport a variety of mate-
rials. Inside the nucleus, the DNA is
84 Part II Biology of the Cell
Centriole
Lysosome
Mitochondrion
Ribosomes
Rough
endoplasmic
reticulum
Cytoplasm
Nucleus
Nucleolus
Nuclear
envelope
Smooth
endoplasmic
reticulum
Cytoskeleton
Golgi
apparatus
Plasma
membrane
Microvilli
(a)
Plasma membrane
Nucleolus
Nucleus
Lysosome
Ribosomes
Rough
endoplasmic
reticulum
Golgi apparatus
Mitochondrion
Smooth
endoplasmic
reticulum
(b)
FIGURE 5.8
Structure of an animal cell. (a) A generalized
diagram of an animal cell. (b) Micrograph of a
human white blood cell (40,500H11003) with
drawings detailing organelles.
wound tightly around proteins and packaged into
compact units called chromosomes. All eukaryotic
cells are supported by an internal protein scaffold, the
cytoskeleton. While the cells of animals and some
protists lack cell walls, the cells of fungi, plants, and
many protists have strong cell walls composed of cel-
lulose or chitin fibers embedded in a matrix of other
polysaccharides and proteins. This composition is
very different from the peptidoglycan that makes up
bacterial cell walls. Let’s now examine the structure
and function of the internal components of eukaryotic
cells in more detail.
Eukaryotic cells contain membrane-bounded
organelles that carry out specialized functions.
Chapter 5 Cell Structure 85
Plasma membrane Mitochondrion
Nucleus
Chloroplast
Central vacuole
(b)
Cell wall
Plasmodesma
FIGURE 5.9
Structure of a plant cell. A generalized illustration (a) and
micrograph (b) of a plant cell. Most mature plant cells
contain large central vacuoles which occupy a major portion
of the internal volume of the cell (14,000H11003).
Cell
wall
Plasma
membrane
Central
vacuole
Mitochondrion
Ribosomes
Golgi
apparatus
Nucleus
Nucleolus
Chloroplasts
Nuclear
envelope
Rough
endoplasmic
reticulum
Cytoplasm
Smooth
endoplasmic
reticulum
Plasmo-
desmata
Lysosome
(a)
The Nucleus: Information Center
for the Cell
The largest and most easily seen organelle within a eukary-
otic cell is the nucleus (Latin, for kernel or nut), first
described by the English botanist Robert Brown in 1831.
Nuclei are roughly spherical in shape and, in animal cells,
they are typically located in the central region of the cell
(figure 5.10). In some cells, a network of fine cytoplasmic
filaments seems to cradle the nucleus in this position. The
nucleus is the repository of the genetic information that
directs all of the activities of a living eukaryotic cell. Most
eukaryotic cells possess a single nucleus, although the cells
of fungi and some other groups may have several to many
nuclei. Mammalian erythrocytes (red blood cells) lose their
nuclei when they mature. Many nuclei exhibit a dark-
staining zone called the nucleolus, which is a region where
intensive synthesis of ribosomal RNA is taking place.
86 Part II Biology of the Cell
5.3 Take a tour of a eukaryotic cell.
(c)
Cytoplasm
Pore
Nucleus
FIGURE 5.10
The nucleus. (a) The nucleus is composed of a double membrane, called a nuclear envelope, enclosing a fluid-filled interior containing
the chromosomes. In cross-section, the individual nuclear pores are seen to extend through the two membrane layers of the envelope; the
dark material within the pore is protein, which acts to control access through the pore. (b) A freeze-fracture scanning electron micrograph
of a cell nucleus showing nuclear pores (9500×). (c) A transmission electron micrograph (see figure 6.6) of the nuclear membrane showing a
nuclear pore.
Nuclear
pores
Nuclear
pore
Nuclear
envelope
Nucleoplasm Outer membrane
Inner membrane
Nucleolus
(a)
Pore
(b)
The Nuclear Envelope: Getting In and Out
The surface of the nucleus is bounded by two phospho-
lipid bilayer membranes, which together make up the
nuclear envelope (see figure 5.10). The outer mem-
brane of the nuclear envelope is continuous with the
cytoplasm’s interior membrane system, called the endo-
plasmic reticulum. Scattered over the surface of the
nuclear envelope, like craters on the moon, are shallow
depressions called nuclear pores. These pores form 50
to 80 nanometers apart at locations where the two mem-
brane layers of the nuclear envelope pinch together.
Rather than being empty, nuclear pores are filled with
proteins that act as molecular channels, permitting certain
molecules to pass into and out of the nucleus. Passage is
restricted primarily to two kinds of molecules: (1) proteins
moving into the nucleus to be incorporated into nuclear
structures or to catalyze nuclear activities; and (2) RNA
and protein-RNA complexes formed in the nucleus and
exported to the cytoplasm.
The Chromosomes: Packaging the DNA
In both bacteria and eukaryotes, DNA contains the hered-
itary information specifying cell structure and function.
However, unlike the circular DNA of bacteria, the DNA
of eukaryotes is divided into several linear chromosomes.
Except when a cell is dividing, its chromosomes are fully
extended into threadlike strands, called chromatin, of
DNA complexed with protein. This open arrangement
allows proteins to attach to specific nucleotide sequences
along the DNA. Without this access, DNA could not
direct the day-to-day activities of the cell. The chromo-
somes are associated with packaging proteins called
histones. When a cell prepares to divide, the DNA coils
up around the histones into a highly condensed form. In
the initial stages of this condensation, units of histone
can be seen with DNA wrapped around like a sash.
Called nucleosomes, these initial aggregations resemble
beads on a string (figure 5.11). Coiling continues until
the DNA is in a compact mass. Under a light micro-
scope, these fully condensed chromosomes are readily
seen in dividing cells as densely staining rods (figure
5.12). After cell division, eukaryotic chromosomes uncoil
and can no longer be individually distinguished with a
light microscope. Uncoiling the chromosomes into a
more extended form permits enzymes to makes RNA
copies of DNA. Only by means of these RNA copies can
the information in the DNA be used to direct the
synthesis of proteins.
The nucleus of a eukaryotic cell contains the cell’s
hereditary apparatus and isolates it from the rest of
the cell. A distinctive feature of eukaryotes is the
organization of their DNA into complex
chromosomes.
Chapter 5 Cell Structure 87
Central histone
Spacer histone
Nucleosome
Chromosome
DNA
FIGURE 5.11
Nucleosomes. Each nucleosome is a region in which the DNA is
wrapped tightly around a cluster of histone proteins.
FIGURE 5.12
Eukaryotic chromosomes. These condensed chromosomes
within an onion root tip are visible under the light microscope
(500×).
The Endoplasmic Reticulum:
Compartmentalizing the Cell
The interior of a eukaryotic cell is packed with membranes
(table 5.1). So thin that they are invisible under the low re-
solving power of light microscopes, this endomembrane
system fills the cell, dividing it into compartments, channel-
ing the passage of molecules through the interior of the cell,
and providing surfaces for the synthesis of lipids and some
proteins. The presence of these membranes in eukaryotic
cells constitutes one of the most fundamental distinctions
between eukaryotes and prokaryotes.
The largest of the internal membranes is called the
endoplasmic reticulum (ER). The term endoplasmic means
“within the cytoplasm,” and the term reticulum is Latin for
“a little net.” Like the plasma membrane, the ER is
composed of a lipid bilayer embedded with proteins. It
weaves in sheets through the interior of the cell, creating a
series of channels between its folds (figure 5.13). Of the
many compartments in eukaryotic cells, the two largest are
the inner region of the ER, called the cisternal space, and
the region exterior to it, the cytosol.
Rough ER: Manufacturing Proteins for Export
The ER surface regions that are devoted to protein synthe-
sis are heavily studded with ribosomes, large molecular
aggregates of protein and ribonucleic acid (RNA) that trans-
late RNA copies of genes into protein (we will examine
ribosomes in detail later in this chapter). Through the elec-
tron microscope, these ribosome-rich regions of the ER
appear pebbly, like the surface of sandpaper, and they are
therefore called rough ER (see figure 5.13).
The proteins synthesized on the surface of the rough ER
are destined to be exported from the cell. Proteins to be ex-
ported contain special amino acid sequences called signal
sequences. As a new protein is made by a free ribosome
(one not attached to a membrane), the signal sequence of
the growing polypeptide attaches to a recognition factor
that carries the ribosome and its partially completed protein
to a “docking site” on the surface of the ER. As the protein
is assembled it passes through the ER membrane into the
interior ER compartment, the cisternal space, from which it
is transported by vesicles to the Golgi apparatus (figure
5.14). The protein then travels within vesicles to the inner
surface of the plasma membrane, where it is released to the
outside of the cell.
88 Part II Biology of the Cell
Table 5.1 Eukaryotic Cell Structures and Their Functions
Structure Description Function
Cell wall
Cytoskeleton
Flagella (cilia)
Plasma membrane
Endoplasmic reticulum
Nucleus
Golgi apparatus
Lysosomes
Microbodies
Mitochondria
Chloroplasts
Chromosomes
Nucleolus
Ribosomes
Outer layer of cellulose or chitin; or absent
Network of protein filaments
Cellular extensions with 9 + 2 arrangement of pairs of
microtubules
Lipid bilayer with embedded proteins
Network of internal membranes
Structure (usually spherical) surrounded by double
membrane that contains chromosomes
Stacks of flattened vesicles
Vesicles derived from Golgi apparatus that contain
hydrolytic digestive enzymes
Vesicles formed from incorporation of lipids and
proteins containing oxidative and other enzymes
Bacteria-like elements with double membrane
Bacteria-like elements with membranes containing
chlorophyll, a photosynthetic pigment
Long threads of DNA that form a complex with
protein
Site of genes for rRNA synthesis
Small, complex assemblies of protein and RNA, often
bound to endoplasmic reticulum
Protection; support
Structural support; cell movement
Motility or moving fluids over surfaces
Regulates what passes into and out of cell;
cell-to-cell recognition
Forms compartments and vesicles;
participates in protein and lipid synthesis
Control center of cell; directs protein
synthesis and cell reproduction
Packages proteins for export from cell;
forms secretory vesicles
Digest worn-out organelles and cell debris;
play role in cell death
Isolate particular chemical activities from
rest of cell
“Power plants” of the cell; sites of oxidative
metabolism
Sites of photosynthesis
Contain hereditary information
Assembles ribosomes
Sites of protein synthesis
Smooth ER: Organizing Internal Activities
Regions of the ER with relatively few bound ribosomes are
referred to as smooth ER. The membranes of the smooth
ER contain many embedded enzymes, most of them active
only when associated with a membrane. Enzymes anchored
within the ER, for example, catalyze the synthesis of a vari-
ety of carbohydrates and lipids. In cells that carry out exten-
sive lipid synthesis, such as those in the testes, intestine, and
brain, smooth ER is particularly abundant. In the liver, the
enzymes of the smooth ER are involved in the detoxification
of drugs including amphetamines, morphine, codeine, and
phenobarbital.
Some vesicles form at the plasma membrane by budding
inward, a process called endocytosis. Some then move into
the cytoplasm and fuse with the smooth endoplasmic reticu-
lum. Others form secondary lysosomes or other interior
vesicles.
The endoplasmic reticulum (ER) is an extensive system
of folded membranes that divides the interior of
eukaryotic cells into compartments and channels.
Rough ER synthesizes proteins, while smooth ER
organizes the synthesis of lipids and other biosynthetic
activities.
Chapter 5 Cell Structure 89
0.08 μm
Ribosomes
Rough
endoplasmic
reticulum
Smooth
endoplasmic
reticulum
FIGURE 5.13
The endoplasmic reticulum. Ribosomes are associated with only one side of the rough ER; the other side is the boundary of a separate
compartment within the cell into which the ribosomes extrude newly made proteins destined for secretion. Smooth endoplasmic reticulum
has few to no bound ribosomes.
Cytoplasm
Lumen
mRNA
Ribosome
Membrane of
endoplasmic reticulum
Polypeptide
Signal
sequence
FIGURE 5.14
Signal sequences direct proteins to their destinations in the
cell. In this example, a sequence of hydrophobic amino acids (the
signal sequence) on a secretory protein attaches them (and the
ribosomes making them) to the membrane of the ER. As the
protein is synthesized, it passes into the lumen (internal chamber)
of the ER. The signal sequence is clipped off after the leading
edge of the protein enters the lumen.
The Golgi Apparatus: Delivery
System of the Cell
At various locations within the endomembrane system,
flattened stacks of membranes called Golgi bodies occur,
often interconnected with one another. These structures
are named for Camillo Golgi, the nineteenth-century
Italian physician who first called attention to them. The
numbers of Golgi bodies a cell contains ranges from 1 or
a few in protists, to 20 or more in animal cells and several
hundred in plant cells. They are especially abundant in
glandular cells, which manufacture and secrete sub-
stances. Collectively the Golgi bodies are referred to as
the Golgi apparatus (figure 5.15).
The Golgi apparatus functions in the collection, packag-
ing, and distribution of molecules synthesized at one place
in the cell and utilized at another location in the cell. A
Golgi body has a front and a back, with distinctly different
membrane compositions at the opposite ends. The front, or
receiving end, is called the cis face, and is usually located
near ER. Materials move to the cis face in transport vesicles
that bud off of the ER. These vesicles fuse with the cis face,
emptying their contents into the interior, or lumen, of the
Golgi apparatus. These ER-synthesized molecules then pass
through the channels of the Golgi apparatus until they
reach the back, or discharging end, called the trans face,
where they are discharged in secretory vesicles (figure 5.16).
Proteins and lipids manufactured on the rough and
smooth ER membranes are transported into the Golgi ap-
paratus and modified as they pass through it. The most
common alteration is the addition or modification of short
sugar chains, forming a glycoprotein when sugars are com-
plexed to a protein and a glycolipid when sugars are bound
to a lipid. In many instances, enzymes in the Golgi appara-
tus modify existing glycoproteins and glycolipids made in
the ER by cleaving a sugar from their sugar chain or modi-
fying one or more of the sugars.
The newly formed or altered glycoproteins and glycol-
ipids collect at the ends of the Golgi bodies, in flattened
stacked membrane folds called cisternae (Latin,
“collecting vessels”). Periodically, the membranes of the
cisternae push together, pinching off small, membrane-
bounded secretory vesicles containing the glycoprotein
and glycolipid molecules. These vesicles then move to
other locations in the cell, distributing the newly
synthesized molecules to their appropriate destinations.
Liposomes are synthetically manufactured vesicles that
contain any variety of desirable substances (such as
drugs), and can be injected into the body. Because the
membrane of liposomes is similar to plasma and organellar
membranes, these liposomes serve as an effective and
natural delivery system to cells and may prove to be of
great therapeutic value.
The Golgi apparatus is the delivery system of the
eukaryotic cell. It collects, packages, modifies, and
distributes molecules that are synthesized at one
location within the cell and used at another.
90 Part II Biology of the Cell
Secretory vesicles
Vesicle
0.57 μm
FIGURE 5.15
The Golgi apparatus. The Golgi apparatus is a smooth, concave membranous structure located near the middle of the cell. It receives
material for processing on one surface and sends the material packaged in vesicles off the other. The substance in a vesicle could be for
export out of the cell or for distribution to another region within the same cell.
Chapter 5 Cell Structure 91
Budding
vesicle
Fusion
of vesicle
with Golgi
apparatus
Migrating
transport
vesicle
Protein
Proteins
Transport
vesicle
Golgi
apparatus
Secretory
vesicle
Smooth
endoplasmic
reticulum
Rough
endoplasmic
reticulum
Nuclear pore
Nucleus
Cisternae
Ribosome
Trans face
Cis face
Cell membrane
Protein expelled
Cytoplasm
Extracellular fluid
FIGURE 5.16
How proteins are transported within
the cell. Proteins are manufactured at the
ribosome and then released into the internal
compartments of the rough ER. If the
newly synthesized proteins are to be used at
a distant location in or outside of the cell,
they are transported within vesicles that bud
off the rough ER and travel to the cis face,
or receiving end, of the Golgi apparatus.
There they are modified and packaged into
secretory vesicles. The secretory vesicles
then migrate from the trans face, or
discharging end, of the Golgi apparatus to
other locations in the cell, or they fuse with
the cell membrane, releasing their contents
to the external cellular environment.
Vesicles: Enzyme
Storehouses
Lysosomes: Intracellular
Digestion Centers
Lysosomes, membrane-bounded diges-
tive vesicles, are also components of the
endomembrane system that arise from
the Golgi apparatus. They contain high
levels of degrading enzymes, which cat-
alyze the rapid breakdown of proteins,
nucleic acids, lipids, and carbohydrates.
Throughout the lives of eukaryotic cells,
lysosomal enzymes break down old or-
ganelles, recycling their component mol-
ecules and making room for newly
formed organelles. For example, mito-
chondria are replaced in some tissues
every 10 days.
The digestive enzymes in lysosomes
function best in an acidic environment.
Lysosomes actively engaged in digestion
keep their battery of hydrolytic enzymes
(enzymes that catalyze the hydrolysis of
molecules) fully active by pumping protons into their inte-
riors and thereby maintaining a low internal pH. Lyso-
somes that are not functioning actively do not maintain an
acidic internal pH and are called primary lysosomes. When a
primary lysosome fuses with a food vesicle or other or-
ganelle, its pH falls and its arsenal of hydrolytic enzymes is
activated; it is then called a secondary lysosome.
In addition to breaking down organelles and other struc-
tures within cells, lysosomes also eliminate other cells that
the cell has engulfed in a process called phagocytosis, a spe-
cific type of endocytosis (see chapter 6). When a white
blood cell, for example, phagocytizes a passing pathogen,
lysosomes fuse with the resulting “food vesicle,” releasing
their enzymes into the vesicle and degrading the material
within (figure 5.17).
Microbodies
Eukaryotic cells contain a variety of enzyme-bearing,
membrane-enclosed vesicles called microbodies. Micro-
bodies are found in the cells of plants, animals, fungi, and
protists. The distribution of enzymes into microbodies is
one of the principal ways in which eukaryotic cells orga-
nize their metabolism.
While lysosomes bud from the endomembrane system,
microbodies grow by incorporating lipids and protein,
then dividing. Plant cells have a special type of microbody
called a glyoxysome that contains enzymes that convert
fats into carbohydrates. Another type of microbody, a
peroxisome, contains enzymes that catalyze the removal
of electrons and associated hydrogen atoms (figure 5.18).
If these oxidative enzymes were not isolated within micro-
bodies, they would tend to short-circuit the metabolism of
the cytoplasm, which often involves adding hydrogen atoms
to oxygen. The name peroxisome refers to the hydrogen
peroxide produced as a by-product of the activities of the
oxidative enzymes in the microbody. Hydrogen peroxide is
dangerous to cells because of its violent chemical reactivity.
However, peroxisomes also contain the enzyme catalase,
which breaks down hydrogen peroxide into harmless
water and oxygen.
Lysosomes and peroxisomes are vesicles that contain
digestive and detoxifying enzymes. The isolation of
these enzymes in vesicles protects the rest of the cell
from inappropriate digestive activity.
92 Part II Biology of the Cell
Cytoplasm
Phagocytosis
Food
vesicle
Golgi
apparatus
Lysosomes
Plasma
membrane
Digestion of
phagocytized
food particles
or cells
Endoplasmic
reticulum
Transport
vesicle
Old or damaged
organelle
Breakdown
of old
organelle
Extracellular
fluid
FIGURE 5.17
Lysosomes. Lysosomes contain hydrolytic enzymes that digest particles or cells taken
into the cell by phagocytosis and break down old organelles.
0.21 μm
FIGURE 5.18
A peroxisome.
Peroxisomes are
spherical organelles
that may contain a
large diamond-shaped
crystal composed of
protein. Peroxisomes
contain digestive and
detoxifying enzymes
that produce hydrogen
peroxide as a by-
product.
Ribosomes: Sites of Protein
Synthesis
Although the DNA in a cell’s nucleus encodes the amino
acid sequence of each protein in the cell, the proteins are
not assembled there. A simple experiment demonstrates
this: if a brief pulse of radioactive amino acid is administered
to a cell, the radioactivity shows up associated with newly
made protein, not in the nucleus, but in the cytoplasm.
When investigators first carried out these experiments, they
found that protein synthesis was associated with large RNA-
protein complexes they called ribosomes.
Ribosomes are made up of several molecules of a special
form of RNA called ribosomal RNA, or rRNA, bound within
a complex of several dozen different proteins. Ribosomes are
among the most complex molecular assemblies found in cells.
Each ribosome is composed of two subunits (figure 5.19).
The subunits join to form a functional ribosome only when
they attach to another kind of RNA, called messenger RNA
(mRNA) in the cytoplasm. To make proteins, the ribosome
attaches to the mRNA, which is a transcribed copy of a
portion of DNA, and uses the information to direct the
synthesis of a protein.
Bacterial ribosomes are smaller than eukaryotic ribo-
somes. Also, a bacterial cell typically has only a few thou-
sand ribosomes, while a metabolically active eukaryotic cell,
such as a human liver cell, contains several million. Proteins
that function in the cytoplasm are made by free ribosomes
suspended there, while proteins bound within membranes
or destined for export from the cell are assembled by ribo-
somes bound to rough ER.
The Nucleolus Manufactures Ribosomal
Subunits
When cells are synthesizing a large number of proteins,
they must first make a large number of ribosomes. To facili-
tate this, many hundreds of copies of the portion of the
DNA encoding the rRNA are clustered together on the
chromosome. By transcribing RNA molecules from this
cluster, the cell rapidly generates large numbers of the mol-
ecules needed to produce ribosomes.
At any given moment, many rRNA molecules dangle
from the chromosome at the sites of these clusters of genes
that encode rRNA. Proteins associate with the dangling
rRNA molecules. These areas where ribosomes are being
assembled are easily visible within the nucleus as one or
more dark-staining regions, called nucleoli (singular, nucle-
olus; figure 5.20). Nucleoli can be seen under the light mi-
croscope even when the chromosomes are extended, unlike
the rest of the chromosomes, which are visible only when
condensed.
Ribosomes are the sites of protein synthesis in the
cytoplasm.
Chapter 5 Cell Structure 93
Small
subunit
Large
subunit
Ribosome
FIGURE 5.19
A ribosome. Ribosomes consist of a large and a small subunit
composed of rRNA and protein. The individual subunits are
synthesized in the nucleolus and then move through the nuclear
pores to the cytoplasm, where they assemble. Ribosomes serve as
sites of protein synthesis.
FIGURE 5.20
The nucleolus. This is the interior of a rat liver cell, magnified
about 6000 times. A single large nucleus occupies the center of
the micrograph. The electron-dense area in the lower left of the
nucleus is the nucleolus, the area where the major components
of the ribosomes are produced. Partly formed ribosomes can be
seen around the nucleolus.
Organelles That
Contain DNA
Among the most interesting cell
organelles are those in addition
to the nucleus that contain
DNA.
Mitochondria: The Cell’s
Chemical Furnaces
Mitochondria (singular, mito-
chondrion) are typically tubular
or sausage-shaped organelles
about the size of bacteria and
found in all types of eukaryotic
cells (figure 5.21). Mitochondria are bounded by two
membranes: a smooth outer membrane and an inner one
folded into numerous contiguous layers called cristae
(singular, crista). The cristae partition the mitochondrion
into two compartments: a matrix, lying inside the inner
membrane; and an outer compartment, or intermem-
brane space, lying between the two mitochondrial mem-
branes. On the surface of the inner membrane, and also
embedded within it, are proteins that carry out oxidative
metabolism, the oxygen-requiring process by which en-
ergy in macromolecules is stored in ATP.
Mitochondria have their own DNA; this DNA contains
several genes that produce proteins essential to the mito-
chondrion’s role in oxidative metabolism. All of these genes
are copied into RNA and used to make proteins within the
mitochondrion. In this process, the mitochondria employ
small RNA molecules and ribosomal components that the
mitochondrial DNA also encodes. However, most of the
genes that produce the enzymes used in oxidative metabo-
lism are located in the nucleus.
A eukaryotic cell does not produce brand new mito-
chondria each time the cell divides. Instead, the mito-
chondria themselves divide in two, doubling in number,
and these are partitioned between the new cells. Most of
the components required for mitochondrial division are
encoded by genes in the nucleus and translated into pro-
teins by cytoplasmic ribosomes. Mitochondrial replica-
tion is, therefore, impossible without nuclear participa-
tion, and mitochondria thus cannot be grown in a
cell-free culture.
Chloroplasts: Where Photosynthesis Takes Place
Plants and other eukaryotic organisms that carry out
photosynthesis typically contain from one to several
hundred chloroplasts. Chloroplasts bestow an obvious
advantage on the organisms that possess them: they can
manufacture their own food. Chloroplasts contain the
photosynthetic pigment chlorophyll that gives most
plants their green color.
The chloroplast body is enclosed, like the mitochon-
drion, within two membranes that resemble those of mito-
chondria (figure 5.22). However, chloroplasts are larger
and more complex than mitochondria. In addition to the
outer and inner membranes, which lie in close association
with each other, chloroplasts have a closed compartment of
stacked membranes called grana (singular, granum), which
lie internal to the inner membrane. A chloroplast may con-
tain a hundred or more grana, and each granum may con-
tain from a few to several dozen disk-shaped structures
called thylakoids. On the surface of the thylakoids are the
light-capturing photosynthetic pigments, to be discussed in
depth in chapter 10. Surrounding the thylakoid is a fluid
matrix called the stroma.
Like mitochondria, chloroplasts contain DNA, but
many of the genes that specify chloroplast components are
also located in the nucleus. Some of the elements used in
the photosynthetic process, including the specific protein
components necessary to accomplish the reaction, are syn-
thesized entirely within the chloroplast.
94 Part II Biology of the Cell
Intermembrane
space
Inner membrane
Outer membrane
Matrix
Crista
membrane
e
(a)
(b)
FIGURE 5.21
Mitochondria. (a) The inner membrane of a mitochondrion is
shaped into folds called cristae, which greatly increase the surface
area for oxidative metabolism. (b) Mitochondria in cross-section
and cut lengthwise (70,000×).
Other DNA-containing organelles in plants are called
leucoplasts, which lack pigment and a complex internal
structure. In root cells and some other plant cells, leu-
coplasts may serve as starch storage sites. A leucoplast that
stores starch (amylose) is sometimes termed an amyloplast.
These organelles—chloroplasts, leucoplasts, and amylo-
plasts—are collectively called plastids. All plastids come
from the division of existing plastids.
Centrioles: Microtubule Assembly Centers
Centrioles are barrel-shaped organelles found in the cells
of animals and most protists. They occur in pairs, usually
located at right angles to each other near the nuclear
membranes (figure 5.23); the region surrounding the pair
in almost all animal cells is referred to as a centrosome.
Although the matter is in some dispute, at least some
centrioles seem to contain DNA, which apparently is in-
volved in producing their structural proteins. Centrioles
help to assemble microtubules, long, hollow cylinders of
the protein tubulin. Microtubules influence cell shape,
move the chromosomes in cell division, and provide the
functional internal structure of flagella and cilia, as we
will discuss later. Centrioles may be contained in areas
called microtubule-organizing centers (MTOCs). The
cells of plants and fungi lack centrioles, and cell biolo-
gists are still in the process of characterizing their
MTOCs.
Both mitochondria and chloroplasts contain specific
genes related to some of their functions, but both
depend on nuclear genes for other functions. Some
centrioles also contain DNA, which apparently helps
control the synthesis of their structural proteins.
Chapter 5 Cell Structure 95
Outer membrane
Inner membrane
Granum
Thylakoid
Stroma
FIGURE 5.22
Chloroplast structure. The inner membrane of a chloroplast is fused to form stacks of closed vesicles called thylakoids. Within these
thylakoids, photosynthesis takes place. Thylakoids are typically stacked one on top of the other in columns called grana.
0.09 μm
Microtubule triplet
FIGURE 5.23
Centrioles. (a) This electron micrograph shows a pair of
centrioles (75,000×). The round shape is a centriole in cross-
section; the rectangular shape is a centriole in longitudinal section.
(b) Each centriole is composed of nine triplets of microtubules.
(a)
(b)
The Cytoskeleton: Interior
Framework of the Cell
The cytoplasm of all eukaryotic cells is crisscrossed by a
network of protein fibers that supports the shape of the
cell and anchors organelles to fixed locations. This net-
work, called the cytoskeleton (figure 5.24), is a dynamic
system, constantly forming and disassembling. Individual
fibers form by polymerization, as identical protein sub-
units attract one another chemically and spontaneously
assemble into long chains. Fibers disassemble in the same
way, as one subunit after another breaks away from one
end of the chain.
Eukaryotic cells may contain three types of cytoskeletal
fibers, each formed from a different kind of subunit:
1. Actin filaments. Actin filaments are long fibers
about 7 nanometers in diameter. Each filament is
composed of two protein chains loosely twined to-
gether like two strands of pearls (figure 5.25a). Each
“pearl,” or subunit, on the chains is the globular pro-
tein actin. Actin molecules spontaneously form these
filaments, even in a test tube; a cell regulates the rate of
their formation through other proteins that act as
switches, turning on polymerization when appropriate.
Actin filaments are responsible for cellular movements
such as contraction, crawling, “pinching” during divi-
sion, and formation of cellular extensions.
2. Microtubules. Microtubules are hollow tubes
about 25 nanometers in diameter, each composed of
a ring of 13 protein protofilaments (figure 5.25b).
Globular proteins consisting of dimers of alpha and
beta tubulin subunits polymerize to form the 13
protofilaments. The protofilaments are arrayed side
by side around a central core, giving the microtubule
its characteristic tube shape. In many cells, micro-
tubules form from MTOC nucleation centers near
the center of the cell and radiate toward the periph-
ery. They are in a constant state of flux, continually
polymerizing and depolymerizing (the average half-
life of a microtubule ranges from 10 minutes in a
nondividing animal cell to as short as 20 seconds in a
dividing animal cell), unless stabilized by the binding
of guanosine triphosphate (GTP) to the ends, which
inhibits depolymerization. The ends of the micro-
tubule are designated as “+” (away from the nucle-
ation center) or “?” (toward the nucleation center).
Along with allowing for cellular movement, micro-
tubules are responsible for moving materials within
the cell itself. Special motor proteins, discussed later
in this chapter, move cellular organelles around the
cell on microtubular “tracks.” Kinesin proteins move
organelles toward the “+” end (toward the cell pe-
riphery), and dyneins move them toward the “?” end.
96 Part II Biology of the Cell
Nuclear
envelope
Nucleolus
Ribosomes
Mitochondrion
Rough
endoplasmic
reticulum
Smooth
endoplasmic
reticulum
Cytoskeleton
FIGURE 5.24
The cytoskeleton. In this diagrammatic cross-section of a
eukaryotic cell, the cytoskeleton, a network of fibers, supports
organelles such as mitochondria.
3. Intermediate filaments. The most durable ele-
ment of the cytoskeleton in animal cells is a system
of tough, fibrous protein molecules twined together
in an overlapping arrangement (figure 5.25c). These
fibers are characteristically 8 to 10 nanometers in di-
ameter, intermediate in size between actin filaments
and microtubules (which is why they are called in-
termediate filaments). Once formed, intermediate
filaments are stable and usually do not break down.
Intermediate filaments constitute a heterogeneous
group of cytoskeletal fibers. The most common
type, composed of protein subunits called vimentin,
provides structural stability for many kinds of cells.
Keratin, another class of intermediate filament, is
found in epithelial cells (cells that line organs and
body cavities) and associated structures such as hair
and fingernails. The intermediate filaments of nerve
cells are called neurofilaments.
As we will discuss in the next section, the cytoskeleton
provides an interior framework that supports the shape of
the cell, stretching the plasma membrane much as the poles
of a circus tent. Changing the relative length of cytoskeleton
filaments allows cells to rapidly alter their shape, extending
projections out or folding inward. Within the cell, the
framework of filaments provides a molecular highway along
which molecules can be transported.
Elements of the cytoskeleton crisscross the cytoplasm,
supporting the cell shape and anchoring organelles in
place. There are three principal types of fibers: actin
filaments, microtubules, and intermediate filaments.
Chapter 5 Cell Structure 97
(b) Microtubule(a) Actin filament
(c) Intermediate filament
Mitochondrion
Microtubule
Intermediate filament
Ribosome
Rough endoplasmic
reticulum
Actin filament
Cell membrane
FIGURE 5.25
Molecules that make up the cytoskeleton. (a) Actin filaments. Actin filaments are made of two strands of the fibrous protein actin
twisted together and usually occur in bundles. Actin filaments are ubiquitous, although they are concentrated below the plasma membrane
in bundles known as stress fibers, which may have a contractile function. (b) Microtubules. Microtubules are composed of 13 stacks of
tubulin protein subunits arranged side by side to form a tube. Microtubules are comparatively stiff cytoskeletal elements that serve to
organize metabolism and intracellular transport in the nondividing cell. (c) Intermediate filaments. Intermediate filaments are composed of
overlapping staggered tetramers of protein. This molecular arrangement allows for a ropelike structure that imparts tremendous
mechanical strength to the cell.
Cell Movement
Essentially all cell motion is tied to the movement of actin
filaments, microtubules, or both. Intermediate filaments act
as intracellular tendons, preventing excessive stretching of
cells, and actin filaments play a major role in determining
the shape of cells. Because actin filaments can form and dis-
solve so readily, they enable some cells to change shape
quickly. If you look at the surfaces of such cells under a mi-
croscope, you will find them alive with motion, as projec-
tions, called microvilli in animal cells, shoot outward from
the surface and then retract, only to shoot out elsewhere
moments later (figure 5.26).
Some Cells Crawl
It is the arrangement of actin filaments within the cell cy-
toplasm that allows cells to “crawl,” literally! Crawling is a
significant cellular phenomenon, essential to inflamma-
tion, clotting, wound healing, and the spread of cancer.
White blood cells in particular exhibit this ability. Pro-
duced in the bone marrow, these cells are released into
the circulatory system and then eventually crawl out of
capillaries and into the tissues to destroy potential
pathogens.
Cells exist in a gel-sol state; that is, at any given time,
some regions of the cell are rigid (gel) and some are more
fluid (sol). The cell is typically more sol-like in its interior,
and more gel-like at its perimeter. To crawl, the cell cre-
ates a weak area in the gel perimeter, and then forces the
fluid (sol) interior through the weak area, forming a
pseudopod (“false foot”). As a result a large section of cy-
toplasm oozes off in a different direction, but still remains
within the plasma membrane. Once extended, the pseudo-
pod stabilizes into a gel state, assembling actin filaments.
Specific membrane proteins in the pseudopod stick to the
surface the cell is crawling on, and the rest of the cell is
dragged in that direction. The pressure required to force
out a developing pseudopod is created when actin filaments
in the trailing end of the cell contract, just as squeezing a
water balloon at one end forces the balloon to bulge out at
the other end.
Moving Material within the Cell
Actin filaments and microtubules often orchestrate their ac-
tivities to affect cellular processes. For example, during cell
reproduction (see chapter 11), newly replicated chromo-
somes move to opposite sides of a dividing cell because they
are attached to shortening microtubules. Then, in animal
cells, a belt of actin pinches the cell in two by contracting
like a purse string. Muscle cells also use actin filaments to
contract their cytoskeletons. The fluttering of an eyelash, the
flight of an eagle, and the awkward crawling of a baby all de-
pend on these cytoskeletal movements within muscle cells.
Not only is the cytoskeleton responsible for the cell’s
shape and movement, but it also provides a scaffold that
holds certain enzymes and other macromolecules in defined
areas of the cytoplasm. Many of the enzymes involved in
cell metabolism, for example, bind to actin filaments; so do
ribosomes. By moving and anchoring particular enzymes
near one another, the cytoskeleton, like the endoplasmic
reticulum, organizes the cell’s activities.
Intracellular Molecular Motors
Certain eukaryotic cells must move materials from one place
to another in the cytoplasm. Most cells use the endomem-
brane system as an intracellular highway; the Golgi appara-
tus packages materials into vesicles that move through the
channels of the endoplasmic reticulum to the far reaches of
the cell. However, this highway is only effective over short
distances. When a cell has to transport materials through
long extensions like the axon of a nerve cell, the ER high-
ways are too slow. For these situations, eukaryotic cells have
developed high-speed locomotives that run along micro-
tubular tracks.
Four components are required: (1) a vesicle or or-
ganelle that is to be transported, (2) a motor molecule
that provides the energy-driven motion, (3) a connector
molecule that connects the vesicle to the motor mole-
cule, and (4) microtubules on which the vesicle will ride
like a train on a rail. For example, embedded within the
membranes of endoplasmic reticulum is a protein called
kinectin that bind the ER vesicles to the motor protein ki-
nesin. As nature’s tiniest motors, these motor proteins lit-
erally pull the transport vesicles along the microtubular
tracks. Kinesin uses ATP to power its movement toward
98 Part II Biology of the Cell
FIGURE 5.26
The surfaces of some cells are in constant motion. This
amoeba, a single-celled protist, is advancing toward you, its
advancing edges extending projections outward. The moving
edges have been said to resemble the ruffled edges of a skirt.
the cell periphery, dragging the vesicle with it as it travels
along the microtubule. Another vesicle protein (or per-
haps a slight modification of kinesin—further research is
needed to determine which) binds vesicles to the motor
protein dynein, which directs movement in the opposite
direction, inward toward the cell’s center. (Dynein is also
involved in the movement of eukaryotic flagella, as dis-
cussed below.) The destination of a particular transport
vesicle and its contents is thus determined by the nature
of the linking protein embedded within the vesicle’s
membrane.
Swimming with Flagella and Cilia
Earlier in this chapter, we described the structure of bacterial
flagella. Eukaryotic cells have a completely different kind of
flagellum, consisting of a circle of nine microtubule pairs
surrounding two central microtubules; this arrangement is
referred to as the 9 + 2 structure (figure 5.27). As pairs of
microtubules move past one another using arms composed
of the motor protein dynein, the eukaryotic flagellum undu-
lates rather than rotates. When examined carefully, each
flagellum proves to be an outward projection of the cell’s
interior, containing cytoplasm and enclosed by the plasma
membrane. The microtubules of the flagellum are derived
from a basal body, situated just below the point where the
flagellum protrudes from the surface of the cell.
The flagellum’s complex microtubular apparatus evolved
early in the history of eukaryotes. Although the cells of many
multicellular and some unicellular eukaryotes today no longer
possess flagella and are nonmotile, an organization similar to
the 9 + 2 arrangement of microtubules can still be found within
them, in structures called cilia (singular, cilium). Cilia are
short cellular projections that are often organized in rows (see
figure 5.1). They are more numerous than flagella on the cell
surface, but have the same internal structure. In many multi-
cellular organisms, cilia carry out tasks far removed from
their original function of propelling cells through water. In
several kinds of vertebrate tissues, for example, the beating of
rows of cilia moves water over the tissue surface. The sensory
cells of the vertebrate ear also contain cilia; sound waves bend
these cilia, the initial sensory input of hearing. Thus, the 9 +
2 structure of flagella and cilia appears to be a fundamental
component of eukaryotic cells.
Some eukaryotic cells use pseudopodia to crawl about
within multicellular organisms, while many protists
swim using flagella and cilia. Materials are transported
within cells by special motor proteins.
Chapter 5 Cell Structure 99
Outer microtubule pair
Microtubules
Flagellum
Basal body
Plasma
membrane
Dynein arm
Radial spoke
Central
microtubule
pair
(a)
(b)
(c) (d)
4.36 μm
FIGURE 5.27
Flagella and cilia. (a) A eukaryotic flagellum originates directly from a basal body. (b) The flagellum has two microtubules in its core
connected by radial spokes to an outer ring of nine paired microtubules with dynein arms. (c) The basal body consists of nine microtubule
triplets connected by short protein segments. The structure of cilia is similar to that of flagella, but cilia are usually shorter. (d) The surface
of this Paramecium is covered with a dense forest of cilia.
Special Things about Plant Cells
Vacuoles: A Central Storage Compartment
The center of a plant cell usually contains a large, appar-
ently empty space, called the central vacuole (figure
5.28). This vacuole is not really empty; it contains large
amounts of water and other materials, such as sugars, ions,
and pigments. The central vacuole functions as a storage
center for these important substances and also helps to in-
crease the surface-to-volume ratio of the plant cell by ap-
plying pressure to the cell membrane. The cell membrane
expands outward under this pressure, thereby increasing
its surface area.
Cell Walls: Protection and Support
Plant cells share a characteristic with bacteria that is not
shared with animal cells—that is, plants have cell walls,
which protect and support the plant cell. Although bacteria
also have cell walls, plant cell walls are chemically and struc-
turally different from bacterial cell walls. Cell walls are also
present in fungi and some protists. In plants, cell walls are
composed of fibers of the polysaccharide cellulose. Primary
walls are laid down when the cell is still growing, and be-
tween the walls of adjacent cells is a sticky substance called
the middle lamella, which glues the cells together (figure
5.29). Some plant cells produce strong secondary walls,
which are deposited inside the primary walls of fully ex-
panded cells.
Plant cells store substances in a large central vacuole,
and encase themselves within a strong cellulose cell wall.
100 Part II Biology of the Cell
Cell
Middle lamella
Primary wall
Secondary wall
FIGURE 5.29
Cell walls in plants. As shown in this drawing (a) and transmission electron micrograph (b), plant cell walls are thicker, stronger, and more
rigid than those of bacteria. Primary cell walls are laid down when the cell is young. Thicker secondary cell walls may be added later when
the cell is fully grown.
Primary
walls
Cell 1
Secondary
wall
Cell 2
(b)
Middle
lamella
(a)
1.83 μm
FIGURE 5.28
The central vacuole. A plant’s central vacuole stores dissolved
substances and can increase in size to increase the surface area of a
plant cell.
Chapter 5 Cell Structure 101
Endosymbiosis
Symbiosis is a close relationship between organisms of differ-
ent species that live together. The theory of endosymbiosis
proposes that some of today’s eukaryotic organelles evolved
by a symbiosis in which one species of prokaryote was en-
gulfed by and lived inside another species of prokaryote that
was a precursor to eukaryotes (figure 5.30). According to the
endosymbiont theory, the engulfed prokaryotes provided
their hosts with certain advantages associated with their spe-
cial metabolic abilities. Two key eukaryotic organelles are
believed to be the descendants of these endosymbiotic
prokaryotes: mitochondria, which are thought to have origi-
nated as bacteria capable of carrying out oxidative metabo-
lism; and chloroplasts, which apparently arose from photo-
synthetic bacteria.
The endosymbiont theory is supported by a wealth of
evidence. Both mitochondria and chloroplasts are sur-
rounded by two membranes; the inner membrane proba-
bly evolved from the plasma membrane of the engulfed
bacterium, while the outer membrane is probably derived
from the plasma membrane or endoplasmic reticulum of
the host cell. Mitochondria are about the same size as most
bacteria, and the cristae formed by their inner membranes
resemble the folded membranes in various groups of bac-
teria. Mitochondrial ribosomes are also similar to bacterial
ribosomes in size and structure. Both mitochondria and
chloroplasts contain circular molecules of DNA similar to
those in bacteria. Finally, mitochondria divide by simple
fission, splitting in two just as bacterial cells do, and they
apparently replicate and partition their DNA in much the
same way as bacteria. Table 5.2 compares and reviews the
features of three types of cells.
Some eukaryotic organelles are thought to have arisen
by endosymbiosis.
5.4 Symbiosis played a key role in the origin of some eukaryotic organelles.
FIGURE 5.30
Endosymbiosis. This figure shows how a double membrane may
have been created during the symbiotic origin of mitochondria or
chloroplasts.
Table 5.2 A Comparison of Bacterial, Animal, and Plant Cells
Bacterium Animal Plant
EXTERIOR STRUCTURES
Cell wall
Cell membrane
Flagella
INTERIOR STRUCTURES
ER
Ribosomes
Microtubules
Centrioles
Golgi apparatus
Nucleus
Mitochondria
Chloroplasts
Chromosomes
Lysosomes
Vacuoles
Present (protein-polysaccharide)
Present
May be present (single strand)
Absent
Present
Absent
Absent
Absent
Absent
Absent
Absent
A single circle of DNA
Absent
Absent
Absent
Present
May be present
Usually present
Present
Present
Present
Present
Present
Present
Absent
Multiple; DNA-protein complex
Usually present
Absent or small
Present (cellulose)
Present
Absent except in sperm of a few species
Usually present
Present
Present
Absent
Present
Present
Present
Present
Multiple; DNA-protein complex
Present
Usually a large single vacuole
Chapter 5
Summary Questions Media Resources
5.1 All organisms are composed of cells.
? The cell is the smallest unit of life. All living things
are made of cells.
? The cell is composed of a nuclear region, which holds
the hereditary apparatus, enclosed within the
cytoplasm.
? In all cells, the cytoplasm is bounded by a membrane
composed of phospholipid and protein.
102 Part II Biology of the Cell
1. What are the three principles
of the cell theory?
2. How does the surface area-
to-volume ratio of cells limit the
size that cells can attain?
5.2 Eukaryotic cells are far more complex than bacterial cells.
? Bacteria, which have prokaryotic cell structure, do
not have membrane-bounded organelles within their
cells. Their DNA molecule is circular.
?The eukaryotic cell is larger and more complex, with
many internal compartments.
3. How are prokaryotes
different from eukaryotes in
terms of their cell walls, interior
organization, and flagella?
?A eukaryotic cell is organized into three principal zones:
the nucleus, the cytoplasm, and the plasma membrane.
Located in the cytoplasm are numerous organelles,
which perform specific functions for the cell.
? Many of these organelles, such as the endoplasmic
reticulum, Golgi apparatus (which gives rise to
lysosomes), and nucleus, are part of a complex
endomembrane system.
? Mitochondria and chloroplasts are part of the energy-
processing system of the cell.
? The cytoskeleton encompasses a variety of fibrous
proteins that provide structural support and perform
other functions for the cell.
? Many eukaryotic cells possess flagella or cilia having a
9 + 2 arrangement of microtubules; sliding of the
microtubules past one another bends these cellular
appendages.
?Cells transport materials long distances within the
cytoplasm by packaging them into vesicles that are
pulled by motor proteins along microtubule tracks.
4. What is the endoplasmic
reticulum? What is its function?
How does rough ER differ from
smooth ER?
5. What is the function of the
Golgi apparatus? How do the
substances released by the Golgi
apparatus make their way to
other locations in the cell?
6. What types of eukaryotic
cells contain mitochondria?
What function do mitochondria
perform?
7. What unique metabolic
activity occurs in chloroplasts?
8. What cellular functions do
centrioles participate in?
9. What kinds of cytoskeleton
fibers are stable and which are
changeable?
10. How do cilia compare with
eukaryotic flagella?
5.3 Take a tour of a eukaryotic cell.
? Present-day mitochondria and chloroplasts probably
evolved as a consequence of early endosymbiosis: the
ancestor of the eukaryotic cell engulfed a bacterium,
and the bacterium continued to function within the
host cell.
11. What is the endosymbiont
theory? What is the evidence
supporting this theory?
5.4 Symbiosis played a key role in the origin of some eukaryotic organelles.
? Exploration: Cell Size
? Surface to Volume
? Art Activities:
-Animal Cell Structure
-Plant Cell Structure
-Nonphotosynthetic
Bacterium
-Cyanobacterium
? Scientists on Science:
The Joy of Discovery
? Art Quizzes:
-Nucleosomes
-Rough ER and
Protein Synthesis
-Protein Transport
? Art Activities:
-Anatomy of the
Nucleus
-Golgi Apparatus
Structure
-Mitochondrion
Structure
-Organization of
Cristae
-Chloroplast Structure
-The Cytoskeleton
-Plant Cell
? Endomembrane
?Energy Organelles
? Cytoskeleton
BIOLOGY
RAVEN
JOHNSON
SIX TH
EDITION
www.mhhe.com/raven6/resources5.mhtml
103
6
Membranes
Concept Outline
6.1 Biological membranes are fluid layers of lipid.
The Phospholipid Bilayer. Cells are encased by
membranes composed of a bilayer of phospholipid.
The Lipid Bilayer Is Fluid. Because individual
phospholipid molecules do not bind to one another, the lipid
bilayer of membranes is a fluid.
6.2 Proteins embedded within the plasma membrane
determine its character.
The Fluid Mosaic Model. A varied collection of proteins
float within the lipid bilayer.
Examining Cell Membranes. Visualizing a plasma
membrane requires a powerful electron microscope.
Kinds of Membrane Proteins. The proteins in a
membrane function in support, transport, recognition, and
reactions.
Structure of Membrane Proteins. Membrane proteins are
anchored into the lipid bilayer by their nonpolar regions.
6.3 Passive transport across membranes moves down
the concentration gradient.
Diffusion. Random molecular motion results in a net
movement of molecules to regions of lower concentration.
Facilitated Diffusion. Passive movement across a
membrane is often through specific carrier proteins.
Osmosis. Polar solutes interact with water and can affect
the movement of water across semipermeable membranes.
6.4 Bulk transport utilizes endocytosis.
Bulk Passage Into and Out of the Cell. To transport large
particles, membranes form vesicles.
6.5 Active transport across membranes is powered by
energy from ATP.
Active Transport. Cells transport molecules up a
concentration gradient using ATP-powered carrier proteins.
Coupled Transport. Active transport of ions drives coupled
uptake of other molecules up their concentration gradients.
A
mong a cell’s most important activities are its interac-
tions with the environment, a give and take that never
ceases. Without it, life could not persist. While living cells
and eukaryotic organelles (figure 6.1) are encased within a
lipid membrane through which few water-soluble sub-
stances can pass, the membrane contains protein passage-
ways that permit specific substances to move in and out of
the cell and allow the cell to exchange information with its
environment. We call this delicate skin of protein mole-
cules embedded in a thin sheet of lipid a plasma mem-
brane. This chapter will examine the structure and func-
tion of this remarkable membrane.
FIGURE 6.1
Membranes within a human cell. Sheets of endoplasmic
reticulum weave through the cell interior. The large oval is a
mitochondrion, itself filled with extensive internal membranes.
just as a layer of oil impedes the passage of a drop of water
(“oil and water do not mix”). This barrier to the passage of
water-soluble substances is the key biological property of
the lipid bilayer. In addition to the phospholipid molecules
that make up the lipid bilayer, the membranes of every cell
also contain proteins that extend through the lipid bilayer,
providing passageways across the membrane.
The basic foundation of biological membranes is a
lipid bilayer, which forms spontaneously. In such a
layer, the nonpolar hydrophobic tails of phospholipid
molecules point inward, forming a nonpolar barrier to
water-soluble molecules.
104 Part II Biology of the Cell
The Phospholipid Bilayer
The membranes that encase all living cells are sheets of
lipid only two molecules thick; more than 10,000 of these
sheets piled on one another would just equal the thickness
of this sheet of paper. The lipid layer that forms the foun-
dation of a cell membrane is composed of molecules called
phospholipids (figure 6.2).
Phospholipids
Like the fat molecules you studied in chapter 3, a phos-
pholipid has a backbone derived from a three-carbon
molecule called glycerol. Attached to this backbone are
fatty acids, long chains of carbon atoms ending in a car-
boxyl (—COOH) group. A fat molecule has three such
chains, one attached to each carbon in the backbone; be-
cause these chains are nonpolar, they do not form hydro-
gen bonds with water, and the fat molecule is not water-
soluble. A phospholipid, by contrast, has only two fatty
acid chains attached to its backbone. The third carbon on
the backbone is attached instead to a highly polar organic
alcohol that readily forms hydrogen bonds with water.
Because this alcohol is attached by a phosphate group,
the molecule is called a phospholipid.
One end of a phospholipid molecule is, therefore,
strongly nonpolar (water-insoluble), while the other end is
strongly polar (water-soluble). The two nonpolar fatty
acids extend in one direction, roughly parallel to each
other, and the polar alcohol group points in the other di-
rection. Because of this structure, phospholipids are often
diagrammed as a polar head with two dangling nonpolar
tails (as in figure 6.2b).
Phospholipids Form Bilayer Sheets
What happens when a collection of phospholipid molecules
is placed in water? The polar water molecules repel the
long nonpolar tails of the phospholipids as the water mole-
cules seek partners for hydrogen bonding. Due to the polar
nature of the water molecules, the nonpolar tails of the
phospholipids end up packed closely together, sequestered
as far as possible from water. Every phospholipid molecule
orients to face its polar head toward water and its nonpolar
tails away. When two layers form with the tails facing each
other, no tails ever come in contact with water. The result-
ing structure is called a lipid bilayer (figure 6.3). Lipid bi-
layers form spontaneously, driven by the tendency of water
molecules to form the maximum number of hydrogen
bonds.
The nonpolar interior of a lipid bilayer impedes the pas-
sage of any water-soluble substances through the bilayer,
6.1 Biological membranes are f luid layers of lipid.
Fatty acid
Phosphorylated
alcohol
(a)
(b)
Polar
(hydrophilic) region
Nonpolar (hydrophobic) region
Fatty acid
G
L
Y
C
E
R
O
L
FIGURE 6.2
Phospholipid structure. (a) A phospholipid is a composite
molecule similar to a triacylglycerol, except that only two fatty
acids are bound to the glycerol backbone; a phosphorylated
alcohol occupies the third position on the backbone. (b) Because
the phosphorylated alcohol usually extends from one end of the
molecule and the two fatty acid chains extend from the other,
phospholipids are often diagrammed as a polar head with two
nonpolar hydrophobic tails.
The Lipid Bilayer Is Fluid
A lipid bilayer is stable because water’s affinity for hydro-
gen bonding never stops. Just as surface tension holds a
soap bubble together, even though it is made of a liquid, so
the hydrogen bonding of water holds a membrane to-
gether. But while water continually drives phospholipid
molecules into this configuration, it does not locate specific
phospholipid molecules relative to their neighbors in the
bilayer. As a result, individual phospholipids and unan-
chored proteins are free to move about within the mem-
brane. This can be demonstrated vividly by fusing cells and
watching their proteins reassort (figure 6.4).
Phospholipid bilayers are fluid, with the viscosity of
olive oil (and like oil, their viscosity increases as the tem-
perature decreases). Some membranes are more fluid than
others, however. The tails of individual phospholipid mole-
cules are attracted to one another when they line up close
together. This causes the membrane to become less fluid,
because aligned molecules must pull apart from one an-
other before they can move about in the membrane. The
greater the degree of alignment, the less fluid the mem-
brane. Some phospholipid tails do not align well because
they contain one or more double bonds between carbon
atoms, introducing kinks in the tail. Membranes containing
such phospholipids are more fluid than membranes that
lack them. Most membranes also contain steroid lipids like
cholesterol, which can either increase or decrease mem-
brane fluidity, depending on temperature.
The lipid bilayer is liquid like a soap bubble, rather than
solid like a rubber balloon.
Chapter 6 Membranes 105
Polar
hydrophilic
heads
Nonpolar
hydrophobic
tails
Polar
hydrophilic
heads
FIGURE 6.3
A phospholipid bilayer. The basic structure of every plasma membrane is a double layer of lipid, in which phospholipids aggregate to
form a bilayer with a nonpolar interior. The phospholipid tails do not align perfectly and the membrane is “fluid.” Individual phospholipid
molecules can move from one place to another in the membrane.
Mouse cell
Fusion of
cells
Intermixed membrane
proteins
Human cell
FIGURE 6.4
Proteins move about in membranes. Protein movement within
membranes can be demonstrated easily by labeling the plasma
membrane proteins of a mouse cell with fluorescent antibodies
and then fusing that cell with a human cell. At first, all of the
mouse proteins are located on the mouse side of the fused cell and
all of the human proteins are located on the human side of the
fused cell. However, within an hour, the labeled and unlabeled
proteins are intermixed throughout the hybrid cell’s plasma
membrane.
The Fluid Mosaic Model
A plasma membrane is composed of both lipids and glob-
ular proteins. For many years, biologists thought the pro-
tein covered the inner and outer surfaces of the phospho-
lipid bilayer like a coat of paint. The widely accepted
Davson-Danielli model, proposed in 1935, portrayed the
membrane as a sandwich: a phospholipid bilayer between
two layers of globular protein. This model, however, was
not consistent with what researchers were learning in the
1960s about the structure of membrane proteins. Unlike
most proteins found within cells, membrane proteins are
not very soluble in water—they possess long stretches of
nonpolar hydrophobic amino acids. If such proteins in-
deed coated the surface of the lipid bilayer, as the
Davson-Danielli model suggests, then their nonpolar por-
tions would separate the polar portions of the phospho-
lipids from water, causing the bilayer to dissolve! Because
this doesn’t happen, there is clearly something wrong
with the model.
In 1972, S. Singer and G. Nicolson revised the model in
a simple but profound way: they proposed that the globular
proteins are inserted into the lipid bilayer, with their nonpo-
lar segments in contact with the nonpolar interior of the
bilayer and their polar portions protruding out from the
membrane surface. In this model, called the fluid mosaic
model, a mosaic of proteins float in the fluid lipid bilayer
like boats on a pond (figure 6.5).
Components of the Cell Membrane
A eukaryotic cell contains many membranes. While they
are not all identical, they share the same fundamental ar-
chitecture. Cell membranes are assembled from four com-
ponents (table 6.1):
1. Lipid bilayer. Every cell membrane is composed of
a phospholipid bilayer. The other components of the
membrane are enmeshed within the bilayer, which
provides a flexible matrix and, at the same time, im-
poses a barrier to permeability.
106 Part II Biology of the Cell
6.2 Proteins embedded within the plasma membrane determine its character.
Extracellular fluid
Carbohydrate
Glycolipid
Transmembrane
protein
Glycoprotein
Peripheral
protein
Cholesterol
Filaments of
cytoskeleton
Cytoplasm
FIGURE 6.5
The fluid mosaic model of the plasma membrane. A variety of proteins protrude through the plasma membrane of animal cells, and
nonpolar regions of the proteins tether them to the membrane’s nonpolar interior. The three principal classes of membrane proteins are
transport proteins, receptors, and cell surface markers. Carbohydrate chains are often bound to the extracellular portion of these proteins,
as well as to the membrane phospholipids. These chains serve as distinctive identification tags, unique to particular cells.
2. Transmembrane proteins. A major component of
every membrane is a collection of proteins that float
on or in the lipid bilayer. These proteins provide pas-
sageways that allow substances and information to
cross the membrane. Many membrane proteins are
not fixed in position; they can move about, as the
phospholipid molecules do. Some membranes are
crowded with proteins, while in others, the proteins
are more sparsely distributed.
3. Network of supporting fibers. Membranes are
structurally supported by intracellular proteins that
reinforce the membrane’s shape. For example, a red
blood cell has a characteristic biconcave shape because
a scaffold of proteins called spectrin links proteins in
the plasma membrane with actin filaments in the cell’s
cytoskeleton. Membranes use networks of other pro-
teins to control the lateral movements of some key
membrane proteins, anchoring them to specific sites.
4. Exterior proteins and glycolipids. Membrane
sections assemble in the endoplasmic reticulum,
transfer to the Golgi complex, and then are trans-
ported to the plasma membrane. The endoplasmic
reticulum adds chains of sugar molecules to mem-
brane proteins and lipids, creating a “sugar coating”
called the glycocalyx that extends from the membrane
on the outside of the cell only. Different cell types ex-
hibit different varieties of these glycoproteins and
glycolipids on their surfaces, which act as cell identity
markers.
The fluid mosaic model proposes that membrane
proteins are embedded within the lipid bilayer.
Membranes are composed of a lipid bilayer within
which proteins are anchored. Plasma membranes are
supported by a network of fibers and coated on the
exterior with cell identity markers.
Chapter 6 Membranes 107
Table 6.1 Components of the Cell Membrane
Component Composition Function How It Works Example
Phospholipid bilayer
Carriers
Channels
Receptors
Spectrins
Clathrins
Glycoproteins
Glycolipid
Provides permeability
barrier, matrix for
proteins
Transport molecules
across membrane against
gradient
Passively transport
molecules across
membrane
Transmit information
into cell
Determine shape of cell
Anchor certain proteins
to specific sites,
especially on the exterior
cell membrane in
receptor-mediated
endocytosis
“Self”-recognition
Tissue recognition
Excludes water-soluble
molecules from nonpolar
interior of bilayer
“Escort” molecules through
the membrane in a series of
conformational changes
Create a tunnel that acts as a
passage through membrane
Signal molecules bind to cell-
surface portion of the receptor
protein; this alters the portion
of the receptor protein within
the cell, inducing activity
Form supporting scaffold
beneath membrane,
anchored to both membrane
and cytoskeleton
Proteins line coated pits and
facilitate binding to specific
molecules
Create a protein/carbohydrate
chain shape characteristic of
individual
Create a lipid/carbohydrate
chain shape characteristic of
tissue
Phospholipid
molecules
Transmembrane
proteins
Interior protein
network
Cell surface
markers
Bilayer of cell is
impermeable to water-
soluble molecules, like
glucose
Glycophorin carrier for
sugar transport
Sodium and potassium
channels in nerve cells
Specific receptors bind
peptide hormones and
neurotransmitters
Red blood cell
Localization of low-
density lipoprotein
receptor within coated
pits
Major histocompatibility
complex protein
recognized by immune
system
A, B, O blood group
markers
Examining Cell
Membranes
Biologists examine the delicate, filmy struc-
ture of a cell membrane using electron mi-
croscopes that provide clear magnification
to several thousand times. We discussed
two types of electron microscopes in chap-
ter 5: the transmission electron microscope
(TEM) and the scanning electron micro-
scope (SEM). When examining cell mem-
branes with electron microscopy, speci-
mens must be prepared for viewing.
In one method of preparing a specimen,
the tissue of choice is embedded in a hard
matrix, usually some sort of epoxy (figure
6.6). The epoxy block is then cut with a
microtome, a machine with a very sharp
blade that makes incredibly thin slices.
The knife moves up and down as the spec-
imen advances toward it, causing transpar-
ent “epoxy shavings” less than 1 microme-
ter thick to peel away from the block of
tissue. These shavings are placed on a grid
and a beam of electrons is directed
through the grid with the TEM. At the
high magnification an electron microscope
provides, resolution is good enough to re-
veal the double layers of a membrane.
Freeze-fracturing a specimen is another
way to visualize the inside of the mem-
brane. The tissue is embedded in a
medium and quick-frozen with liquid ni-
trogen. The frozen tissue is then “tapped”
with a knife, causing a crack between the
phospholipid layers of membranes. Pro-
teins, carbohydrates, pits, pores, channels,
or any other structure affiliated with the
membrane will pull apart (whole, usually)
and stick with one side of the split mem-
brane. A very thin coating of platinum is
then evaporated onto the fractured surface
forming a replica of “cast” of the surface.
Once the topography of the membrane has
been preserved in the “cast,” the actual tis-
sue is dissolved away, and the “cast” is ex-
amined with electron microscopy, creating
a strikingly different view of the mem-
brane (see figure 5.10b).
Visualizing a plasma membrane
requires a very powerful electron
microscope. Electrons can either be
passed through a sample or bounced
off it.
108 Part II Biology of the Cell
1. A small chunk of tissue
containing cells of interest
is preserved chemically.
3. A diamond knife sections the
tissue-epoxy block like a loaf of
bread, creating slices 25 nm thick.
2. The tissue is embedded in
epoxy and allowed to harden.
Knife
Forceps
Grid
Section
Tissue
Wax paper
Grid
Section
Lead "stain"
Tissue
Epoxy
4. A tissue section is
mounted on a small grid.
5. The section on the grid is
"stained" with an electron-
dense element (such as
lead).
6. The section is examined by
directing a beam of electrons
through the grid in the transmission
electron microscope (TEM).
7. The high resolution of the TEM
allows detailed examination of
ultrathin sections of tissues and cells.
FIGURE 6.6
Thin section preparation for viewing membranes with electron microscopy.
Kinds of Membrane Proteins
As we’ve seen, the plasma membrane is a complex assem-
bly of proteins enmeshed in a fluid array of phospholipid
molecules. This enormously flexible design permits a
broad range of interactions with the environment, some
directly involving membrane proteins (figure 6.7). Though
cells interact with their environment through their plasma
membranes in many ways, we will focus on six key classes
of membrane protein in this and the following chapter
(chapter 7).
1. Transporters. Membranes are very selective, al-
lowing only certain substances to enter or leave the
cell, either through channels or carriers. In some in-
stances, they take up molecules already present in the
cell in high concentration.
2. Enzymes. Cells carry out many chemical reactions
on the interior surface of the plasma membrane,
using enzymes attached to the membrane.
3. Cell surface receptors. Membranes are exquisitely
sensitive to chemical messages, detecting them with re-
ceptor proteins on their surfaces that act as antennae.
4. Cell surface identity markers. Membranes carry
cell surface markers that identify them to other cells.
Most cell types carry their own ID tags, specific com-
binations of cell surface proteins characteristic of that
cell type.
5. Cell adhesion proteins. Cells use specific proteins
to glue themselves to one another. Some act like Vel-
cro, while others form a more permanent bond.
6. Attachments to the cytoskeleton. Surface pro-
teins that interact with other cells are often anchored
to the cytoskeleton by linking proteins.
The many proteins embedded within a membrane carry
out a host of functions, many of which are associated
with transport of materials or information across the
membrane.
Chapter 6 Membranes 109
Outside
Plasma membrane
Inside
Transporter Cell surface receptorEnzyme
Cell surface identity
marker
Attachment to the
cytoskeleton
Cell adhesion
Figure 6.7
Functions of plasma membrane proteins. Membrane proteins act as transporters, enzymes, cell surface receptors, and cell surface
markers, as well as aiding in cell-to-cell adhesion and securing the cytoskeleton.
Structure of Membrane Proteins
If proteins float on lipid bilayers like ships on the sea, how
do they manage to extend through the membrane to create
channels, and how can certain proteins be anchored into
particular positions on the cell membrane?
Anchoring Proteins in the Bilayer
Many membrane proteins are attached to the surface of the
membrane by special molecules that associate with phos-
pholipids and thereby anchor the protein to the membrane.
Like a ship tied up to a floating dock, these proteins are
free to move about on the surface of the membrane teth-
ered to a phospholipid.
In contrast, other proteins actually traverse the lipid bi-
layer. The part of the protein that extends through the
lipid bilayer, in contact with the nonpolar interior, consists
of one or more nonpolar helices or several β-pleated sheets
of nonpolar amino acids (figure 6.8). Because water avoids
nonpolar amino acids much as it does nonpolar lipid
chains, the nonpolar portions of the protein are held within
the interior of the lipid bilayer. Although the polar ends of
the protein protrude from both sides of the membrane, the
protein itself is locked into the membrane by its nonpolar
segments. Any movement of the protein out of the mem-
brane, in either direction, brings the nonpolar regions of
the protein into contact with water, which “shoves” the
protein back into the interior.
Extending Proteins across the Bilayer
Cells contain a variety of different transmembrane pro-
teins, which differ in the way they traverse the bilayer, de-
pending on their functions.
Anchors. A single nonpolar segment is adequate to an-
chor a protein in the membrane. Anchoring proteins of this
sort attach the spectrin network of the cytoskeleton to the
interior of the plasma membrane (figure 6.9). Many pro-
teins that function as receptors for extracellular signals are
also “single-pass” anchors that pass through the membrane
only once. The portion of the receptor that extends out
from the cell surface binds to specific hormones or other
molecules when the cell encounters them; the binding in-
duces changes at the other end of the protein, in the cell’s
interior. In this way, information outside the cell is trans-
lated into action within the cell. The mechanisms of cell
signaling will be addressed in detail in chapter 7.
Channels. Other proteins have several helical segments
that thread their way back and forth through the mem-
brane, forming a channel like the hole in a doughnut. For
example, bacteriorhodopsin is one of the key transmem-
brane proteins that carries out photosynthesis in bacteria. It
contains seven nonpolar helical segments that traverse the
membrane, forming a circular pore through which protons
pass during the light-driven pumping of protons (figure
6.10). Other transmembrane proteins do not create chan-
nels but rather act as carriers to transport molecules across
the membrane. All water-soluble molecules or ions that
enter or leave the cell are either transported by carriers or
pass through channels.
Pores. Some transmembrane proteins have extensive
nonpolar regions with secondary configurations of β-
pleated sheets instead of α helices. The β sheets form a
characteristic motif, folding back and forth in a circle so the
sheets come to be arranged like the staves of a barrel. This
so-called β barrel, open on both ends, is a common feature
of the porin class of proteins that are found within the
outer membrane of some bacteria (figure 6.11).
Transmembrane proteins are anchored into the bilayer
by their nonpolar segments. While anchor proteins may
pass through the bilayer only once, many channels and
pores are created by proteins that pass back and forth
through the bilayer repeatedly, creating a circular hole
in the bilayer.
110 Part II Biology of the Cell
Phospholipids
Polar areas
of protein
Cholesterol
Nonpolar
areas of
protein
FIGURE 6.8
How nonpolar regions lock proteins into membranes. A
spiral helix of nonpolar amino acids (red) extends across the
nonpolar lipid interior, while polar (purple) portions of the
protein protrude out from the bilayer. The protein cannot move
in or out because such a movement would drag nonpolar
segments of the protein into contact with water.
Chapter 6 Membranes 111
Cytoplasmic side
of cell membrane
Cytoskeletal
proteins
Junctional
complex
100 nm
Ankyrin
Actin
Glycophorin
Spectrin
Linker
protein
FIGURE 6.9
Anchoring proteins. Spectrin extends as a
mesh anchored to the cytoplasmic side of a
red blood cell plasma membrane. The
spectrin protein is represented as a twisted
dimer, attached to the membrane by special
proteins such as junctional complexes and
ankyrin; glycophorins can also be involved in
attachments. This cytoskeletal protein
network confers resiliency to cells like the
red blood cell.
NH
2
H
+
H
+
COOH
Cytoplasm
Retinal
chromophore
Nonpolar
(hydrophobic)
H9251-helices in the
cell membrane
FIGURE 6.10
A channel protein. This transmembrane protein mediates photosynthesis in
the bacterium Halobacterium halobium. The protein traverses the membrane
seven times with hydrophobic helical strands that are within the hydrophobic
center of the lipid bilayer. The helical regions form a channel across the bilayer
through which protons are pumped by the retinal chromophore (green).
Bacterial
outer
membrane
Porin monomer
H9252-pleated sheets
FIGURE 6.11
A pore protein. The bacterial transmembrane protein
porin creates large open tunnels called pores in the outer
membrane of a bacterium. Sixteen strands of β-pleated
sheets run antiparallel to each other, creating a β barrel
in the bacterial outer cell membrane. The tunnel allows
water and other materials to pass through the membrane.
112 Part II Biology of the Cell
Diffusion
Molecules and ions dissolved in water are in constant mo-
tion, moving about randomly. This random motion causes
a net movement of these substances from regions where
their concentration is high to regions where their concen-
tration is lower, a process called diffusion (figure 6.12).
Net movement driven by diffusion will continue until the
concentrations in all regions are the same. You can demon-
strate diffusion by filling a jar to the brim with ink, capping
it, placing it at the bottom of a bucket of water, and then
carefully removing the cap. The ink molecules will slowly
diffuse out from the jar until there is a uniform concentra-
tion in the bucket and the jar. This uniformity in the con-
centration of molecules is a type of equilibrium.
Facilitated Transport
Many molecules that cells require, including glucose and
other energy sources, are polar and cannot pass through
the nonpolar interior of the phospholipid bilayer. These
molecules enter the cell through specific channels in the
plasma membrane. The inside of the channel is polar and
thus “friendly” to the polar molecules, facilitating their
transport across the membrane. Each type of biomolecule
that is transported across the plasma membrane has its own
type of transporter (that is, it has its own channel which fits
it like a glove and cannot be used by other molecules). Each
channel is said to be selective for that type of molecule, and
thus to be selectively permeable, as only molecules admit-
ted by the channels it possesses can enter it. The plasma
membrane of a cell has many types of channels, each selec-
tive for a different type of molecule.
Diffusion of Ions through Channels
One of the simplest ways for a substance to diffuse across a
cell membrane is through a channel, as ions do. Ions are
solutes (substances dissolved in water) with an unequal
number of protons and electrons. Those with an excess of
protons are positively charged and called cations. Ions with
more electrons are negatively charged and called anions.
Because they are charged, ions interact well with polar
molecules like water but are repelled by the nonpolar inte-
rior of a phospholipid bilayer. Therefore, ions cannot move
between the cytoplasm of a cell and the extracellular fluid
without the assistance of membrane transport proteins. Ion
channels possess a hydrated interior that spans the mem-
brane. Ions can diffuse through the channel in either direc-
tion without coming into contact with the hydrophobic
tails of the phospholipids in the membrane, and the trans-
ported ions do not bind to or otherwise interact with the
channel proteins. Two conditions determine the direction
of net movement of the ions: their relative concentrations
on either side of the membrane, and the voltage across the
membrane (a topic we’ll explore in chapter 54). Each type
of channel is specific for a particular ion, such as calcium
(Ca
++
) or chloride (Cl
–
), or in some cases for a few kinds of
ions. Ion channels play an essential role in signaling by the
nervous system.
Diffusion is the net movement of substances to regions
of lower concentration as a result of random
spontaneous motion. It tends to distribute substances
uniformly. Membrane transport proteins allow only
certain molecules and ions to diffuse through the
plasma membrane.
6.3 Passive transport across membranes moves down the concentration gradient.
Lump
of sugar
Sugar
molecule
FIGURE 6.12
Diffusion. If a lump of sugar is dropped into a beaker of water (a), its molecules dissolve (b) and diffuse (c). Eventually, diffusion results in
an even distribution of sugar molecules throughout the water (d).
(a)
(b)
(c)
(d)
Facilitated Diffusion
Carriers, another class of membrane
proteins, transport ions as well as
other solutes like sugars and amino
acids across the membrane. Like
channels, carriers are specific for a
certain type of solute and can trans-
port substances in either direction
across the membrane. Unlike chan-
nels, however, they facilitate the
movement of solutes across the mem-
brane by physically binding to them
on one side of the membrane and re-
leasing them on the other. Again, the
direction of the solute’s net movement
simply depends on its concentration
gradient across the membrane. If the
concentration is greater in the cyto-
plasm, the solute is more likely to
bind to the carrier on the cytoplasmic
side of the membrane and be released
on the extracellular side. This will cause a net movement
from inside to outside. If the concentration is greater in
the extracellular fluid, the net movement will be from out-
side to inside. Thus, the net movement always occurs from
areas of high concentration to low, just as it does in simple
diffusion, but carriers facilitate the process. For this rea-
son, this mechanism of transport is sometimes called facil-
itated diffusion (figure 6.13).
Facilitated Diffusion in Red Blood Cells
Several examples of facilitated diffusion by carrier proteins
can be found in the membranes of vertebrate red blood
cells (RBCs). One RBC carrier protein, for example, trans-
ports a different molecule in each direction: Cl
–
in one di-
rection and bicarbonate ion (HCO
3
–
) in the opposite direc-
tion. As you will learn in chapter 52, this carrier is
important in transporting carbon dioxide in the blood.
A second important facilitated diffusion carrier in RBCs
is the glucose transporter. Red blood cells keep their inter-
nal concentration of glucose low through a chemical trick:
they immediately add a phosphate group to any entering
glucose molecule, converting it to a highly charged glucose
phosphate that cannot pass back across the membrane.
This maintains a steep concentration gradient for glucose,
favoring its entry into the cell. The glucose transporter that
carries glucose into the cell does not appear to form a
channel in the membrane for the glucose to pass through.
Instead, the transmembrane protein appears to bind the
glucose and then flip its shape, dragging the glucose
through the bilayer and releasing it on the inside of the
plasma membrane. Once it releases the glucose, the glucose
transporter reverts to its original shape. It is then available
to bind the next glucose molecule that approaches the out-
side of the cell.
Transport through Selective Channels Saturates
A characteristic feature of transport through selective chan-
nels is that its rate is saturable. In other words, if the con-
centration gradient of a substance is progressively in-
creased, its rate of transport will also increase to a certain
point and then level off. Further increases in the gradient
will produce no additional increase in rate. The explanation
for this observation is that there are a limited number of
carriers in the membrane. When the concentration of the
transported substance rises high enough, all of the carriers
will be in use and the capacity of the transport system will
be saturated. In contrast, substances that move across the
membrane by simple diffusion (diffusion through channels
in the bilayer without the assistance of carriers) do not
show saturation.
Facilitated diffusion provides the cell with a ready way
to prevent the buildup of unwanted molecules within the
cell or to take up needed molecules, such as sugars, that
may be present outside the cell in high concentrations. Fa-
cilitated diffusion has three essential characteristics:
1. It is specific. Any given carrier transports only cer-
tain molecules or ions.
2. It is passive. The direction of net movement is de-
termined by the relative concentrations of the trans-
ported substance inside and outside the cell.
3. It saturates. If all relevant protein carriers are in
use, increases in the concentration gradient do not in-
crease the transport rate.
Facilitated diffusion is the transport of molecules and
ions across a membrane by specific carriers in the
direction of lower concentration of those molecules or
ions.
Chapter 6 Membranes 113
Outside of cell
Inside of cell
FIGURE 6.13
Facilitated diffusion is a carrier-mediated transport process. Molecules bind to a
receptor on the extracellular side of the cell and are conducted through the plasma
membrane by a membrane protein.
Osmosis
The cytoplasm of a cell contains ions and molecules, such
as sugars and amino acids, dissolved in water. The mixture
of these substances and water is called an aqueous solu-
tion. Water, the most common of the molecules in the
mixture, is the solvent, and the substances dissolved in the
water are solutes. The ability of water and solutes to dif-
fuse across membranes has important consequences.
Molecules Diffuse down a Concentration
Gradient
Both water and solutes diffuse from regions of high con-
centration to regions of low concentration; that is, they dif-
fuse down their concentration gradients. When two re-
gions are separated by a membrane, what happens depends
on whether or not the solutes can pass freely through that
membrane. Most solutes, including ions and sugars, are not
lipid-soluble and, therefore, are unable to cross the lipid bi-
layer of the membrane.
Even water molecules, which are very polar, cannot
cross a lipid bilayer. Water flows through aquaporins,
which are specialized channels for water. A simple experi-
ment demonstrates this. If you place an amphibian egg in
hypotonic spring water, it does not swell. If you then inject
aquaporin mRNA into the egg, the channel proteins are ex-
pressed and the egg then swells.
Dissolved solutes interact with water molecules, which
form hydration shells about the charged solute. When there
is a concentration gradient of solutes, the solutes will move
from a high to a low concentration, dragging with them their
hydration shells of water molecules. When a membrane sepa-
rates two solutions, hydration shell water molecules move
with the diffusing ions, creating a net movement of water to-
wards the low solute. This net water movement across a
membrane by diffusion is called osmosis (figure 6.14).
The concentration of all solutes in a solution determines
the osmotic concentration of the solution. If two solu-
tions have unequal osmotic concentrations, the solution
with the higher concentration is hyperosmotic (Greek
hyper, “more than”), and the solution with the lower con-
centration is hypoosmotic (Greek hypo, “less than”). If the
osmotic concentrations of two solutions are equal, the solu-
tions are isosmotic (Greek iso, “the same”).
In cells, a plasma membrane separates two aqueous solu-
tions, one inside the cell (the cytoplasm) and one outside
114 Part II Biology of the Cell
3% salt solution
Selectively
permeable
membrane
Distilled
water
Salt solution
rising
Solution stops rising
when weight of column
equals osmotic
pressure
(a) (b) (c)
FIGURE 6.14
An experiment demonstrating osmosis. (a) The end of a tube
containing a salt solution is closed by stretching a selectively
permeable membrane across its face; the membrane allows the
passage of water molecules but not salt ions. (b) When this tube is
immersed in a beaker of distilled water, the salt cannot cross the
membrane, but water can. The water entering the tube causes the
salt solution to rise in the tube. (c) Water will continue to enter the
tube from the beaker until the weight of the column of water in the
tube exerts a downward force equal to the force drawing water
molecules upward into the tube. This force is referred to as
osmotic pressure.
Shriveled cells Normal cells Cells swell and
eventually burst
Cell body shrinks
from cell wall
Flaccid cell Normal turgid cell
Human red blood cells
Plant cells
Hyperosmotic
solution
Isosmotic
solution
Hypoosmotic
solution
FIGURE 6.15
Osmosis. In a hyperosmotic solution water moves out of the cell
toward the higher concentration of solutes, causing the cell to
shrivel. In an isosmotic solution, the concentration of solutes on
either side of the membrane is the same. Osmosis still occurs, but
water diffuses into and out of the cell at the same rate, and the cell
doesn’t change size. In a hypoosmotic solution the concentration of
solutes is higher within the cell than without, so the net movement
of water is into the cell.
(the extracellular fluid). The direction of the net diffusion
of water across this membrane is determined by the os-
motic concentrations of the solutions on either side (figure
6.15). For example, if the cytoplasm of a cell were hypoos-
motic to the extracellular fluid, water would diffuse out of
the cell, toward the solution with the higher concentration
of solutes (and, therefore, the lower concentration of un-
bound water molecules). This loss of water from the cyto-
plasm would cause the cell to shrink until the osmotic con-
centrations of the cytoplasm and the extracellular fluid
become equal.
Osmotic Pressure
What would happen if the cell’s cytoplasm were hyperos-
motic to the extracellular fluid? In this situation, water
would diffuse into the cell from the extracellular fluid,
causing the cell to swell. The pressure of the cytoplasm
pushing out against the cell membrane, or hydrostatic
pressure, would increase. On the other hand, the osmotic
pressure (figure 6.16), defined as the pressure that must be
applied to stop the osmotic movement of water across a
membrane, would also be at work. If the membrane were
strong enough, the cell would reach an equilibrium, at
which the osmotic pressure, which tends to drive water into
the cell, is exactly counterbalanced by the hydrostatic pres-
sure, which tends to drive water back out of the cell. How-
ever, a plasma membrane by itself cannot withstand large
internal pressures, and an isolated cell under such condi-
tions would burst like an overinflated balloon. Accordingly,
it is important for animal cells to maintain isosmotic condi-
tions. The cells of bacteria, fungi, plants, and many pro-
tists, in contrast, are surrounded by strong cell walls. The
cells of these organisms can withstand high internal pres-
sures without bursting.
Maintaining Osmotic Balance
Organisms have developed many solutions to the osmotic
dilemma posed by being hyperosmotic to their environment.
Extrusion. Some single-celled eukaryotes like the protist
Paramecium use organelles called contractile vacuoles to re-
move water. Each vacuole collects water from various parts
of the cytoplasm and transports it to the central part of the
vacuole, near the cell surface. The vacuole possesses a small
pore that opens to the outside of the cell. By contracting
rhythmically, the vacuole pumps the water out of the cell
through the pore.
Isosmotic Solutions. Some organisms that live in the
ocean adjust their internal concentration of solutes to
match that of the surrounding seawater. Isosmotic with re-
spect to their environment, there is no net flow of water
into or out of these cells. Many terrestrial animals solve the
problem in a similar way, by circulating a fluid through
their bodies that bathes cells in an isosmotic solution. The
blood in your body, for example, contains a high concen-
tration of the protein albumin, which elevates the solute
concentration of the blood to match your cells.
Turgor. Most plant cells are hyperosmotic to their im-
mediate environment, containing a high concentration of
solutes in their central vacuoles. The resulting internal hy-
drostatic pressure, known as turgor pressure, presses the
plasma membrane firmly against the interior of the cell
wall, making the cell rigid. The newer, softer portions of
trees and shrubs depend on turgor pressure to maintain
their shape, and wilt when they lack sufficient water.
Osmosis is the diffusion of water, but not solutes,
across a membrane.
Chapter 6 Membranes 115
Urea
molecule
Water
molecules
Semipermeable
membrane
FIGURE 6.16
How solutes create osmotic pressure.
Charged or polar substances are soluble in
water because they form hydrogen bonds with
water molecules clustered around them. When
a polar solute (illustrated here with urea) is
added to the solution on one side of a
membrane, the water molecules that gather
around each urea molecule are no longer free
to diffuse across the membrane; in effect, the
polar solute has reduced the number of free
water molecules on that side of the membrane
increasing the osmotic pressure. Because the
hypoosmotic side of the membrane (on the
right, with less solute) has more unbound
water molecules than the hyperosmotic side
(on the left, with more solute), water moves by
diffusion from the right to the left.
Bulk Passage Into and
Out of the Cell
Endocytosis
The lipid nature of their biological membranes raises a
second problem for cells. The substances cells use as fuel
are for the most part large, polar molecules that cannot
cross the hydrophobic barrier a lipid bilayer creates. How
do organisms get these substances into their cells? One
process many single-celled eukaryotes employ is endocy-
tosis (figure 6.17). In this process the plasma membrane
extends outward and envelops food particles. Cells use
three major types of endocytosis: phagocytosis, pinocyto-
sis, and receptor-mediated endocytosis.
Phagocytosis and Pinocytosis. If the material the cell
takes in is particulate (made up of discrete particles), such
as an organism or some other fragment of organic matter
(figure 6.17a), the process is called phagocytosis (Greek
phagein, “to eat” + cytos, “cell”). If the material the cell takes
in is liquid (figure 6.17b), it is called pinocytosis (Greek
pinein, “to drink”). Pinocytosis is common among animal
cells. Mammalian egg cells, for example, “nurse” from sur-
rounding cells; the nearby cells secrete nutrients that the
maturing egg cell takes up by pinocytosis. Virtually all eu-
karyotic cells constantly carry out these kinds of endocyto-
sis, trapping particles and extracellular fluid in vesicles and
ingesting them. Endocytosis rates vary from one cell type
to another. They can be surprisingly high: some types of
white blood cells ingest 25% of their cell volume each
hour!
Receptor-Mediated Endocytosis. Specific molecules
are often transported into eukaryotic cells through
receptor-mediated endocytosis. Molecules to be trans-
ported first bind to specific receptors on the plasma mem-
brane. The transport process is specific because only that
molecule has a shape that fits snugly into the receptor. The
plasma membrane of a particular kind of cell contains a
characteristic battery of receptor types, each for a different
kind of molecule.
The interior portion of the receptor molecule resembles
a hook that is trapped in an indented pit coated with the
protein clathrin. The pits act like molecular mousetraps,
closing over to form an internal vesicle when the right mol-
ecule enters the pit (figure 6.18). The trigger that releases
the trap is a receptor protein embedded in the membrane
of the pit, which detects the presence of a particular target
molecule and reacts by initiating endocytosis. The process
is highly specific and very fast.
One type of molecule that is taken up by receptor-
mediated endocytosis is called a low density lipoprotein
(LDL). The LDL molecules bring cholesterol into the cell
where it can be incorporated into membranes. Cholesterol
plays a key role in determining the stiffness of the body’s
membranes. In the human genetic disease called hyper-
cholesteremia, the receptors lack tails and so are never
caught in the clathrin-coated pits and, thus, are never
taken up by the cells. The cholesterol stays in the blood-
stream of affected individuals, coating their arteries and
leading to heart attacks.
Fluid-phase endocytosis is the receptor-mediated
pinocytosis of fluids. It is important to understand that en-
docytosis in itself does not bring substances directly into
the cytoplasm of a cell. The material taken in is still sepa-
rated from the cytoplasm by the membrane of the vesicle.
116 Part II Biology of the Cell
6.4 Bulk transport utilizes endocytosis.
Cytoplasm
Phagocytosis
Pinocytosis
Plasma membrane
Plasma membrane
Nucleus
Cytoplasm
Nucleus
FIGURE 6.17
Endocytosis. Both phagocytosis (a) and pinocytosis (b) are forms
of endocytosis.
(a)
(b)
Exocytosis
The reverse of endocytosis is exocytosis, the discharge of
material from vesicles at the cell surface (figure 6.19). In
plant cells, exocytosis is an important means of exporting
the materials needed to construct the cell wall through the
plasma membrane. Among protists, contractile vacuole dis-
charge is a form of exocytosis. In animal cells, exocytosis
provides a mechanism for secreting many hormones, neuro-
transmitters, digestive enzymes, and other substances.
Cells import bulk materials by engulfing them with
their plasma membranes in a process called endocytosis;
similarly, they extrude or secrete material through
exocytosis.
Chapter 6 Membranes 117
Coated pit
Target molecule
Clathrin
Receptor protein
Coated vesicle
(a)
FIGURE 6.18
Receptor-mediated
endocytosis. (a) Cells that
undergo receptor-mediated
endocytosis have pits coated
with the protein clathrin that
initiate endocytosis when
target molecules bind to
receptor proteins in the
plasma membrane. (b) A
coated pit appears in the
plasma membrane of a
developing egg cell, covered
with a layer of proteins
(80,000×). When an
appropriate collection of
molecules gathers in the
coated pit, the pit deepens (c)
and seals off (d) to form a
coated vesicle, which carries
the molecules into the cell.
(b) (c) (d)
Cytoplasm
Secretory
vesicle
Secretory
product
Plasma
membrane
(a) (b)
FIGURE 6.19
Exocytosis. (a) Proteins and other molecules are secreted from cells in small packets called vesicles, whose membranes fuse with the
plasma membrane, releasing their contents to the cell surface. (b) A transmission electron micrograph showing exocytosis.
Active Transport
While diffusion, facilitated diffusion, and osmosis are pas-
sive transport processes that move materials down their
concentration gradients, cells can also move substances
across the membrane up their concentration gradients.
This process requires the expenditure of energy, typically
ATP, and is therefore called active transport. Like facili-
tated diffusion, active transport involves highly selective
protein carriers within the membrane. These carriers bind
to the transported substance, which could be an ion or a
simple molecule like a sugar (figure 6.20), an amino acid, or
a nucleotide to be used in the synthesis of DNA.
Active transport is one of the most important functions
of any cell. It enables a cell to take up additional molecules
of a substance that is already present in its cytoplasm in
concentrations higher than in the extracellular fluid. With-
out active transport, for example, liver cells would be un-
able to accumulate glucose molecules from the blood
plasma, as the glucose concentration is often higher inside
the liver cells than it is in the plasma. Active transport also
enables a cell to move substances from its cytoplasm to the
extracellular fluid despite higher external concentrations.
The Sodium-Potassium Pump
The use of ATP in active transport may be direct or indi-
rect. Lets first consider how ATP is used directly to move
ions against their concentration gradient. More than one-
third of all of the energy expended by an animal cell that is
not actively dividing is used in the active transport of
sodium (Na
+
) and potassium (K
+
) ions. Most animal cells
have a low internal concentration of Na
+
, relative to their
surroundings, and a high internal concentration of K
+
.
They maintain these concentration differences by actively
pumping Na
+
out of the cell and K
+
in. The remarkable
protein that transports these two ions across the cell mem-
brane is known as the sodium-potassium pump (figure
6.21). The cell obtains the energy it needs to operate the
pump from adenosine triphosphate (ATP), a molecule we’ll
learn more about in chapter 8.
The important characteristic of the sodium-potassium
pump is that it is an active transport process, transporting
Na
+
and K
+
from areas of low concentration to areas of
high concentration. This transport up their concentration
gradients is the opposite of the passive transport in diffu-
sion; it is achieved only by the constant expenditure of
metabolic energy. The sodium-potassium pump works
through a series of conformational changes in the trans-
membrane protein:
Step 1. Three sodium ions bind to the cytoplasmic
side of the protein, causing the protein to change its
conformation.
Step 2. In its new conformation, the protein binds a
molecule of ATP and cleaves it into adenosine diphos-
phate and phosphate (ADP + P
i
). ADP is released, but
the phosphate group remains bound to the protein. The
protein is now phosphorylated.
Step 3. The phosphorylation of the protein induces a
second conformational change in the protein. This
change translocates the three Na
+
across the membrane,
118 Part II Biology of the Cell
6.5 Active transport across membranes is powered by energy from ATP.
Exterior
Cytoplasm
Glucose-binding
site
Hydrophobic
Hydrophilic
Charged amino
acids
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
–
–
–
–
–
–
–
–
–
–
–
–
–
––
–
,
FIGURE 6.20
A glucose transport channel. The molecular structure of this
particular glucose transport channel is known in considerable
detail. The protein’s 492 amino acids form a folded chain that
traverses the lipid membrane 12 times. Amino acids with charged
groups are less stable in the hydrophobic region of the lipid
bilayer and are thus exposed to the cytoplasm or the extracellular
fluid. Researchers think the center of the protein consists of five
helical segments with glucose-binding sites (in red) facing
inward. A conformational change in the protein transports
glucose across the membrane by shifting the position of the
glucose-binding sites.
so they now face the exterior. In this new conformation,
the protein has a low affinity for Na
+
, and the three
bound Na
+
dissociate from the protein and diffuse into
the extracellular fluid.
Step 4. The new conformation has a high affinity for
K
+
, two of which bind to the extracellular side of the
protein as soon as it is free of the Na
+
.
Step 5. The binding of the K
+
causes another confor-
mational change in the protein, this time resulting in the
dissociation of the bound phosphate group.
Step 6. Freed of the phosphate group, the protein re-
verts to its original conformation, exposing the two K
+
to the cytoplasm. This conformation has a low affinity
for K
+
, so the two bound K
+
dissociate from the protein
and diffuse into the interior of the cell. The original
conformation has a high affinity for Na
+
; when these
ions bind, they initiate another cycle.
Three Na
+
leave the cell and two K
+
enter in every
cycle. The changes in protein conformation that occur
during the cycle are rapid, enabling each carrier to
transport as many as 300 Na
+
per second. The sodium-
potassium pump appears to be ubiquitous in animal cells,
although cells vary widely in the number of pump pro-
teins they contain.
Active transport moves a solute across a membrane up
its concentration gradient, using protein carriers driven
by the expenditure of chemical energy.
Chapter 6 Membranes 119
P
P
P
A
P
P
P
A
Na
+
Extracellular
Intracellular
ATP
ATP
P
P
P
A
ATP
P
P
A
P
ADP
1. Protein in membrane binds intracellular
sodium.
2. ATP phosphorylates protein with bound
sodium.
3. Phosphorylation causes conformational
change in protein, allowing sodium to
leave.
P
P
A
P
ADP
4. Extracellular potassium binds to exposed
sites.
K
+
P
P
A
P
ADP+P
i
5. Binding of potassium causes dephos-
phorylation of protein.
6. Dephosphorylation of protein triggers
change back to original conformation,
potassium moves into cell, and the cycle
repeats.
FIGURE 6.21
The sodium-potassium pump. The protein channel known as the sodium-potassium pump transports sodium (Na
+
) and potassium (K
+
)
ions across the cell membrane. For every three Na
+
that are transported out of the cell, two K
+
are transported into the cell. The sodium-
potassium pump is fueled by ATP.
Coupled Transport
Many molecules are transported into cells up a concentration
gradient through a process that uses ATP indirectly. The
molecules move hand-in-hand with sodium ions or protons
that are moving down their concentration gradients. This type
of active transport, called cotransport, has two components:
1. Establishing the down gradient. ATP is used to
establish the sodium ion or proton down gradient,
which is greater than the up gradient of the molecule
to be transported.
2. Traversing the up gradient. Cotransport proteins
(also called coupled transport proteins) carry the mol-
ecule and either a sodium ion or a proton together
across the membrane.
Because the down gradient of the sodium ion or proton is
greater than the up gradient of the molecule to be trans-
ported, the net movement across the membrane is in the
direction of the down gradient, typically into the cell.
Establishing the Down Gradient
Either the sodium-potassium pump or the proton pump es-
tablishes the down gradient that powers most active trans-
port processes of the cell.
The Sodium-Potassium Pump. The sodium-potassium
pump actively pumps sodium ions out of the cell, powered
by energy from ATP. This establishes a sodium ion con-
centration gradient that is lower inside the cell.
The Proton Pump. The proton pump pumps protons
(H
+
ions) across a membrane using energy derived from
energy-rich molecules or from photosynthesis. This cre-
ates a proton gradient, in which the concentration of pro-
tons is higher on one side of the membrane than the other.
Membranes are impermeable to protons, so the only way
protons can diffuse back down their concentration gradi-
ent is through a second cotransport protein.
Traversing the Up Gradient
Animal cells accumulate many amino acids and sugars against
a concentration gradient: the molecules are transported into
the cell from the extracellular fluid, even though their con-
centrations are higher inside the cell. These molecules couple
with sodium ions to enter the cell down the Na
+
concentra-
tion gradient established by the sodium-potassium pump. In
this cotransport process, Na
+
and a specific sugar or amino
acid simultaneously bind to the same transmembrane protein
on the outside of the cell, called a symport (figure 6.22).
Both are then translocated to the inside of the cell, but in the
process Na
+
moves down its concentration gradient while the
sugar or amino acid moves up its concentration gradient. In
effect, the cell uses some of the energy stored in the Na
+
con-
centration gradient to accumulate sugars and amino acids.
In a related process, called countertransport, the in-
ward movement of Na
+
is coupled with the outward move-
ment of another substance, such as Ca
++
or H
+
. As in co-
transport, both Na
+
and the other substance bind to the
same transport protein, in this case called an antiport, but
in this case they bind on opposite sides of the membrane
and are moved in opposite directions. In countertransport,
the cell uses the energy released as Na
+
moves down its
concentration gradient into the cell to extrude a substance
up its concentration gradient.
The cell uses the proton down gradient established by
the proton pump (figure 6.23) in ATP production. The
movement of protons through their cotransport protein is
coupled to the production of ATP, the energy-storing mol-
ecule we mentioned earlier. Thus, the cell expends energy
to produce ATP, which provides it with a convenient en-
ergy storage form that it can employ in its many activities.
The coupling of the proton pump to ATP synthesis, called
chemiosmosis, is responsible for almost all of the ATP
produced from food (see chapter 9) and all of the ATP pro-
duced by photosynthesis (see chapter 10). We know that
proton pump proteins are ancient because they are present
in bacteria as well as in eukaryotes. The mechanisms for
transport across plasma membranes are summarized in
table 6.2.
Many molecules are cotransported into cells up their
concentration gradients by coupling their movement to
that of sodium ions or protons moving down their
concentration gradients.
120 Part II Biology of the Cell
Outside of cell
Inside of cell
Na
+
Coupled
transport
protein
Sugar
K
+
Na/K
pump
FIGURE 6.22
Cotransport through a coupled transport protein. A
membrane protein transports sodium ions into the cell, down
their concentration gradient, at the same time it transports a sugar
molecule into the cell. The gradient driving the Na
+
entry is so
great that sugar molecules can be brought in against their
concentration gradient.
Chapter 6 Membranes 121
Conformation A
Extracellular
fluid
Cytoplasm
H
+
Conformation AConformation B
H
+
H
+
H
+
H
+
H
+
ATP
ADP+P
i
FIGURE 6.23
The proton pump. In this general model of energy-driven proton pumping, the transmembrane protein that acts as a proton pump is
driven through a cycle of two conformations: A and B. The cycle A→B→A goes only one way, causing protons to be pumped from the
inside to the outside of the membrane. ATP powers the pump.
Table 6.2 Mechanisms for Transport across Cell Membranes
Passage through
Process Membrane How It Works Example
PASSIVE PROCESSES
Diffusion
Facilitated diffusion
Osmosis
ACTIVE PROCESSES
Endocytosis
Phagocytosis
Pinocytosis
Carrier-mediated
endocytosis
Exocytosis
Active transport
Na
+
/K
+
pump
Proton pump
Direct
Protein carrier
Direct
Membrane vesicle
Membrane vesicle
Membrane vesicle
Membrane vesicle
Protein carrier
Protein carrier
Random molecular motion produces net
migration of molecules toward region of lower
concentration
Molecule binds to carrier protein in membrane
and is transported across; net movement is
toward region of lower concentration
Diffusion of water across differentially
permeable membrane
Particle is engulfed by membrane, which folds
around it and forms a vesicle
Fluid droplets are engulfed by membrane,
which forms vesicles around them
Endocytosis triggered by a specific receptor
Vesicles fuse with plasma membrane and eject
contents
Carrier expends energy to export Na
+
against
a concentration gradient
Carrier expends energy to export protons
against a concentration gradient
Movement of oxygen into cells
Movement of glucose into cells
Movement of water into cells
placed in a hypotonic solution
Ingestion of bacteria by white
blood cells
“Nursing” of human egg cells
Cholesterol uptake
Secretion of mucus
Coupled uptake of glucose into
cells against its concentration
gradient
Chemiosmotic generation of ATP
122 Part II Biology of the Cell
Chapter 6
Summary Questions Media Resources
6.1 Biological membranes are fluid layers of lipid.
? Every cell is encased within a fluid bilayer sheet of
phospholipid molecules called the plasma membrane.
1. How would increasing the
number of phospholipids with
double bonds between carbon
atoms in their tails affect the
fluidity of a membrane?
? Proteins that are embedded within the plasma
membrane have their hydrophobic regions exposed to
the hydrophobic interior of the bilayer, and their
hydrophilic regions exposed to the cytoplasm or the
extracellular fluid.
? Membrane proteins can transport materials into or
out of the cell, they can mark the identity of the cell,
or they can receive extracellular information.
2. Describe the two basic types
of structures that are
characteristic of proteins that
span membranes.
6.2 Proteins embedded within the plasma membrane determine its character.
? Diffusion is the kinetic movement of molecules or
ions from an area of high concentration to an area of
low concentration.
? Osmosis is the diffusion of water. Because all
organisms are composed of mostly water, maintaining
osmotic balance is essential to life.
3. If a cell’s cytoplasm were
hyperosmotic to the extracellular
fluid, how would the
concentration of solutes in the
cytoplasm compare with that in
the extracellular fluid?
6.3 Passive transport across membranes moves down the concentration gradient.
? Materials or volumes of fluid that are too large to
pass directly through the cell membrane can move
into or out of cells through endocytosis or exocytosis,
respectively.
? In these processes, the cell expends energy to change
the shape of its plasma membrane, allowing the cell
to engulf materials into a temporary vesicle
(endocytosis), or eject materials by fusing a filled
vesicle with the plasma membrane (exocytosis).
4. How do phagocytosis and
pinocytosis differ?
5. Describe the mechanism of
receptor-mediated endocytosis.
6.4 Bulk transport utilizes endocytosis.
? Cells use active transport to move substances across
the plasma membrane against their concentration
gradients, either accumulating them within the cell or
extruding them from the cell. Active transport
requires energy from ATP, either directly or
indirectly.
6. In what two ways does
facilitated diffusion differ from
simple diffusion across a
membrane?
7. How does active transport
differ from facilitated diffusion?
How is it similar to facilitated
diffusion?
6.5 Active transport across membranes is powered by energy from ATP.
? Membrane Structure
? Art Activity: Fluid
Mosaic Model
? Art Activity:
Membrane Protein
Diversity
? Diffusion
? Osmosis
? Diffusion
? Diffusion
? Osmosis
? Student Research:
Understanding
Membrane Transport
? Exocystosis/
endocytosis
? Exocystosis/
endocytosis
? Exploration: Active
Transport
? Active Transport
? Active Transport
http://www.mhhe.com/raven6e http://www.biocourse.com
123
7
Cell-Cell Interactions
Concept Outline
7.1 Cells signal one another with chemicals.
Receptor Proteins and Signaling between Cells.
Receptor proteins embedded in the plasma membrane
change shape when they bind specific signal molecules,
triggering a chain of events within the cell.
Types of Cell Signaling. Cell signaling can occur between
adjacent cells, although chemical signals called hormones act
over long distances.
7.2 Proteins in the cell and on its surface receive
signals from other cells.
Intracellular Receptors. Some receptors are located
within the cell cytoplasm. These receptors respond to lipid-
soluble signals, such as steroid hormones.
Cell Surface Receptors. Many cell-to-cell signals are
water-soluble and cannot penetrate membranes. Instead, the
signals are received by transmembrane proteins protruding
out from the cell surface.
7.3 Follow the journey of information into the cell.
Initiating the Intracellular Signal. Cell surface receptors
often use “second messengers” to transmit a signal to the
cytoplasm.
Amplifying the Signal: Protein Kinase Cascades.
Surface receptors and second messengers amplify signals as
they travel into the cell, often toward the cell nucleus.
7.4 Cell surface proteins mediate cell-cell interactions.
The Expression of Cell Identity. Cells possess on their
surfaces a variety of tissue-specific identity markers that
identify both the tissue and the individual.
Intercellular Adhesion. Cells attach themselves to one
another with protein links. Some of the links are very strong,
others more transient.
Tight Junctions. Adjacent cells form a sheet when
connected by tight junctions, and molecules are encouraged
to flow through the cells, not between them.
Anchoring Junctions. The cytoskeleton of a cell is
connected by an anchoring junction to the cytoskeleton of
another cell or to the extracellular matrix.
Communicating Junctions. Many adjacent cells have
direct passages that link their cytoplasms, permitting the
passage of ions and small molecules.
D
id you know that each of the 100 trillion cells of your
body shares one key feature with the cells of tigers,
bumblebees, and persimmons (figure 7.1)—a feature that
most bacteria and protists lack? Your cells touch and com-
municate with one another. Sending and receiving a variety
of chemical signals, they coordinate their behavior so that
your body functions as an integrated whole, rather than as a
massive collection of individual cells acting independently.
The ability of cells to communicate with one another is the
hallmark of multicellular organisms. In this chapter we will
look in detail at how the cells of multicellular organisms in-
teract with one another, first exploring how they signal one
another with chemicals and then examining the ways in
which their cell surfaces interact to organize tissues and
body structures.
FIGURE 7.1
Persimmon cells in close contact with one another. These
plant cells and all cells, no matter what their function, interact
with their environment, including the cells around them.
Monoclonal antibodies. The first method uses mono-
clonal antibodies. An antibody is an immune system pro-
tein that, like a receptor, binds specifically to another
molecule. Each individual immune system cell can make
only one particular type of antibody, which can bind to
only one specific target molecule. Thus, a cell-line de-
rived from a single immune system cell (a clone) makes
one specific antibody (a monoclonal antibody). Mono-
clonal antibodies that bind to particular receptor pro-
teins can be used to isolate those proteins from the
thousands of others in the cell.
Gene analysis. The study of mutants and isolation of
gene sequences has had a tremendous impact on the
field of receptor analysis. In chapter 19 we will present a
detailed account of how this is done. These advances
make it feasible to identify and isolate the many genes
that encode for various receptor proteins.
Remarkably, these techniques have revealed that the
enormous number of receptor proteins can be grouped into
just a handful of “families” containing many related recep-
tors. Later in this chapter we will meet some of the mem-
bers of these receptor families.
Cells in a multicellular organism communicate with
others by releasing signal molecules that bind to
receptor proteins on the surface of the other cells.
Recent advances in protein isolation have yielded a
wealth of information about the structure and function
of these proteins.
124 Part II Biology of the Cell
Receptor Proteins and
Signaling between Cells
Communication between cells is com-
mon in nature. Cell signaling occurs
in all multicellular organisms, provid-
ing an indispensable mechanism for
cells to influence one another. The
cells of multicellular organisms use a
variety of molecules as signals, includ-
ing not only peptides, but also large
proteins, individual amino acids, nu-
cleotides, steroids and other lipids.
Even dissolved gases are used as
signals. Nitric oxide (NO) plays a role
in mediating male erections (Viagra
functions by stimulating NO release).
Some of these molecules are at-
tached to the surface of the signaling
cell; others are secreted through the
plasma membrane or released by
exocytosis.
Cell Surface Receptors
Any given cell of a multicellular organism is exposed to a
constant stream of signals. At any time, hundreds of differ-
ent chemical signals may be in the environment surround-
ing the cell. However, each cell responds only to certain
signals and ignores the rest (figure 7.2), like a person fol-
lowing the conversation of one or two individuals in a
noisy, crowded room. How does a cell “choose” which sig-
nals to respond to? Located on or within the cell are re-
ceptor proteins, each with a three-dimensional shape that
fits the shape of a specific signal molecule. When a signal
molecule approaches a receptor protein of the right shape,
the two can bind. This binding induces a change in the re-
ceptor protein’s shape, ultimately producing a response in
the cell. Hence, a given cell responds to the signal mole-
cules that fit the particular set of receptor proteins it pos-
sesses, and ignores those for which it lacks receptors.
The Hunt for Receptor Proteins
The characterization of receptor proteins has presented a
very difficult technical problem, because of their relative
scarcity in the cell. Because these proteins may constitute
less than 0.01% of the total mass of protein in a cell, puri-
fying them is analogous to searching for a particular grain
of sand in a sand dune! However, two recent techniques
have enabled cell biologists to make rapid progress in this
area.
7.1 Cells signal one another with chemicals.
Cytoplasm
Signal
molecules
Extracellular
surface
FIGURE 7.2
Cell surface receptors recognize only specific molecules. Signal molecules will bind
only to those cells displaying receptor proteins with a shape into which they can fit snugly.
Types of Cell Signaling
Cells communicate through any of four
basic mechanisms, depending primarily
on the distance between the signaling
and responding cells (figure 7.3). In ad-
dition to using these four basic mecha-
nisms, some cells actually send signals
to themselves, secreting signals that
bind to specific receptors on their own
plasma membranes. This process,
called autocrine signaling, is thought
to play an important role in reinforcing
developmental changes.
Direct Contact
As we saw in chapter 6, the surface of a
eukaryotic cell is a thicket of proteins,
carbohydrates, and lipids attached to
and extending outward from the
plasma membrane. When cells are very
close to one another, some of the mol-
ecules on the cells’ plasma membranes
may bind together in specific ways.
Many of the important interactions
between cells in early development
occur by means of direct contact be-
tween cell surfaces (figure 7.3a). We’ll
examine contact-dependent interac-
tions more closely later in this chapter.
Paracrine Signaling
Signal molecules released by cells can diffuse through the
extracellular fluid to other cells. If those molecules are
taken up by neighboring cells, destroyed by extracellular
enzymes, or quickly removed from the extracellular fluid in
some other way, their influence is restricted to cells in the
immediate vicinity of the releasing cell. Signals with such
short-lived, local effects are called paracrine signals (figure
7.3b). Like direct contact, paracrine signaling plays an im-
portant role in early development, coordinating the activi-
ties of clusters of neighboring cells.
Endocrine Signaling
If a released signal molecule remains in the extracellular fluid,
it may enter the organism’s circulatory system and travel
widely throughout the body. These longer lived signal mole-
cules, which may affect cells very distant from the releasing
cell, are called hormones, and this type of intercellular com-
munication is known as endocrine signaling (figure 7.3c).
Chapter 58 discusses endocrine signaling in detail. Both ani-
mals and plants use this signaling mechanism extensively.
Synaptic Signaling
In animals, the cells of the nervous system provide rapid
communication with distant cells. Their signal molecules,
neurotransmitters, do not travel to the distant cells
through the circulatory system like hormones do. Rather,
the long, fiberlike extensions of nerve cells release neuro-
transmitters from their tips very close to the target cells
(figure 7.3d). The narrow gap between the two cells is
called a chemical synapse. While paracrine signals move
through the fluid between cells, neurotransmitters cross the
synapse and persist only briefly. We will examine synaptic
signaling more fully in chapter 54.
Adjacent cells can signal others by direct contact, while
nearby cells that are not touching can communicate
through paracrine signals. Two other systems mediate
communication over longer distances: in endocrine
signaling the blood carries hormones to distant cells,
and in synaptic signaling nerve cells secrete
neurotransmitters from long cellular extensions close to
the responding cells.
Chapter 7 Cell-Cell Interactions 125
(d) Synaptic signaling
Nerve cell
Neurotransmitter
Synaptic gap
Target cell
(c) Endocrine signaling
Hormone secretion into
blood by endocrine gland
Blood vessel
Distant target cells
Gap
junction
(b) Paracrine signaling
Adjacent target cells
Secretory cell
(a) Direct contact
FIGURE 7.3
Four kinds of cell signaling. Cells communicate in several ways. (a) Two cells in direct
contact with each other may send signals across gap junctions. (b) In paracrine signaling,
secretions from one cell have an effect only on cells in the immediate area. (c) In endocrine
signaling, hormones are released into the circulatory system, which carries them to the target
cells. (d) Chemical synaptic signaling involves transmission of signal molecules, called
neurotransmitters, from a neuron over a small synaptic gap to the target cell.
Intracellular Receptors
All cell signaling pathways share certain common elements,
including a chemical signal that passes from one cell to an-
other and a receptor that receives the signal in or on the
target cell. We’ve looked at the sorts of signals that pass
from one cell to another. Now let’s consider the nature of
the receptors that receive the signals. Table 7.1 summarizes
the types of receptors we will discuss in this chapter.
Many cell signals are lipid-soluble or very small mole-
cules that can readily pass across the plasma membrane of
the target cell and into the cell, where they interact with a
receptor. Some bind to protein receptors located in the cy-
toplasm; others pass across the nuclear membrane as well
and bind to receptors within the nucleus. These intracel-
lular receptors (figure 7.4) may trigger a variety of re-
sponses in the cell, depending on the receptor.
126 Part II Biology of the Cell
7.2 Proteins in the cell and on its surface receive signals from other cells.
Signal molecule-
binding siteInhibitor
protein
DNA binding
domain
Transcription-activating
domain
Signal molecule
Signal molecule-
binding domain
FIGURE 7.4
Basic structure of a gene-regulating intracellular receptor.
These receptors are located within the cell and function in the
reception of signals such as steroid hormones, vitamin D, and
thyroid hormone.
Table 7.1 Cell Communicating Mechanisms
Mechanism Structure Function Example
INTRACELLULAR RECEPTORS No extracellular signal-binding Receives signals from lipid-soluble Receptors for NO, steroid
site or noncharged, nonpolar small hormone, vitamin D, and
molecules thyroid hormone
CELL SURFACE RECEPTORS
Chemically gated Multipass transmembrane Molecular “gates” triggered Neurons
ion channels protein forming a central pore chemically to open or close
Enzymic receptors Single-pass transmembrane Binds signal extracellularly, Phosphorylation of protein
protein catalyzes response intracellularly kinases
G-protein-linked Seven-pass transmembrane Binding of signal to receptor causes Peptide hormones, rod
receptors protein with cytoplasmic GTP to bind a G protein; G protein, cells in the eyes
binding site for G protein with attached GTP, detaches to
deliver the signal inside the cell
PHYSICAL CONTACT WITH OTHER CELLS
Surface markers Variable; integral proteins or Identify the cell MHC complexes, blood
glycolipids in cell membrane groups, antibodies
Tight junctions Tightly bound, leakproof, Organizing junction: holds cells Junctions between
fibrous protein “belt” that together such that material passes epithelial cells in the gut
surrounds cell through but not between the cells
Desmosomes Intermediate filaments of Anchoring junction: “buttons” Epithelium
cytoskeleton linked to adjoining cells together
cells through cadherins
Adherens junctions Transmembrane fibrous Anchoring junction: “roots” Tissues with high
proteins extracellular matrix to cytoskeleton mechanical stress, such as
the skin
Gap junctions Six transmembrane connexon Communicating junction: allows Excitable tissue such as
proteins creating a “pipe” passage of small molecules from heart muscle
that connects cells cell to cell in a tissue
Plasmodesmata Cytoplasmic connections Communicating junction between Plant tissues
between gaps in adjoining plant cells
plant cell walls
Receptors That Act as Gene Regulators
Some intracellular receptors act as regulators of gene
transcription. Among them are the receptors for steroid
hormones, such as cortisol, estrogen, and progesterone, as
well as the receptors for a number of other small, lipid-
soluble signal molecules, such as vitamin D and thyroid
hormone. All of these receptors have similar structures;
the genes that code for them may well be the evolutionary
descendants of a single ancestral gene. Because of their
structural similarities, they are all part of the intracellular
receptor superfamily.
Each of these receptors has a binding site for DNA. In
its inactive state, the receptor typically cannot bind DNA
because an inhibitor protein occupies the binding site.
When the signal molecule binds to another site on the re-
ceptor, the inhibitor is released and the DNA binding site
is exposed (figure 7.5). The receptor then binds to a spe-
cific nucleotide sequence on the DNA, which activates (or,
in a few instances, suppresses) a particular gene, usually lo-
cated adjacent to the regulatory site.
The lipid-soluble signal molecules that intracellular re-
ceptors recognize tend to persist in the blood far longer
than water-soluble signals. Most water-soluble hormones
break down within minutes, and neurotransmitters within
seconds or even milliseconds. A steroid hormone like corti-
sol or estrogen, on the other hand, persists for hours.
The target cell’s response to a lipid-soluble cell signal
can vary enormously, depending on the nature of the cell.
This is true even when different target cells have the
same intracellular receptor, for two reasons: First, the
binding site for the receptor on the target DNA differs
from one cell type to another, so that different genes are
affected when the signal-receptor complex binds to the
DNA, and second, most eukaryotic genes have complex
controls. We will discuss them in detail in chapter 16,
but for now it is sufficient to note that several different
regulatory proteins are usually involved in reading a eu-
karyotic gene. Thus the intracellular receptor interacts
with different signals in different tissues. Depending on
the cell-specific controls operating in different tissues,
the effect the intracellular receptor produces when it
binds with DNA will vary.
Receptors That Act as Enzymes
Other intracellular receptors act as enzymes. A very inter-
esting example is the receptor for the signal molecule, ni-
tric oxide (NO). A small gas molecule, NO diffuses readily
out of the cells where it is produced and passes directly into
neighboring cells, where it binds to the enzyme guanylyl
cyclase. Binding of NO activates the enzyme, enabling it to
catalyze the synthesis of cyclic guanosine monophosphate
(GMP), an intracellular messenger molecule that produces
cell-specific responses such as the relaxation of smooth
muscle cells.
NO has only recently been recognized as a signal mole-
cule in vertebrates. Already, however, a wide variety of
roles have been documented. For example, when the brain
sends a nerve signal relaxing the smooth muscle cells lining
the walls of vertebrate blood vessels, the signal molecule
acetylcholine released by the nerve near the muscle does
not interact with the muscle cell directly. Instead, it causes
nearby epithelial cells to produce NO, which then causes
the smooth muscle to relax, allowing the vessel to expand
and thereby increase blood flow.
Many target cells possess intracellular receptors, which
are activated by substances that pass through the
plasma membrane.
Chapter 7 Cell-Cell Interactions 127
DNA binding
site blocked
Transcription
activating
domain
DNA binding
site exposed
Cortisol
Inhibitor
Signal molecule-
binding domain
FIGURE 7.5
How intracellular receptors regulate gene transcription. In
this model, the binding of the steroid hormone cortisol to a DNA
regulatory protein causes it to alter its shape. The inhibitor is
released, exposing the DNA binding site of the regulatory
protein. The DNA binds to the site, positioning a specific
nucleotide sequence over the transcription activating domain of
the receptor and initiating transcription.
Cell Surface Receptors
Most signal molecules are water-soluble, including neuro-
transmitters, peptide hormones, and the many proteins
that multicellular organisms employ as “growth factors”
during development. Water-soluble signals cannot diffuse
through cell membranes. Therefore, to trigger responses
in cells, they must bind to receptor proteins on the sur-
face of the cell. These cell surface receptors (figure 7.6)
convert the extracellular signal to an intracellular one, re-
sponding to the binding of the signal molecule by produc-
ing a change within the cell’s cytoplasm. Most of a cell’s
receptors are cell surface receptors, and almost all of them
belong to one of three receptor superfamilies: chemically
gated ion channels, enzymic receptors, and G-protein-
linked receptors.
Chemically Gated Ion Channels
Chemically gated ion channels are receptor proteins that
ions pass through. The receptor proteins that bind many
neurotransmitters have the same basic structure (figure
7.6a). Each is a “multipass” transmembrane protein, mean-
ing that the chain of amino acids threads back and forth
across the plasma membrane several times. In the center of
the protein is a pore that connects the extracellular fluid
with the cytoplasm. The pore is big enough for ions to pass
through, so the protein functions as an ion channel. The
channel is said to be chemically gated because it opens
when a chemical (the neurotransmitter) binds to it. The
type of ion (sodium, potassium, calcium, chloride, for ex-
ample) that flows across the membrane when a chemically
gated ion channel opens depends on the specific three-
dimensional structure of the channel.
Enzymic Receptors
Many cell surface receptors either act as enzymes or are di-
rectly linked to enzymes (figure 7.6b). When a signal mole-
cule binds to the receptor, it activates the enzyme. In al-
most all cases, these enzymes are protein kinases, enzymes
that add phosphate groups to proteins. Most enzymic re-
ceptors have the same general structure. Each is a single-
pass transmembrane protein (the amino acid chain passes
through the plasma membrane only once); the portion that
binds the signal molecule lies outside the cell, and the por-
tion that carries out the enzyme activity is exposed to the
cytoplasm.
128 Part II Biology of the Cell
(a) Chemically gated ion channel
Signal
G protein Activated
G protein
Enzyme or
ion channel
Activated
enzyme or
ion channel
Ions
(b) Enzymic receptor
(c) G-protein-linked receptor
Signal
Inactive
catalytic
domain
Active
catalytic
domain
FIGURE 7.6
Cell surface receptors. (a) Chemically gated ion channels are multipass transmembrane proteins that form a pore in the cell membrane.
This pore is opened or closed by chemical signals. (b) Enzymic receptors are single-pass transmembrane proteins that bind the signal on
the extracellular surface. A catalytic region on their cytoplasmic portion then initiates enzymatic activity inside the cell. (c) G-protein-
linked receptors bind to the signal outside the cell and to G proteins inside the cell. The G protein then activates an enzyme or ion
channel, mediating the passage of a signal from the cell’s surface to its interior.
G-Protein-Linked Receptors
A third class of cell surface receptors acts indirectly on
enzymes or ion channels in the plasma membrane with
the aid of an assisting protein, called a guanosine triphos-
phate (GTP)-binding protein, or G protein (figure 7.6c).
Receptors in this category use G proteins to mediate pas-
sage of the signal from the membrane surface into the
cell interior.
How G-Protein-Linked Receptors Work. G pro-
teins are mediators that initiate a diffusible signal in the
cytoplasm. They form a transient link between the recep-
tor on the cell surface and the signal pathway within the
cytoplasm. Importantly, this signal has a relatively short
life span whose active age is determined by GTP. When
a signal arrives, it finds the G protein nestled into the G-
protein-linked receptor on the cytoplasmic side of the
plasma membrane. Once the signal molecule binds to the
receptor, the G-protein-linked receptor changes shape.
This change in receptor shape twists the G protein, caus-
ing it to bind GTP. The G protein can now diffuse away
from the receptor. The “activated” complex of a G pro-
tein with attached GTP is then free to initiate a number
of events. However, this activation is short-lived, because
GTP has a relatively short life span (seconds to minutes).
This elegant arrangement allows the G proteins to acti-
vate numerous pathways, but only in a transient manner.
In order for a pathway to “stay on,” there must be a con-
tinuous source of incoming extracellular signals. When
the rate of external signal drops off, the pathway shuts
down.
The Largest Family of Cell Surface Receptors. Sci-
entists have identified more than 100 different G-
protein-linked receptors, more than any other kind of
cell surface receptor. They mediate an incredible range
of cell signals, including peptide hormones, neurotrans-
mitters, fatty acids, and amino acids. Despite this great
variation in specificity, however, all G-protein-linked re-
ceptors whose amino acid sequences are known have a
similar structure. They are almost certainly closely re-
lated in an evolutionary sense, arising from a single an-
cestral sequence. Each of these G-protein-linked recep-
tors is a seven-pass transmembrane protein (figure
7.7)—a single polypeptide chain that threads back and
forth across the lipid bilayer seven times, creating a chan-
nel through the membrane.
Evolutionary Origin of G-Protein-Linked Receptors.
As research revealed the structure of G-protein-linked re-
ceptors, an interesting pattern emerged: the same seven-
pass structural motif is seen again and again, in sensory re-
ceptors such as the light-activated rhodopsin protein in the
vertebrate eye, in the light-activated bacteriorhodopsin
proton pump that plays a key role in bacterial photosynthe-
sis, in the receptor that recognizes the yeast mating factor
protein discussed earlier, and in many other sensory recep-
tors. Vertebrate rhodopsin is in fact a G-protein-linked re-
ceptor and utilizes a G protein. Bacteriorhodopsin is not.
The occurrence of the seven-pass structural motif in both,
and in so many other G-protein-linked receptors, suggests
that this motif is a very ancient one, and that G-protein-
linked receptors may have evolved from sensory receptors
of single-celled ancestors.
Discovery of G Proteins. Martin Rodbell of the Na-
tional Institute of Environmental Health Sciences and
Alfred Gilman of the University of Texas Southwestern
Medical Center received the 1994 Nobel Prize for Medi-
cine or Physiology for their work on G proteins. Rodbell
and Gilman’s work has proven to have significant ramifi-
cations. G proteins are involved in the mechanism em-
ployed by over half of all medicines in use today. Study-
ing G proteins will vastly expand our understanding of
how these medicines work. Furthermore, the investiga-
tion of G proteins should help elucidate how cells com-
municate in general and how they contribute to the over-
all physiology of organisms. As Gilman says, G proteins
are “involved in everything from sex in yeast to cognition
in humans.”
Most receptors are located on the surface of the plasma
membrane. Chemically gated ion channels open or
close when signal molecules bind to the channel,
allowing specific ions to diffuse through. Enzyme
receptors typically activate intracellular proteins by
phosphorylation. G-protein-linked receptors activate an
intermediary protein, which then effects the
intracellular change.
Chapter 7 Cell-Cell Interactions 129
Protein signal-binding site
G-protein-binding sites
COOH
NH
2
FIGURE 7.7
The G-protein-linked receptor is a seven-pass transmembrane
protein.
Initiating the
Intracellular Signal
Some enzymic receptors and most G-
protein-linked receptors carry the sig-
nal molecule’s message into the target
cell by utilizing other substances to
relay the message within the cyto-
plasm. These other substances, small
molecules or ions commonly called
second messengers or intracellular
mediators, alter the behavior of partic-
ular proteins by binding to them and
changing their shape. The two most
widely used second messengers are
cyclic adenosine monophosphate
(cAMP) and calcium.
cAMP
All animal cells studied thus far use
cAMP as a second messenger (chap-
ter 56 discusses cAMP in detail). To
see how cAMP typically works as a
messenger, let’s examine what hap-
pens when the hormone epinephrine
binds to a particular type of G-
protein-linked receptor called the β-
adrenergic receptor (figure 7.8).
When epinephrine binds with this re-
ceptor, it activates a G protein, which
then stimulates the enzyme adenylyl
cyclase to produce large amounts of cAMP within the
cell (figure 7.9a). The cAMP then binds to and activates
the enzyme α-kinase, which adds phosphates to specific
proteins in the cell. The effect this phosphorylation has
on cell function depends on the identity of the cell and
the proteins that are phosphorylated. In muscle cells, for
example, the α-kinase phosphorylates and thereby acti-
vates enzymes that stimulate the breakdown of glycogen
into glucose and inhibit the synthesis of glycogen from
glucose. Glucose is then more available to the muscle
cells for metabolism.
Calcium
Calcium (Ca
++
) ions serve even more widely than cAMP
as second messengers. Ca
++
levels inside the cytoplasm of a
cell are normally very low (less than 10
H110027
M), while outside
the cell and in the endoplasmic reticulum Ca
++
levels are
quite high (about 10
H110023
M). Chemically gated calcium chan-
nels in the endoplasmic reticulum membrane act as
switches; when they open, Ca
++
rushes into the cytoplasm
and triggers proteins sensitive to Ca
++
to initiate a variety of
activities. For example, the efflux of Ca
++
from the endo-
plasmic reticulum causes skeletal muscle cells to contract
and some endocrine cells to secrete hormones.
The gated Ca
++
channels are opened by a G-protein-
linked receptor. In response to signals from other cells,
the receptor activates its G protein, which in turn acti-
vates the enzyme, phospholipase C. This enzyme catalyzes
the production of inositol trisphosphate (IP
3
) from phospho-
lipids in the plasma membrane. The IP
3
molecules diffuse
through the cytoplasm to the endoplasmic reticulum and
bind to the Ca
++
channels. This opens the channels and
allows Ca
++
to flow from the endoplasmic reticulum into
the cytoplasm (figure 7.9b).
Ca
++
initiates some cellular responses by binding to
calmodulin, a 148-amino acid cytoplasmic protein that con-
tains four binding sites for Ca
++
(figure 7.10). When four
Ca
++
ions are bound to calmodulin, the calmodulin/Ca
++
complex binds to other proteins, and activates them.
Cyclic AMP and Ca
++
often behave as second
messengers, intracellular substances that relay
messages from receptors to target proteins.
130 Part II Biology of the Cell
7.3 Follow the journey of information into the cell.
Extracellular
Intracellular
NH
3
+
COO
-
Oligosaccharide
unit
FIGURE 7.8
Structure of the β-adrenergic receptor. The receptor is a G-protein-linked molecule
which, when it binds to an extracellular signal molecule, stimulates voluminous production
of cAMP inside the cell, which then effects the cellular change.
Chapter 7 Cell-Cell Interactions 131
Signal molecule Signal molecule
Cell surface receptor
cAMP pathway Ca
++
pathway
Adenylyl cyclase
G protein
cAMP
Target protein
Nucleus
Cell
membrane
Cytoplasm
Nucleus
Cell
membrane
Cytoplasm
Cell surface receptor
Phospholipase C
G protein
Ca
++
Target
protein
Inositol
trisphosphate
intermediaryEndoplasmic
reticulum
ATP
(a) (b)
FIGURE 7.9
How second messengers work. (a) The cyclic AMP (cAMP) pathway. An extracellular receptor binds to a signal molecule and, through a
G protein, activates the membrane-bound enzyme, adenylyl cyclase. This enzyme catalyzes the synthesis of cAMP, which binds to the
target protein to initiate the cellular change. (b) The calcium (Ca
++
) pathway. An extracellular receptor binds to another signal molecule
and, through another G protein, activates the enzyme phospholipase C. This enzyme stimulates the production of inositol trisphosphate,
which binds to and opens calcium channels in the membrane of the endoplasmic reticulum. Ca
++
is released into the cytoplasm, effecting a
change in the cell.
Ca
++
Ca
++
Ca
++
Ca
++
Ca
++
Inactive
protein
Active
protein
Calmodulin
Calmodulin
(a) (b)
FIGURE 7.10
Calmodulin. (a)
Calmodulin is a protein
containing 148 amino acid
residues that mediates
Ca
++
function. (b) When
four Ca
++
are bound to the
calmodulin molecule, it
undergoes a
conformational change
that allows it to bind to
other cytoplasmic proteins
and effect cellular
responses.
Amplifying the Signal: Protein
Kinase Cascades
Both enzyme-linked and G-protein-linked receptors re-
ceive signals at the surface of the cell, but as we’ve seen, the
target cell’s response rarely takes place there. In most cases
the signals are relayed to the cytoplasm or the nucleus by
second messengers, which influence the activity of one or
more enzymes or genes and so alter the behavior of the
cell. But most signaling molecules are found in such low
concentrations that their diffusion across the cytoplasm
would take a great deal of time unless the signal is ampli-
fied. Therefore, most enzyme-linked and G-protein-linked
receptors use a chain of other protein messengers to am-
plify the signal as it is being relayed to the nucleus.
How is the signal amplified? Imagine a relay race where,
at the end of each stage, the finishing runner tags five new
runners to start the next stage. The number of runners
would increase dramatically as the race progresses: 1, then
5, 25, 125, and so on. The same sort of process takes place
as a signal is passed from the cell surface to the cytoplasm
or nucleus. First the receptor activates a stage-one protein,
almost always by phosphorylating it. The receptor either
adds a phosphate group directly, or, it activates a G protein
that goes on to activate a second protein that does the
phosphorylation. Once activated, each of these stage-one
proteins in turn activates a large number of stage-two pro-
132 Part II Biology of the Cell
Signal
molecule
Receptor
protein
Activated
adenylyl cyclase
Amplification
Amplification
Amplification
Amplification
GTP
G protein
2
1
3
4
5
6
7
Enzymatic product
Enzyme
Protein kinase
cAMP
Not yet
activated
FIGURE 7.11
Signal amplification. Amplification at many steps of the cell-signaling process can ultimately produce a large response by the cell. One
cell surface receptor (1), for example, may activate many G protein molecules (2), each of which activates a molecule of adenylyl cyclase
(3), yielding an enormous production of cAMP (4). Each cAMP molecule in turn will activate a protein kinase (5), which can
phosphorylate and thereby activate several copies of a specific enzyme (6). Each of those enzymes can then catalyze many chemical
reactions (7). Starting with 10
H1100210
M of signaling molecule, one cell surface receptor can trigger the production of 10
H110026
M of one of the
products, an amplification of four orders of magnitude.
teins; then each of them activates a large number of stage-
three proteins, and so on (figure 7.11). A single cell surface
receptor can thus stimulate a cascade of protein kinases to
amplify the signal.
The Vision Amplification Cascade
Let’s trace a protein amplification cascade to see exactly
how one works. In vision, a single light-activated rhodopsin
(a G-protein-linked receptor) activates hundreds of mole-
cules of the G protein transducin in the first stage of the
relay. In the second stage, each transducin causes an en-
zyme to modify thousands of molecules of a special inside-
the-cell messenger called cyclic GMP (figure 7.12). (We
will discuss cyclic GMP in more detail later.) In about 1
second, a single rhodopsin signal passing through this two-
step cascade splits more than 10
5
(100,000) cyclic GMP
molecules (figure 7.13)! The rod cells of humans are suffi-
ciently sensitive to detect brief flashes of 5 photons.
The Cell Division Amplification Cascade
The amplification of signals traveling from the plasma
membrane to the nucleus can be even more complex than
the process we’ve just described. Cell division, for example,
is controlled by a receptor that acts as a protein kinase. The
receptor responds to growth-promoting signals by phos-
phorylating an intracellular protein called ras, which then
activates a series of interacting phosphorylation cascades,
some with five or more stages. If the ras protein becomes
hyperactive for any reason, the cell acts as if it is being con-
stantly stimulated to divide. Ras proteins were first discov-
ered in cancer cells. A mutation of the gene that encodes
ras had caused it to become hyperactive, resulting in unre-
strained cell proliferation. Almost one-third of human can-
cers have such a mutation in a ras gene.
A small number of surface receptors can generate a vast
intracellular response, as each stage of the pathway
amplifies the next.
Chapter 7 Cell-Cell Interactions 133
Sugar
Guanine
Phosphate
CH
2
O
–
O OH
O
O
O
NH
2N
N
N
N
O
P
Na
+
Na
+
One rhodopsin molecule
absorbs one photon, which
activates 500 transducin
molecules, which
activate 500 phosphodiesterase
molecules, which
hydrolyze 10
5
cyclic GMP
molecules, which
Na
+
close 250 Na
+
channels, preventing
10
6
–10
7
Na
+
per second from entering
the cell for a period of 1 second, which
hyperpolarizes the rod cell membrane
by 1 mV, sending a visual signal to the brain.
FIGURE 7.13
The role of signal amplification in vision. In this vertebrate rod
cell (the cells of the eye responsible for interpreting light and
dark), one single rhodopsin pigment molecule, when excited by a
photon, ultimately yields 100,000 split cGMP molecules, which
will then effect a change in the membrane of the rod cell, which
will be interpreted by the organism as a visual event.
FIGURE 7.12
Cyclic GMP. Cyclic GMP is a
guanosine monophosphate nucleotide
molecule with the single phosphate
group attached to a sugar residue in
two places (this cyclic
part is shown in
yellow). Cyclic GMP
is an important
second messenger
linking G proteins
to signal
transduction
pathways within
the cytoplasm.
The Expression of Cell Identity
With the exception of a few primitive types of organisms,
the hallmark of multicellular life is the development of
highly specialized groups of cells called tissues, such as
blood and muscle. Remarkably, each cell within a tissue
performs the functions of that tissue and no other, even
though all cells of the body are derived from a single fertil-
ized cell and contain the same genetic information. How
do cells sense where they are, and how do they “know”
which type of tissue they belong to?
Tissue-Specific Identity Markers
As it develops, each animal cell type acquires a unique set
of cell surface molecules. These molecules serve as markers
proclaiming the cells’ tissue-specific identity. Other cells
that make direct physical contact with them “read” the
markers.
Glycolipids. Most tissue-specific cell surface markers are
glycolipids, lipids with carbohydrate heads. The glycolipids
on the surface of red blood cells are also responsible for the
differences among A, B, and O blood types. As the cells in a
tissue divide and differentiate, the population of cell surface
glycolipids changes dramatically.
MHC Proteins. The immune system uses other cell sur-
face markers to distinguish between “self” and “nonself”
cells. All of the cells of a given individual, for example, have
the same “self” markers, called major histocompatibility com-
plex (MHC) proteins. Because practically every individual
makes a different set of MHC proteins, they serve as dis-
tinctive identity tags for each individual. The MHC pro-
teins and other self-identifying markers are single-pass pro-
teins anchored in the plasma membrane, and many of them
are members of a large superfamily of receptors, the im-
munoglobulins (figure 7.14). Cells of the immune system
continually inspect the other cells they encounter in the
body, triggering the destruction of cells that display foreign
or “nonself” identity markers.
The immune systems of vertebrates, described in detail in
chapter 57, shows an exceptional ability to distinguish self
from nonself. However, other vertebrates and even some
simple animals like sponges are able to make this distinc-
tion to some degree, even though they lack a complex im-
mune system.
Every cell contains a specific array of marker proteins
on its surface. These markers identify each type of cell
in a very precise way.
134 Part II Biology of the Cell
7.4 Cell surface proteins mediate cell-cell interactions.
Constant region
Variable region
Disulfide bond
s
s
s
s
s
s
s s
s
s
H9251 chain
H9252 chain
T Receptor
ss
ss
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
s
B Receptor
Light chain
Light chain
Heavy
chains
Cell
membrane
s
s
s
s
s
s
MHC-II
H9251 chain
H9252 chain
s
s
s
s
s
s
MHC-I
H9251 chain
H9252-2
microglobulin
SS
s
FIGURE 7.14
Structure of the immunoglobulin family of cell surface marker proteins. T and B cell receptors help mediate the immune response in
organisms by recognizing and binding to foreign cell markers. MHC antigens label cells as “self,” so that the immune system attacks only
invading entities, such as bacteria, viruses, and usually even the cells of transplanted organs!
Intercellular Adhesion
Not all physical contacts between cells in a multicellular
organism are fleeting touches. In fact, most cells are in
physical contact with other cells at all times, usually as
members of organized tissues such as those in the lungs,
heart, or gut. These cells and the mass of other cells clus-
tered around them form long-lasting or permanent connec-
tions with each other called cell junctions (figure 7.15).
The nature of the physical connections between the cells of
a tissue in large measure determines what the tissue is like.
Indeed, a tissue’s proper functioning often depends criti-
cally upon how the individual cells are arranged within it.
Just as a house cannot maintain its structure without nails
and cement, so a tissue cannot maintain its characteristic
architecture without the appropriate cell junctions.
Cells attach themselves to one another with long-
lasting bonds called cell junctions.
Chapter 7 Cell-Cell Interactions 135
Microvilli
Tight junction
Adherens junction
(anchoring junction)
Intermediate
filament
Desmosome
(anchoring junction)
Gap junction
(communicating junction)
Hemidesmosome
(anchoring junction)
Basal lamina
FIGURE 7.15
A summary of cell junction types. Gut epithelial cells are used here to illustrate the comparative structures and locations of common
cell junctions.
Tight Junctions
Cell junctions are divided into three
categories, based upon the functions
they serve (figure 7.16): tight junc-
tions, anchoring junctions, and com-
municating junctions.
Sometimes called occluding junc-
tions, tight junctions connect the
plasma membranes of adjacent cells in
a sheet, preventing small molecules
from leaking between the cells and
through the sheet (figure 7.17). This
allows the sheet of cells to act as a
wall within the organ, keeping mole-
cules on one side or the other.
Creating Sheets of Cells
The cells that line an animal’s diges-
tive tract are organized in a sheet
only one cell thick. One surface of
the sheet faces the inside of the tract
and the other faces the extracellular
space where blood vessels are lo-
cated. Tight junctions encircle each
cell in the sheet, like a belt cinched
around a pair of pants. The junc-
tions between neighboring cells are so securely attached
that there is no space between them for leakage. Hence,
nutrients absorbed from the food in the digestive tract
must pass directly through the cells in the sheet to enter
the blood.
Partitioning the Sheet
The tight junctions between the cells lining the digestive
tract also partition the plasma membranes of these cells
into separate compartments. Transport proteins in the
membrane facing the inside of the tract carry nutrients
from that side to the cytoplasm of the cells. Other proteins,
located in the membrane on the opposite side of the cells,
transport those nutrients from the cytoplasm to the extra-
cellular fluid, where they can enter the blood. For the sheet
to absorb nutrients properly, these proteins must remain in
the correct locations within the fluid membrane. Tight
junctions effectively segregate the proteins on opposite
sides of the sheet, preventing them from drifting within the
membrane from one side of the sheet to the other. When
tight junctions are experimentally disrupted, just this sort
of migration occurs.
Tight junctions connect the plasma membranes of
adjacent cells into sheets.
136 Part II Biology of the Cell
ER
Tight
junction
Cell
1
Cell 2
Cell
3
Lumen
(a) Tight junction
Cell
1 2
Cytoskeletal
filament
Inter-
cellular
space
Extracellular matrix
Intracellular
attachment
proteins
Plasma
membranes
Transmembrane
linking proteins
(b) Anchoring junction
Central
tubule
Smooth
Cell
1
Cell
2
Primary cell
wall
Middle
lamella
Plasma
membrane
(c) Communicating junction
Cell
FIGURE 7.16
The three types of cell junctions. These three models represent current thinking on how
the structures of the three major types of cell junctions facilitate their function: (a) tight
junction; (b) anchoring junction; (c) communicating junction.
Cell
1
Cell
2
Cell
3
Blood
Glucose
Apical surface
Lumen of gut
Tight junction
Plasma membranes
of adjacent cells
Intercellular space
Extracellular
fluid
FIGURE 7.17
Tight junctions. Encircling the cell like a tight belt, these
intercellular contacts ensure that materials move through the
cells rather than between them.
Anchoring Junctions
Anchoring junctions mechanically attach the cytoskele-
ton of a cell to the cytoskeletons of other cells or to the
extracellular matrix. They are commonest in tissues sub-
ject to mechanical stress, such as muscle and skin
epithelium.
Cadherin and Intermediate Filaments:
Desmosomes
Anchoring junctions called desmosomes connect the cy-
toskeletons of adjacent cells (figure 7.18), while
hemidesmosomes anchor epithelial cells to a basement
membrane. Proteins called cadherins, most of which are
single-pass transmembrane glycoproteins, create the criti-
cal link. A variety of attachment proteins link the short cy-
toplasmic end of a cadherin to the intermediate filaments in
the cytoskeleton. The other end of the cadherin molecule
projects outward from the plasma membrane, joining di-
rectly with a cadherin protruding from an adjacent cell in a
firm handshake binding the cells together.
Connections between proteins tethered to the interme-
diate filaments are much more secure than connections be-
tween free-floating membrane proteins. Proteins are sus-
pended within the membrane by relatively weak
interactions between the nonpolar portions of the protein
and the membrane lipids. It would not take much force to
pull an untethered protein completely out of the mem-
brane, as if pulling an unanchored raft out of the water.
Chapter 7 Cell-Cell Interactions 137
Adjacent plasma
membranes
Cytoplasmic protein
plaque
Cadherin
Intercellular
space
Cytoskeletal filaments
anchored to
cytoplasmic plaque
0.1 μm(a)
FIGURE 7.18
Desmosomes. (a) Desmosomes anchor adjacent cells to each
other. (b) Cadherin proteins create the adhering link between
adjoining cells.
(b)
Cadherin and Actin Filaments
Cadherins can also connect the actin frame-
works of cells in cadherin-mediated junc-
tions (figure 7.19). When they do, they form
less stable links between cells than when
they connect intermediate filaments. Many
kinds of actin-linking cadherins occur in dif-
ferent tissues, as well as in the same tissue at
different times. During vertebrate develop-
ment, the migration of neurons in the em-
bryo is associated with changes in the type of
cadherin expressed on their plasma mem-
branes. This suggests that gene-controlled
changes in cadherin expression may provide
the migrating cells with a “roadmap” to their
destination.
Integrin-Mediated Links
Anchoring junctions called adherens junc-
tions are another type of junction that con-
nects the actin filaments of one cell with
those of neighboring cells or with the extra-
cellular matrix (figure 7.20). The linking
proteins in these junctions are members of a
large superfamily of cell surface receptors
called integrins. Each integrin is a trans-
membrane protein composed of two differ-
ent glycoprotein subunits that extend out-
ward from the plasma membrane. Together,
these subunits bind a protein component of
the extracellular matrix, like two hands
clasping a pole. There appear to be many
different kinds of integrin (cell biologists
have identified 20), each with a slightly dif-
ferent shaped “hand.” The exact component
of the matrix that a given cell binds to de-
pends on which combination of integrins
that cell has in its plasma membrane.
Anchoring junctions attach the
cytoskeleton of a cell to the matrix
surrounding the cell, or to the
cytoskeleton of another cell.
138 Part II Biology of the Cell
β
β
α γ
x
Extracellular
domains of
cadherin protein
Adjoining cell
membrane
Plasma
membrane
Cytoplasm
Cytoplasm
Cadherin of
adjoining cell
Actin
10 nm
COOH
Intracellular
attachment proteins
NH
2
FIGURE 7.19
A cadherin-mediated junction. The cadherin molecule is anchored to actin in the
cytoskeleton and passes through the membrane to interact with the cadherin of an
adjoining cell.
Extracellular
matrix protein
Plasma
membrane
Cytoplasm
Integrin
subunitIntegrin
subunit
Actin
10 nm
COOHHOOC
S
S
FIGURE 7.20
An integrin-mediated junction. These
adherens junctions link the actin filaments
inside cells to their neighbors and to the
extracellular matrix.
Communicating Junctions
Many cells communicate with adjacent cells through direct
connections, called communicating junctions. In these
junctions, a chemical signal passes directly from one cell to
an adjacent one. Communicating junctions establish direct
physical connections that link the cytoplasms of two cells
together, permitting small molecules or ions to pass from
one to the other. In animals, these direct communication
channels between cells are called gap junctions. In plants,
they are called plasmodesmata.
Gap Junctions in Animals
Communicating junctions called gap junctions are com-
posed of structures called connexons, complexes of six
identical transmembrane proteins (figure 7.21). The pro-
teins in a connexon are arranged in a circle to create a
channel through the plasma membrane that protrudes sev-
eral nanometers from the cell surface. A gap junction forms
when the connexons of two cells align perfectly, creating an
open channel spanning the plasma membranes of both
cells. Gap junctions provide passageways large enough to
permit small substances, such as simple sugars and amino
acids, to pass from the cytoplasm of one cell to that of the
next, yet small enough to prevent the passage of larger
molecules such as proteins. The connexons hold the plasma
membranes of the paired cells about 4 nanometers apart, in
marked contrast to the more-or-less direct contact between
the lipid bilayers in a tight junction.
Gap junction channels are dynamic structures that can
open or close in response to a variety of factors, including
Ca
++
and H
+
ions. This gating serves at least one important
function. When a cell is damaged, its plasma membrane
often becomes leaky. Ions in high concentrations outside
the cell, such as Ca
++
, flow into the damaged cell and shut
its gap junction channels. This isolates the cell and so pre-
vents the damage from spreading to other cells.
Plasmodesmata in Plants
In plants, cell walls separate every cell from all others. Cell-
cell junctions occur only at holes or gaps in the walls,
where the plasma membranes of adjacent cells can come
into contact with each other. Cytoplasmic connections that
form across the touching plasma membranes are called
plasmodesmata (figure 7.22). The majority of living cells
within a higher plant are connected with their neighbors by
these junctions. Plasmodesmata function much like gap
junctions in animal cells, although their structure is more
complex. Unlike gap junctions, plasmodesmata are lined
with plasma membrane and contain a central tubule that
connects the endoplasmic reticulum of the two cells.
Communicating junctions permit the controlled
passage of small molecules or ions between cells.
Chapter 7 Cell-Cell Interactions 139
Two adjacent connexons
forming an open channel
between cells
Adjacent plasma
membranes
Connexon
Channel
(diameter 1.5 nm)
Intercellular
space
"Gap" of
2-4 nm
FIGURE 7.21
Gap junctions. Connexons in gap junctions create passageways
that connect the cytoplasms of adjoining cells. Gap junctions
readily allow the passage of small molecules and ions required for
rapid communication (such as in heart tissue), but do not allow
the passage of larger molecules like proteins.
Vacuoles
Cytoplasm
Primary cell wall
Middle lamella
Plasmodesmata
Nuclei
FIGURE 7.22
Plasmodesmata. Plant cells can communicate through specialized
openings in their cell walls, called plasmodesmata, where the
cytoplasms of adjoining cells are connected.
140 Part II Biology of the Cell
Chapter 7
Summary Questions Media Resources
7.1 Cells signal one another with chemicals.
? Cell signaling is accomplished through the
recognition of signal molecules by target cells.
1. What determines which signal
molecules in the extracellular
environment a cell will respond
to?
2. How do paracrine, endocrine,
and synaptic signaling differ?
? The binding of a signal molecule to an intracellular
receptor usually initiates transcription of specific
regions of DNA, ultimately resulting in the
production of specific proteins.
? Cell surface receptors bind to specific molecules in
the extracellular fluid. In some cases, this binding
causes the receptor to enzymatically alter other
(usually internal) proteins, typically through
phosphorylation.
? G proteins behave as intracellular shuttles, moving
from an activated receptor to other areas in the cell.
3. Describe two of the ways in
which intracellular receptors
control cell activities.
4. What structural features are
characteristic of chemically
gated ion channels, and how are
these features related to the
function of the channels?
5. What are G proteins? How
do they participate in cellular
responses mediated by G-
protein-linked receptors?
7.2 Proteins in the cell and on its surface receive signals from other cells.
? There are usually several amplifying steps between
the binding of a signal molecule to a cell surface
receptor and the response of the cell. These steps
often involve phosphorylation by protein kinases.
6. How does the binding of a
single signal molecule to a cell
surface receptor result in an
amplified response within the
target cell?
7.3 Follow the journey of information into the cell.
? Tight junctions and desmosomes enable cells to
adhere in tight, leakproof sheets, holding the cells
together such that materials cannot pass between
them.
? Gap junctions (in animals) and plasmodesmata (in
plants) permit small substances to pass directly from
cell to cell through special passageways.
7. What are the functions of
tight junctions? What are the
functions of desmosomes and
adherens junctions, and what
proteins are involved in these
junctions?
8. What are the molecular
components that make up gap
junctions? What sorts of
substances can pass through gap
junctions?
9. Where are plasmodesmata
found? What cellular
constituents are found in
plasmodesmata?
7.4 Cell surface proteins mediate cell-cell interactions.
http://www.mhhe.com/raven6e http://www.biocourse.com
? Cell Interactions
? Student Research:
Retrograde
Messengers between
Nerve Cells
? Student Research:
Vertebrate Limb
formation
? Exploration: Cell-Cell
Interactions
? Scientists on Science:
G Proteins
141
How Do Proteins Help Chlorophyll
Carry Out Photosynthesis?
Much public attention in recent years has been focused on
high-profile science—headline-creating advances in the
Human Genome Project, genetic engineering, and the battle
against AIDS and cancer. Meanwhile, great advances have
been made more quietly in other areas of biology. Among the
greatest of these achievements has been the unmasking in the
last decade of the underlying mechanism of photosynthesis.
In photosynthesis, photons of light are absorbed by
chlorophyll molecules, causing them to donate a high-
energy electron that is put to work making NADPH and
pumping protons to produce ATP.
When researchers looked at the light-absorbing chloro-
phylls that carry out photosynthesis more closely, they
found the chlorophylls to be arranged in clusters called
photosystems, supported by proteins and accessory pig-
ments. Within a photosystem, hundreds of chlorophyll
molecules act like antennae, absorbing light and passing the
energy they capture inward to a single chlorophyll mole-
cule that acts as the reaction center. This chlorophyll acts
as the primary electron donor of photosynthesis. Once it
releases a light-energized electron, the complex series of
chemical events we call photosynthesis begins, and, like a
falling row of dominos, is difficult to stop.
Plants possess two kinds of photosystems that work
together to harvest light energy. One of them, called photo-
system I, is similar to a simpler photosystem found in pho-
tosynthetic bacteria, and is thought to have evolved from it.
Photosystem I has been the subject of intense research.
In its reaction center, a pair of chlorophyll molecules act as
the trap for photon energy, passing an excited electron on
to an acceptor molecule outside the reaction center. This
moves the photon energy away from the chlorophylls, and
is the key conversion of light energy to chemical energy,
the very heart of photosynthesis.
Because the pair of chlorophyll molecules in the reaction
center of photosystem I absorb light at a wavelength of
700 nm, they are together given the name P
700
. The P
700
dimer is positioned within the photosystem by two related
proteins that act as scaffolds. These proteins, discovered less
than 10 years ago, turn out to play a pivotal role in the pho-
tosynthetic process. Passing back and forth across the inter-
nal chloroplast membranes 11 times, they form a molecular
frame that positions P
700
to accept energy from other
chlorophyll molecules of the photosystem, and to donate a
photo-excited electron to an acceptor molecule outside the
photosystem.
Recent research suggests that the role of these scaffold
proteins, called PsaA and PsaB, is far more active than the
passive support provided by a scaffold. Analysis of highly
purified photosystems carried out in 1995 revealed that
the distribution of electric charge over the two halves of
the P
700
dimer is highly asymmetric—one chlorophyll
molecule exhibits a far greater charge density than the
other. Because the two chlorophyll molecules of P
700
are
themselves identical, this suggests that the PsaA and PsaB
proteins are actively modulating the physicochemical
properties of the chlorophyll.
How can a protein pull off this physical-chemical
sleight-of-hand? Just what are these proteins doing to the
chlorophyll molecules? To look more closely at what is
going on, you have to first figure out what part of the pro-
tein to look at. One way to get a handle on this problem is
to compare the amino acid sequences of PsaA and PsaB
with that of the bacterial photosystem from which they are
thought to have evolved. It is likely that such an important
part of the sequence would have been conserved and will
be found in all three.
Several sequences are indeed conserved, but most of
them prove not to interact directly with chlorophyll. One,
however, is a promising candidate. A single amino acid in
the helix X domain (that is, the tenth pass of the PsaB
protein across the membrane), dubbed His-656, is con-
served in all sequences, and is positioned right where the
PsaB protein touches the P
700
chlorophyll (see above).
This amino acid, a histidine, has become the focus of
recent efforts to clarify how proteins help chlorophyll
carry out photosynthesis.
Part
Stroma
Thylakoid
space
XI
700
Psa B protein
650
736
600
450
500
IX
VIII
VII
398
X
550
H P
700
III
Energetics
The proposed antenna complex of the PsaB protein. Position
656 is a histidine (H) in the tenth pass (helix X) of the PsaB
protein across the thylakoid membrane within chloroplasts. This
histidine is where the PsaB protein makes contact with a P
700
chlorophyll molecule.
Real People Doing Real Science
The Experiment
To determine the importance of His-656, and more gener-
ally of the helix X domain of the PsaB protein, Professor
Andrew Webber of Arizona State University, working with
his research team and the group of Professor Wolfgang
Lubitz at Technische Universitat Berlin, has created site-
directed mutations of His-656 in the photosynthetic protist
Chlamydomonas reinhardtii. C. reinhardtii is widely used to
study photosynthesis because of the ease with which lab ex-
periments can be done.
Webber and his collaborators set out to change the
amino acid located at position 656 of PsaB, and then to
look and see what effect the change had on photosynthesis.
If His-656 indeed plays a critical role in modifying the P
700
chlorophylls, then a change at that position to a different
amino acid should have profound effects.
Creating PsaB Proteins Mutant at Position 656. The
first and key step in Webber’s experimental approach was
to genetically alter the chloroplast of C. reinhardtii, intro-
ducing a mutation of the PsaB gene at the His-656 posi-
tion. To do this, the team employed site-specific mutagen-
esis to construct mutant plasmids pHN(B656) and
pHS(B656), inserting a gene carrying either the Ser or Asn
amino acids in place of His. The two mutation-carrying
plasmids were then cloned into C. reinhardtii, cells carrying
the mutant plasmids isolated, and presence of the mutated
gene directly confirmed by sequencing the DNA.
Characterizing the Effects of 656 Mutations. Once re-
searchers confirmed that the C. reinhardtii chloroplast
DNA now contained the mutant forms of the PsaB gene,
they proceeded to test the function of the mutated PsaB
protein in coordinating P
700
, examining interior thylakoid
membranes isolated from the chloroplasts. To do this, the
researchers measured the oxidation midpoint potentials of
the P
700
complexes, an indication of how tightly the chloro-
phyll molecules are holding onto their electrons.
The researchers further characterized the P
700
com-
plexes by measuring the changes in absorbance of the mu-
tants versus the wild type to see if the mutations altered the
spectral properties of the P
700
chlorophylls.
The Results
The results of the examination of the oxidation midpoint
potentials revealed that the influence of the PsaB protein
on P
700
had been profoundly altered by the mutations.
The midpoint potential of P
700
in the wild type was deter-
mined to be 447+6 mV, while the midpoint potential had
increased to 487+6 mV in both the PsaB mutant I,
HN(B656), and the PsaB mutant II, HS(B656) (see graph
a). This increase in the oxidation midpoint potential by ap-
proximately 40 mV indicates that the mutations to the His
residue significantly altered the redox property of P
700
and,
therefore, that His-656 is closely interacting with one of
the chlorophyll molecules of the P
700
dimer.
These results and this conclusion are further supported by
changes observed in the spectral properties of the mutants
and wild type (see graph b). There is a reduction and a slight
shift in the 696 nm bleaching band (dip in absorbance) in
PsaB mutant I toward the blue end of the spectrum and a
new bleaching band appearing at 667 nm, both changes in
the spectral properties of chlorophyll induced by the muta-
tional changes in the PsaB protein.
Ultimately, the researchers conclude that the His-656 of
PsaB directly coordinates the central magnesium atom of
one of the two chlorophyll molecules of P
700
. Their results
are consistent with a model of photosystem I in which the
first six spans of PsaB constitute an antenna domain for
receiving energy from other chlorophylls and the last five
membrane spans interact with the P
700
reaction complex.
Relative absorbance dif
ference at 826 nm
Relative absorbance change
0
H110021
(b)(a)
Wild type
Mutant I
Wild type
Mutant I
Mutant II
1.0
0.8
0.6
0.4
0.2
0.0
H110020.2
0.25 0.30 0.35 0.40 0.45
Potential (volts) Wavelength (nm)
0.50 0.55 0.60 0.65 500 550 600 650 700
Effect of altering position 656. (a) When P
700
interacts with normal and mutant forms of PsaB, the midpoint potentials are 447+6 mV in
the wild type and 487+6 in the mutants, the mutant value being about 40 mV higher. (b) The bleaching band (dip in the absorbance) of
P
700
is shifted to the blue (left) and exhibits a new bleaching band at 667 nm when interacting with mutant forms of PsaB.
To explore this experiment further,
go to the Virtual Lab at
www.mhhe.com/raven6/vlab3.mhtml
143
8
Energy and Metabolism
Concept Outline
8.1 The laws of thermodynamics describe how energy
changes.
The Flow of Energy in Living Things. Potential energy
is present in the valance electrons of atoms, and so can be
transferred from one molecule to another.
The Laws of Thermodynamics. Energy is never lost but
as it is transferred, more and more of it dissipates as heat, a
disordered form of energy.
Free Energy. In a chemical reaction, the energy released
or supplied is the difference in bond energies between
reactants and products, corrected for disorder.
Activation Energy. To start a chemical reaction, an input
of energy is required to destabilize existing chemical bonds.
8.2 Enzymes are biological catalysts.
Enzymes. Globular proteins called enzymes catalyze
chemical reactions within cells.
How Enzymes Work. Enzymes have sites on their surface
shaped to fit their substrates snugly, forcing chemically
reactive groups close enough to facilitate a reaction.
Enzymes Take Many Forms. Some enzymes are
associated in complex groups; others are not even proteins.
Factors Affecting Enzyme Activity. Each enzyme works
most efficiently at its optimal temperature and pH. Metal
ions or other substances often help enzymes carry out their
catalysis.
8.3 ATP is the energy currency of life.
What Is ATP? Cells store and release energy from the
phosphate bonds of ATP, the energy currency of the cell.
8.4 Metabolism is the chemical life of a cell.
Biochemical Pathways: The Organizational Units of
Metabolism. Biochemical pathways are the organizational
units of metabolism.
The Evolution of Metabolism. The major metabolic
processes evolved over a long period, building on what came
before.
L
ife can be viewed as a constant flow of energy, channeled
by organisms to do the work of living. Each of the signif-
icant properties by which we define life—order, growth, re-
production, responsiveness, and internal regulation—
requires a constant supply of energy (figure 8.1). Deprived of
a source of energy, life stops. Therefore, a comprehensive
study of life would be impossible without discussing bioener-
getics, the analysis of how energy powers the activities of liv-
ing systems. In this chapter, we will focus on energy—on
what it is and how organisms capture, store, and use it.
FIGURE 8.1
Lion at lunch. Energy that this lion extracts from its meal of
giraffe will be used to power its roar, fuel its running, and build a
bigger lion.
Oxidation-Reduction
Energy flows into the biological world from the sun, which
shines a constant beam of light on the earth. It is estimated
that the sun provides the earth with more than 13 × 10
23
calories per year, or 40 million billion calories per second!
Plants, algae, and certain kinds of bacteria capture a frac-
tion of this energy through photosynthesis. In photosyn-
thesis, energy garnered from sunlight is used to combine
small molecules (water and carbon dioxide) into more com-
plex molecules (sugars). The energy is stored as potential
energy in the covalent bonds between atoms in the sugar
molecules. Recall from chapter 2 that an atom consists of a
central nucleus surrounded by one or more orbiting elec-
trons, and a covalent bond forms when two atomic nuclei
share valence electrons. Breaking such a bond requires en-
ergy to pull the nuclei apart. Indeed, the strength of a cova-
lent bond is measured by the amount of energy required to
break it. For example, it takes 98.8 kcal to break one mole
(6.023 × 10
23
) of carbon-hydrogen (C—H) bonds.
During a chemical reaction, the energy stored in
chemical bonds may transfer to new bonds. In some of
these reactions, electrons actually pass from one atom or
molecule to another. When an atom or molecule loses an
electron, it is said to be oxidized, and the process by
which this occurs is called oxidation. The name reflects
the fact that in biological systems oxygen, which attracts
electrons strongly, is the most common electron accep-
144 Part III Energetics
The Flow of Energy in Living Things
Energy is defined as the capacity to do work. It can be con-
sidered to exist in two states. Kinetic energy is the energy
of motion (figure 8.2). Moving objects perform work by
causing other matter to move. Potential energy is stored
energy. Objects that are not actively moving but have the
capacity to do so possess potential energy. A boulder
perched on a hilltop has potential energy; as it begins to
roll downhill, some of its potential energy is converted into
kinetic energy. Much of the work that living organisms
carry out involves transforming potential energy to kinetic
energy.
Energy can take many forms: mechanical energy, heat,
sound, electric current, light, or radioactive radiation. Be-
cause it can exist in so many forms, there are many ways to
measure energy. The most convenient is in terms of heat,
because all other forms of energy can be converted into
heat. In fact, the study of energy is called thermodynam-
ics, meaning heat changes. The unit of heat most com-
monly employed in biology is the kilocalorie (kcal). One
kilocalorie is equal to 1000 calories (cal), and one calorie is
the heat required to raise the temperature of one gram of
water one degree Celsius (°C). (It is important not to con-
fuse calories with a term related to diets and nutrition, the
Calorie with a capital C, which is actually another term for
kilocalorie.) Another energy unit, often used in physics, is
the joule; one joule equals 0.239 cal.
8.1 The laws of thermodynamics describe how energy changes.
(a) Potential energy (b) Kinetic energy
FIGURE 8.2
Potential and kinetic energy. (a) Objects that have the capacity to move but are not moving have potential energy. The energy required
to move the ball up the hill is stored as potential energy. (b) Objects that are in motion have kinetic energy. The stored energy is released as
kinetic energy as the ball rolls down the hill.
tor. Conversely, when an atom or molecule gains an elec-
tron, it is said to be reduced, and the process is called re-
duction. Oxidation and reduction always take place to-
gether, because every electron that is lost by an atom
through oxidation is gained by some other atom through
reduction. Therefore, chemical reactions of this sort are
called oxidation-reduction (redox) reactions (figure
8.3). Energy is transferred from one molecule to another
via redox reactions. The reduced form of a molecule thus
has a higher level of energy than the oxidized form
(figure 8.4).
Oxidation-reduction reactions play a key role in the flow
of energy through biological systems because the electrons
that pass from one atom to another carry energy with
them. The amount of energy an electron possesses depends
on how far it is from the nucleus and how strongly the nu-
cleus attracts it. Light (and other forms of energy) can add
energy to an electron and boost it to a higher energy level.
When this electron departs from one atom (oxidation) and
moves to another (reduction), the electron’s added energy
is transferred with it, and the electron orbits the second
atom’s nucleus at the higher energy level. The added en-
ergy is stored as potential chemical energy that the atom
can later release when the electron returns to its original
energy level.
Energy is the capacity to do work, either actively
(kinetic energy) or stored for later use (potential
energy). Energy is transferred with electrons. Oxidation
is the loss of an electron; reduction is the gain of one.
Chapter 8 Energy and Metabolism 145
Product
NAD
+
H
H
H
H
NAD
+
NAD
+
NAD
NAD
H
Energy-rich molecule
1. Enzymes that harvest hydrogen
atoms have a binding site for NAD
+
located near the substrate binding
site. NAD
+
and an energy-rich
molecule bind to the enzyme.
3. NADH then diffuses away and
is available to other molecules.
2. In an oxidation-reduction reaction,
a hydrogen atom is transferred to
NAD
+
, forming NADH.
FIGURE 8.3
An oxidation-reduction reaction. Cells use a chemical called NAD
+
to carry out oxidation-reduction reactions. Energetic electrons are
often paired with a proton as a hydrogen atom. Molecules that gain energetic electrons are said to be reduced, while ones that lose
energetic electrons are said to be oxidized. NAD
+
oxidizes energy-rich molecules by acquiring their hydrogens (in the figure, this proceeds
1→2→3) and then reduces other molecules by giving the hydrogens to them (in the figure, this proceeds 3→2→1).
FIGURE 8.4
Redox reactions. Oxidation is the loss of an electron; reduction is
the gain of an electron. In this example, the charges of molecules
A and B are shown in small circles to the upper right of each
molecule. Molecule A loses energy as it loses an electron, while
molecule B gains energy as it gains an electron.
Gain of electron (reduction)
Low energy
e
–
AB
High energy
Loss of electron (oxidation)
A+
oo
B +
+ –
A* B*
The Laws of
Thermodynamics
Running, thinking, singing, reading these
words—all activities of living organisms
involve changes in energy. A set of univer-
sal laws we call the Laws of Thermody-
namics govern all energy changes in the
universe, from nuclear reactions to the
buzzing of a bee.
The First Law of Thermodynamics
The first of these universal laws, the
First Law of Thermodynamics, con-
cerns the amount of energy in the uni-
verse. It states that energy cannot be cre-
ated or destroyed; it can only change
from one form to another (from potential
to kinetic, for example). The total
amount of energy in the universe remains
constant.
The lion eating a giraffe in figure 8.1
is in the process of acquiring energy.
Rather than creating new energy or cap-
turing the energy in sunlight, the lion is
merely transferring some of the potential
energy stored in the giraffe’s tissues to its
own body (just as the giraffe obtained the
potential energy stored in the plants it
ate while it was alive). Within any living
organism, this chemical potential energy can be shifted to
other molecules and stored in different chemical bonds,
or it can convert into other forms, such as kinetic energy,
light, or electricity. During each conversion, some of the
energy dissipates into the environment as heat, a measure
of the random motions of molecules (and, hence, a mea-
sure of one form of kinetic energy). Energy continuously
flows through the biological world in one direction, with
new energy from the sun constantly entering the system
to replace the energy dissipated as heat.
Heat can be harnessed to do work only when there is a
heat gradient, that is, a temperature difference between two
areas (this is how a steam engine functions). Cells are too
small to maintain significant internal temperature differ-
ences, so heat energy is incapable of doing the work of
cells. Thus, although the total amount of energy in the uni-
verse remains constant, the energy available to do work de-
creases, as progressively more of it dissipates as heat.
The Second Law of Thermodynamics
The Second Law of Thermodynamics concerns this trans-
formation of potential energy into heat, or random molecular
motion. It states that the disorder (more formally called en-
tropy) in the universe is continuously increasing. Put simply,
disorder is more likely than order. For example, it is much
more likely that a column of bricks will tumble over than that
a pile of bricks will arrange themselves spontaneously to form
a column. In general, energy transformations proceed sponta-
neously to convert matter from a more ordered, less stable
form, to a less ordered, more stable form (figure 8.5).
Entropy
Entropy is a measure of the disorder of a system, so the
Second Law of Thermodynamics can also be stated simply
as “entropy increases.” When the universe formed, it held
all the potential energy it will ever have. It has become pro-
gressively more disordered ever since, with every energy
exchange increasing the amount of entropy.
The First Law of Thermodynamics states that energy
cannot be created or destroyed; it can only undergo
conversion from one form to another. The Second Law
of Thermodynamics states that disorder (entropy) in
the universe is increasing. As energy is used, more and
more of it is converted to heat, the energy of random
molecular motion.
146 Part III Energetics
Disorder happens “spontaneously”
Organization requires energy
FIGURE 8.5
Entropy in action. As time elapses, a child’s room becomes more disorganized. It takes
effort to clean it up.
Free Energy
It takes energy to break the chemical bonds that hold the
atoms in a molecule together. Heat energy, because it in-
creases atomic motion, makes it easier for the atoms to pull
apart. Both chemical bonding and heat have a significant
influence on a molecule, the former reducing disorder and
the latter increasing it. The net effect, the amount of en-
ergy actually available to break and subsequently form
other chemical bonds, is called the free energy of that
molecule. In a more general sense, free energy is defined as
the energy available to do work in any system. In a mole-
cule within a cell, where pressure and volume usually do
not change, the free energy is denoted by the symbol G
(for “Gibbs’ free energy,” which limits the system being
considered to the cell). G is equal to the energy contained
in a molecule’s chemical bonds (called enthalpy and desig-
nated H) minus the energy unavailable because of disorder
(called entropy and given the symbol S) times the absolute
temperature, T, in degrees Kelvin (K = °C + 273):
G = H – TS
Chemical reactions break some bonds in the reactants
and form new bonds in the products. Consequently, reac-
tions can produce changes in free energy. When a chemical
reaction occurs under conditions of constant temperature,
pressure, and volume—as do most biological reactions—
the change in free energy (?G) is simply:
?G = ?H – T ?S
The change in free energy, or ?G, is a fundamental
property of chemical reactions.
In some reactions, the ?G is positive. This means that
the products of the reaction contain more free energy than
the reactants; the bond energy (H) is higher or the disorder
(S) in the system is lower. Such reactions do not proceed
spontaneously because they require an input of energy. Any
reaction that requires an input of energy is said to be en-
dergonic (“inward energy”).
In other reactions, the ?G is negative. The products of
the reaction contain less free energy than the reactants; ei-
ther the bond energy is lower or the disorder is higher, or
both. Such reactions tend to proceed spontaneously. Any
chemical reaction will tend to proceed spontaneously if the
difference in disorder (T ?S) is greater than the difference
in bond energies between reactants and products (?H).
Note that spontaneous does not mean the same thing as in-
stantaneous. A spontaneous reaction may proceed very
slowly. These reactions release the excess free energy as
heat and are thus said to be exergonic (“outward energy”).
Figure 8.6 sums up these reactions.
Free energy is the energy available to do work. Within
cells, the change in free energy (?G) is the difference
in bond energies between reactants and products (?H),
minus any change in the degree of disorder of the
system (T ?S). Any reaction whose products contain
less free energy than the reactants (?G is negative) will
tend to proceed spontaneously.
Chapter 8 Energy and Metabolism 147
Reactant Reactant
Product
Product
ExergonicEndergonic
Energy released
Energy supplied
Energy is
released.
Energy
must be
supplied.
FIGURE 8.6
Energy in chemical reactions. (a) In an endergonic reaction, the products of the reaction contain more energy than the reactants, and the
extra energy must be supplied for the reaction to proceed. (b) In an exergonic reaction, the products contain less energy than the reactants,
and the excess energy is released.
(a) (b)
Activation Energy
If all chemical reactions that release free energy tend to
occur spontaneously, why haven’t all such reactions already
occurred? One reason they haven’t is that most reactions
require an input of energy to get started. Before it is possi-
ble to form new chemical bonds, even bonds that contain
less energy, it is first necessary to break the existing bonds,
and that takes energy. The extra energy required to desta-
bilize existing chemical bonds and initiate a chemical reac-
tion is called activation energy (figure 8.7a).
The rate of an exergonic reaction depends on the acti-
vation energy required for the reaction to begin. Reac-
tions with larger activation energies tend to proceed
more slowly because fewer molecules succeed in over-
coming the initial energy hurdle. Activation energies are
not constant, however. Stressing particular chemical
bonds can make them easier to break. The process of in-
fluencing chemical bonds in a way that lowers the activa-
tion energy needed to initiate a reaction is called cataly-
sis, and substances that accomplish this are known as
catalysts (figure 8.7b).
Catalysts cannot violate the basic laws of thermody-
namics; they cannot, for example, make an endergonic re-
action proceed spontaneously. By reducing the activation
energy, a catalyst accelerates both the forward and the re-
verse reactions by exactly the same amount. Hence, it
does not alter the proportion of reactant ultimately con-
verted into product.
148 Part III Energetics
To grasp this, imagine a bowling ball resting in a shal-
low depression on the side of a hill. Only a narrow rim of
dirt below the ball prevents it from rolling down the hill.
Now imagine digging away that rim of dirt. If you remove
enough dirt from below the ball, it will start to roll down
the hill—but removing dirt from below the ball will never
cause the ball to roll UP the hill! Removing the lip of dirt
simply allows the ball to move freely; gravity determines
the direction it then travels. Lowering the resistance to the
ball’s movement will promote the movement dictated by its
position on the hill.
Similarly, the direction in which a chemical reaction
proceeds is determined solely by the difference in free en-
ergy. Like digging away the soil below the bowling ball on
the hill, catalysts reduce the energy barrier preventing the
reaction from proceeding. Catalysts don’t favor endergonic
reactions any more than digging makes the hypothetical
bowling ball roll uphill. Only exergonic reactions can pro-
ceed spontaneously, and catalysts cannot change that.
What catalysts can do is make a reaction proceed much
faster.
The rate of a reaction depends on the activation energy
necessary to initiate it. Catalysts reduce the activation
energy and so increase the rates of reactions, although
they do not change the final proportions of reactants
and products.
Activation
energy
Reactant
Product
Activation
energy
Catalyzed
Uncatalyzed
Product
Energy released
Energy supplied
Reactant
FIGURE 8.7
Activation energy and catalysis. (a) Exergonic reactions do not necessarily proceed rapidly because energy must be supplied to destabilize
existing chemical bonds. This extra energy is the activation energy for the reaction. (b) Catalysts accelerate particular reactions by lowering
the amount of activation energy required to initiate the reaction.
(a) (b)
Enzymes
The chemical reactions within living organisms are regu-
lated by controlling the points at which catalysis takes
place. Life itself is, therefore, regulated by catalysts. The
agents that carry out most of the catalysis in living organ-
isms are proteins called enzymes. (There is increasing evi-
dence that some types of biological catalysis are carried out
by RNA molecules.) The unique three-dimensional shape
of an enzyme enables it to stabilize a temporary association
between substrates, the molecules that will undergo the
reaction. By bringing two substrates together in the correct
orientation, or by stressing particular chemical bonds of a
substrate, an enzyme lowers the activation energy required
for new bonds to form. The reaction thus proceeds much
more quickly than it would without the enzyme. Because
the enzyme itself is not changed or consumed in the reac-
tion, only a small amount of an enzyme is needed, and it
can be used over and over.
As an example of how an enzyme works, let’s consider
the reaction of carbon dioxide and water to form carbonic
acid. This important enzyme-catalyzed reaction occurs in
vertebrate red blood cells:
CO
2
+ H
2
O ?→ H
2
CO
3
carbon water carbonic
dioxide acid
This reaction may proceed in either direction, but because
it has a large activation energy, the reaction is very slow in
the absence of an enzyme: perhaps 200 molecules of car-
bonic acid form in an hour in a cell. Reactions that proceed
this slowly are of little use to a cell. Cells overcome this
problem by employing an enzyme within their cytoplasm
called carbonic anhydrase (enzyme names usually end in
“–ase”). Under the same conditions, but in the presence of
carbonic anhydrase, an estimated 600,000 molecules of car-
bonic acid form every second! Thus, the enzyme increases
the reaction rate more than 10 million times.
Thousands of different kinds of enzymes are known,
each catalyzing one or a few specific chemical reactions.
By facilitating particular chemical reactions, the enzymes
in a cell determine the course of metabolism—the collec-
tion of all chemical reactions—in that cell. Different types
of cells contain different sets of enzymes, and this differ-
ence contributes to structural and functional variations
among cell types. The chemical reactions taking place
within a red blood cell differ from those that occur within
a nerve cell, in part because the cytoplasm and membranes
of red blood cells and nerve cells contain different arrays
of enzymes.
Cells use proteins called enzymes as catalysts to lower
activation energies.
Chapter 8 Energy and Metabolism 149
8.2 Enzymes are biological catalysts.
Catalysis: A Closer
Look at Carbonic
Anhydrase
of the enzyme is a deep cleft traversing the
enzyme, as if it had been cut with the blade
of an ax. Deep within the cleft, some 1.5
nm from the surface, are located three his-
tidines, their imidazole (nitrogen ring)
groups all pointed at the same place in the
center of the cleft. Together they hold a
zinc ion firmly in position. This zinc ion
will be the cutting blade of the catalytic
process.
Here is how the zinc catalyzes the reac-
tion. Immediately adjacent to the position
of the zinc atom in the cleft are a group of
amino acids that recognize and bind carbon
dioxide. The zinc atom interacts with this
carbon dioxide molecule, orienting it in the
plane of the cleft. Meanwhile, water bound
to the zinc is rapidly converted to hydroxide
ion. This hydroxide ion is now precisely po-
sitioned to attack the carbon dioxide. When
it does so, HCO
3
–
is formed—and the en-
zyme is unchanged (figure 8.A).
Carbonic anhydrase is an effective cata-
lyst because it brings its two substrates into
close proximity and optimizes their orien-
tation for reaction. Other enzymes use
other mechanisms. Many, for example, use
charged amino acids to polarize substrates
or electronegative amino acids to stress
particular bonds. Whatever the details of
the reaction, however, the precise posi-
tioning of substrates achieved by the par-
ticular shape of the enzyme always plays a
key role.
One of the most rapidly acting enzymes in
the human body is carbonic anhydrase,
which plays a key role in blood by convert-
ing dissolved CO
2
into carbonic acid, which
dissociates into bicarbonate and hydrogen
ions:
CO
2
+ H
2
O → H
2
CO
3
→ HCO
3
–
+ H
+
Fully 70% of the CO
2
transported by
the blood is transported as bicarbonate ion.
This reaction is exergonic, but its energy of
activation is significant, so that little con-
version to bicarbonate occurs sponta-
neously. In the presence of the enzyme car-
bonic anhydrase, however, the rate of the
reaction accelerates by a factor of more
than 10 million!
How does carbonic anhydrase catalyze
this reaction so effectively? The active site
Zn
++
Zn
++
HIS
HIS
HIS
HIS
HIS
HIS
H
O
–
C
O
O
H
O
C
O
–
O
FIGURE 8.A
How Enzymes Work
Most enzymes are globular proteins with one or more
pockets or clefts on their surface called active sites (figure
8.8). Substrates bind to the enzyme at these active sites,
forming an enzyme-substrate complex. For catalysis to
occur within the complex, a substrate molecule must fit
precisely into an active site. When that happens, amino
acid side groups of the enzyme end up in close proximity to
certain bonds of the substrate. These side groups interact
chemically with the substrate, usually stressing or distorting
a particular bond and consequently lowering the activation
energy needed to break the bond. The substrate, now a
product, then dissociates from the enzyme.
Proteins are not rigid. The binding of a substrate in-
duces the enzyme to adjust its shape slightly, leading to a
better induced fit between enzyme and substrate (figure 8.9).
This interaction may also facilitate the binding of other
substrates; in such cases, the substrate itself “activates” the
enzyme to receive other substrates.
Enzymes typically catalyze only one or a few similar
chemical reactions because they are specific in their
choice of substrates. This specificity is due to the active
site of the enzyme, which is shaped so that only a
certain substrate molecule will fit into it.
150 Part III Energetics
Active site
(a)
Substrate
(b)
FIGURE 8.8
How the enzyme lysozyme works. (a) A groove runs through
lysozyme that fits the shape of the polysaccharide (a chain of
sugars) that makes up bacterial cell walls. (b) When such a chain of
sugars, indicated in yellow, slides into the groove, its entry
induces the protein to alter its shape slightly and embrace the
substrate more intimately. This induced fit positions a glutamic
acid residue in the protein next to the bond between two adjacent
sugars, and the glutamic acid “steals” an electron from the bond,
causing it to break.
The substrate, sucrose, consists
of glucose and fructose bonded
together.
1
The substrate binds to the enzyme,
forming an enzyme-substrate
complex.
2
The binding of the
substrate and enzyme
places stress on the
glucose-fructose bond,
and the bond breaks.
3
Products are released,
and the enzyme is free
to bind other substrates.
4
Bond
Enzyme
Active site
H
2
O
Glucose Fructose
FIGURE 8.9
The catalytic cycle of an enzyme. Enzymes increase the speed with which chemical reactions occur, but they are not altered
themselves as they do so. In the reaction illustrated here, the enzyme sucrase is splitting the sugar sucrose (present in most candy) into
two simpler sugars: glucose and fructose. (1) First, the sucrose substrate binds to the active site of the enzyme, fitting into a depression
in the enzyme surface. (2) The binding of sucrose to the active site forms an enzyme-substrate complex and induces the sucrase
molecule to alter its shape, fitting more tightly around the sucrose. (3) Amino acid residues in the active site, now in close proximity to
the bond between the glucose and fructose components of sucrose, break the bond. (4) The enzyme releases the resulting glucose and
fructose fragments, the products of the reaction, and is then ready to bind another molecule of sucrose and run through the catalytic
cycle once again. This cycle is often summarized by the equation: E + S ?
[
ES
]
? E + P, where E = enzyme, S = substrate, ES =
enzyme-substrate complex, and P = products.
Enzymes Take
Many Forms
While many enzymes are suspended
in the cytoplasm of cells, free to
move about and not attached to any
structure, other enzymes function as
integral parts of cell structures and
organelles.
Multienzyme Complexes
Often in cells the several enzymes
catalyzing the different steps of a
sequence of reactions are loosely
associated with one another in non-
covalently bonded assemblies called
multienzyme complexes. The bacter-
ial pyruvate dehydrogenase mul-
tienzyme complex seen in figure
8.10 contains enzymes that carry
out three sequential reactions in ox-
idative metabolism. Each complex has multiple copies of
each of the three enzymes—60 protein subunits in all.
The many subunits work in concert, like a tiny factory.
Multienzyme complexes offer significant advantages in
catalytic efficiency:
1. The rate of any enzyme reaction is limited by the fre-
quency with which the enzyme collides with its sub-
strate. If a series of sequential reactions occurs within
a multienzyme complex, the product of one reaction
can be delivered to the next enzyme without releasing
it to diffuse away.
2. Because the reacting substrate never leaves the com-
plex during its passage through the series of reactions,
the possibility of unwanted side reactions is eliminated.
3. All of the reactions that take place within the mul-
tienzyme complex can be controlled as a unit.
In addition to pyruvate dehydrogenase, which controls
entry to the Krebs cycle, several other key processes in
the cell are catalyzed by multienzyme complexes. One
well-studied system is the fatty acid synthetase complex
that catalyzes the synthesis of fatty acids from two-carbon
precursors. There are seven different enzymes in this
multienzyme complex, and the reaction intermediates re-
main associated with the complex for the entire series of
reactions.
Not All Biological Catalysts Are Proteins
Until a few years ago, most biology textbooks contained
statements such as “Enzymes are the catalysts of biological
systems.” We can no longer make that statement without
qualification. As discussed in chapter 4, Tom Cech and his
colleagues at the University of Colorado reported in 1981
that certain reactions involving RNA molecules appear to
be catalyzed in cells by RNA itself, rather than by enzymes.
This initial observation has been corroborated by addi-
tional examples of RNA catalysis in the last few years. Like
enzymes, these RNA catalysts, which are loosely called “ri-
bozymes,” greatly accelerate the rate of particular biochem-
ical reactions and show extraordinary specificity with re-
spect to the substrates on which they act.
There appear to be at least two sorts of ribozymes.
Those that carry out intramolecular catalysis have folded
structures and catalyze reactions on themselves. Those
that carry out intermolecular catalysis act on other mole-
cules without themselves being changed in the process.
Many important cellular reactions involve small RNA
molecules, including reactions that chip out unnecessary
sections from RNA copies of genes, that prepare ribo-
somes for protein synthesis, and that facilitate the replica-
tion of DNA within mitochondria. In all of these cases,
the possibility of RNA catalysis is being actively investi-
gated. It seems likely, particularly in the complex process
of photosynthesis, that both enzymes and RNA play im-
portant catalytic roles.
The ability of RNA, an informational molecule, to act as
a catalyst has stirred great excitement among biologists, as
it appears to provide a potential answer to the “chicken-
and-egg” riddle posed by the spontaneous origin of life hy-
pothesis discussed in chapter 3. Which came first, the pro-
tein or the nucleic acid? It now seems at least possible that
RNA may have evolved first and catalyzed the formation of
the first proteins.
Not all biological catalysts float free in the cytoplasm.
Some are part of other structures, and some are not
even proteins.
Chapter 8 Energy and Metabolism 151
(a)
FIGURE 8.10
The enzyme pyruvate dehydrogenase. The enzyme (model, a) that carries out the oxidation
of pyruvate is one of the most complex enzymes known—it has 60 subunits, many of which
can be seen in the electron micrograph (b) (200,000×).
Factors Affecting Enzyme Activity
The rate of an enzyme-catalyzed reaction is affected by the
concentration of substrate, and of the enzyme that works
on it. In addition, any chemical or physical factor that alters
the enzyme’s three-dimensional shape—such as tempera-
ture, pH, salt concentration, and the binding of specific
regulatory molecules—can affect the enzyme’s ability to
catalyze the reaction.
Temperature
Increasing the temperature of an uncatalyzed reaction will
increase its rate because the additional heat represents an
increase in random molecular movement. The rate of an
enzyme-catalyzed reaction also increases with temperature,
but only up to a point called the temperature optimum (fig-
ure 8.11a). Below this temperature, the hydrogen bonds
and hydrophobic interactions that determine the enzyme’s
shape are not flexible enough to permit the induced fit that
is optimum for catalysis. Above the temperature optimum,
these forces are too weak to maintain the enzyme’s shape
against the increased random movement of the atoms in
the enzyme. At these higher temperatures, the enzyme de-
natures, as we described in chapter 3. Most human enzymes
have temperature optima between 35°C and 40°C, a range
that includes normal body temperature. Bacteria that live in
hot springs have more stable enzymes (that is, enzymes
held together more strongly), so the temperature optima
for those enzymes can be 70°C or higher.
pH
Ionic interactions between oppositely charged amino acid
residues, such as glutamic acid (–) and lysine (+), also hold
enzymes together. These interactions are sensitive to the
hydrogen ion concentration of the fluid the enzyme is dis-
solved in, because changing that concentration shifts the
balance between positively and negatively charged amino
acid residues. For this reason, most enzymes have a pH op-
timum that usually ranges from pH 6 to 8. Those enzymes
able to function in very acid environments are proteins that
maintain their three-dimensional shape even in the pres-
ence of high levels of hydrogen ion. The enzyme pepsin,
for example, digests proteins in the stomach at pH 2, a very
acidic level (figure 8.11b).
Inhibitors and Activators
Enzyme activity is sensitive to the presence of specific sub-
stances that bind to the enzyme and cause changes in its
shape. Through these substances, a cell is able to regulate
which enzymes are active and which are inactive at a partic-
ular time. This allows the cell to increase its efficiency and
to control changes in its characteristics during develop-
ment. A substance that binds to an enzyme and decreases its
activity is called an inhibitor. Very often, the end product
of a biochemical pathway acts as an inhibitor of an early re-
action in the pathway, a process called feedback inhibition (to
be discussed later).
Enzyme inhibition occurs in two ways: competitive
inhibitors compete with the substrate for the same bind-
ing site, displacing a percentage of substrate molecules
from the enzymes; noncompetitive inhibitors bind to
the enzyme in a location other than the active site,
changing the shape of the enzyme and making it unable
to bind to the substrate (figure 8.12). Most noncompeti-
tive inhibitors bind to a specific portion of the enzyme
called an allosteric site (Greek allos, “other” + steros,
“form”). These sites serve as chemical on/off switches;
the binding of a substance to the site can switch the en-
zyme between its active and inactive configurations. A
substance that binds to an allosteric site and reduces en-
zyme activity is called an allosteric inhibitor (figure
8.12b). Alternatively, activators bind to allosteric sites
and keep the enzymes in their active configurations,
thereby increasing enzyme activity.
152 Part III Energetics
30
Optimum temperature
for human enzyme
Optimal temperature
for enzyme from
hotsprings bacterium
Optimum pH
for pepsin
Rate of reaction
Temperature of reaction (°C)
pH of reaction
Rate of reaction
Optimum pH
for trypsin
40 50 60 70 80
1 2 3 4 5 6 7 8 9
(a)
(b)
FIGURE 8.11
Enzymes are sensitive to their environment. The activity of
an enzyme is influenced by both (a) temperature and (b) pH.
Most human enzymes, such as the protein-degrading enzyme
trypsin, work best at temperatures of about 40°C and within a
pH range of 6 to 8.
Enzyme Cofactors
Enzyme function is often assisted by additional chemical
components known as cofactors. For example, the active
sites of many enzymes contain metal ions that help draw
electrons away from substrate molecules. The enzyme
carboxypeptidase digests proteins by employing a zinc ion
(Zn
++
) in its active site to remove electrons from the
bonds joining amino acids. Other elements, such as
molybdenum and manganese, are also used as cofactors.
Like zinc, these substances are required in the diet in
small amounts. When the cofactor is a nonprotein organic
molecule, it is called a coenzyme. Many vitamins are
parts of coenzymes.
In numerous oxidation-reduction reactions that are cat-
alyzed by enzymes, the electrons pass in pairs from the ac-
tive site of the enzyme to a coenzyme that serves as the
electron acceptor. The coenzyme then transfers the elec-
trons to a different enzyme, which releases them (and the
energy they bear) to the substrates in another reaction.
Often, the electrons pair with protons (H
+
) as hydrogen
atoms. In this way, coenzymes shuttle energy in the form of
hydrogen atoms from one enzyme to another in a cell.
One of the most important coenzymes is the hydrogen
acceptor nicotinamide adenine dinucleotide (NAD
+
)
(figure 8.13). The NAD
+
molecule is composed of two
nucleotides bound together. As you may recall from
chapter 3, a nucleotide is a five-carbon sugar with one or
more phosphate groups attached to one end and an or-
ganic base attached to the other end. The two nu-
cleotides that make up NAD
+
, nicotinamide monophos-
phate (NMP) and adenine monophosphate (AMP), are
joined head-to-head by their phosphate groups. The two
nucleotides serve different functions in the NAD
+
mole-
cule: AMP acts as the core, providing a shape recognized
by many enzymes; NMP is the active part of the mole-
cule, contributing a site that is readily reduced (that is,
easily accepts electrons).
When NAD
+
acquires an electron and a hydrogen atom
(actually, two electrons and a proton) from the active site of
an enzyme, it is reduced to NADH. The NADH molecule
now carries the two energetic electrons and the proton.
The oxidation of energy-containing molecules, which pro-
vides energy to cells, involves stripping electrons from
those molecules and donating them to NAD
+
. As we’ll see,
much of the energy of NADH is transferred to another
molecule.
Enzymes have an optimum temperature and pH, at
which the enzyme functions most effectively. Inhibitors
decrease enzyme activity, while activators increase it.
The activity of enzymes is often facilitated by cofactors,
which can be metal ions or other substances. Cofactors
that are nonprotein organic molecules are called
coenzymes.
Chapter 8 Energy and Metabolism 153
(a) Competitive inhibition (b) Noncompetitive inhibition
Competitive
inhibitor
inteferes with
active site of
enzyme so
substrate
cannot bind
Allosteric inhibitor
changes shape of
enzyme so it cannot
bind to substrate
Enzyme
Substrate
Enzyme
Substrate
FIGURE 8.12
How enzymes can be inhibited. (a) In competitive inhibition, the
inhibitor interferes with the active site of the enzyme. (b) In
noncompetitive inhibition, the inhibitor binds to the enzyme at a
place away from the active site, effecting a conformational change
in the enzyme so that it can no longer bind to its substrate.
N
+
OCH
2
HH
P
O
–
O
O
OPO
O
–
OH OH
OH OH
HH
O
CH
2
HH
HH
O
C
O
NH
2
N
N
N
N
NH
2
NMP
reactive
group
AMP
group
FIGURE 8.13
The chemical structure of nicotinamide adenine dinucleotide
(NAD
+
). This key cofactor is composed of two nucleotides, NMP
and AMP, attached head-to-head.
What Is ATP?
The chief energy currency all cells use is a molecule called
adenosine triphosphate (ATP). Cells use their supply of
ATP to power almost every energy-requiring process they
carry out, from making sugars, to supplying activation en-
ergy for chemical reactions, to actively transporting sub-
stances across membranes, to moving through their envi-
ronment and growing.
Structure of the ATP Molecule
Each ATP molecule is a nucleotide composed of three
smaller components (figure 8.14). The first component is a
five-carbon sugar, ribose, which serves as the backbone to
which the other two subunits are attached. The second
component is adenine, an organic molecule composed of
two carbon-nitrogen rings. Each of the nitrogen atoms in
the ring has an unshared pair of electrons and weakly at-
tracts hydrogen ions. Adenine, therefore, acts chemically as
a base and is usually referred to as a nitrogenous base (it is
one of the four nitrogenous bases found in DNA and
RNA). The third component of ATP is a triphosphate
group (a chain of three phosphates).
How ATP Stores Energy
The key to how ATP stores energy lies in its triphosphate
group. Phosphate groups are highly negatively charged, so
they repel one another strongly. Because of the electrosta-
tic repulsion between the charged phosphate groups, the
two covalent bonds joining the phosphates are unstable.
The ATP molecule is often referred to as a “coiled spring,”
the phosphates straining away from one another.
The unstable bonds holding the phosphates together in
the ATP molecule have a low activation energy and are
easily broken. When they break, they can transfer a consid-
erable amount of energy. In most reactions involving ATP,
only the outermost high-energy phosphate bond is hy-
drolyzed, cleaving off the phosphate group on the end.
When this happens, ATP becomes adenosine diphos-
phate (ADP), and energy equal to 7.3 kcal/mole is released
under standard conditions. The liberated phosphate group
usually attaches temporarily to some intermediate mole-
cule. When that molecule is dephosphorylated, the phos-
phate group is released as inorganic phosphate (P
i
).
How ATP Powers Energy-Requiring Reactions
Cells use ATP to drive endergonic reactions. Such reac-
tions do not proceed spontaneously, because their products
possess more free energy than their reactants. However, if
the cleavage of ATP’s terminal high-energy bond releases
more energy than the other reaction consumes, the overall
energy change of the two coupled reactions will be exer-
gonic (energy releasing) and they will both proceed. Be-
cause almost all endergonic reactions require less energy
than is released by the cleavage of ATP, ATP can provide
most of the energy a cell needs.
The same feature that makes ATP an effective energy
donor—the instability of its phosphate bonds—precludes
it from being a good long-term energy storage molecule.
Fats and carbohydrates serve that function better. Most
cells do not maintain large stockpiles of ATP. Instead,
they typically have only a few seconds’ supply of ATP at
any given time, and they continually produce more from
ADP and P
i
.
The instability of its phosphate bonds makes ATP an
excellent energy donor.
154 Part III Energetics
8.3 ATP is the energy currency of life.
Triphosphate
group
High-energy
bonds
O == P — O
–
O
–
—
O
O == P — O
–
O
O == P — O
O
–
—
AMP
core
CH
2
HH
OH OH
HH
O
N
N
N
N
NH
2
Adenine
Ribose
(a)
(b)
FIGURE 8.14
The ATP molecule. (a) The model and (b) structural diagram
both show that like NAD
+
, ATP has a core of AMP. In ATP the
reactive group added to the end of the AMP phosphate group is
not another nucleotide but rather a chain of two additional
phosphate groups. The bonds connecting these two phosphate
groups to each other and to AMP are energy-storing bonds.
Biochemical Pathways: The
Organizational Units of Metabolism
This living chemistry, the total of all chemical reactions
carried out by an organism, is called metabolism (Greek
metabole, “change”). Those reactions that expend energy to
make or transform chemical bonds are called anabolic reac-
tions, or anabolism. Reactions that harvest energy when
chemical bonds are broken are called catabolic reactions, or
catabolism.
Organisms contain thousands of different kinds of en-
zymes that catalyze a bewildering variety of reactions.
Many of these reactions in a cell occur in sequences called
biochemical pathways. In such pathways, the product of
one reaction becomes the substrate for the next (figure
8.15). Biochemical pathways are the organizational units of
metabolism, the elements an organism controls to achieve
coherent metabolic activity. Most sequential enzyme steps
in biochemical pathways take place in specific compart-
ments of the cell; the steps of the citric acid cycle (chapter
9), for example, occur inside mitochondria. By determining
where many of the enzymes that catalyze these steps are lo-
cated, we can “map out” a model of metabolic processes in
the cell.
How Biochemical Pathways Evolved
In the earliest cells, the first biochemical processes proba-
bly involved energy-rich molecules scavenged from the en-
vironment. Most of the molecules necessary for these
processes are thought to have existed in the “organic soup”
of the early oceans. The first catalyzed reactions are
thought to have been simple, one-step reactions that
brought these molecules together in various combinations.
Eventually, the energy-rich molecules became depleted in
the external environment, and only organisms that had
evolved some means of making those molecules from other
substances in the environment could survive. Thus, a hypo-
thetical reaction,
F
+ ?→ H
G
where two energy-rich molecules (F and G) react to pro-
duce compound H and release energy, became more com-
plex when the supply of F in the environment ran out. A
new reaction was added in which the depleted molecule, F,
is made from another molecule, E, which was also present
in the environment:
E ?→ F
+ ?→ H
G
When the supply of E in turn became depleted, organisms
that were able to make it from some other available precur-
sor, D, survived. When D became depleted, those organ-
isms in turn were replaced by ones able to synthesize D
from another molecule, C:
C ?→ D ?→ E ?→ F
+ ?→ H
G
This hypothetical biochemical pathway would have
evolved slowly through time, with the final reactions in the
pathway evolving first and earlier reactions evolving later.
Looking at the pathway now, we would say that the organ-
ism, starting with compound C, is able to synthesize H by
means of a series of steps. This is how the biochemical
pathways within organisms are thought to have evolved—
not all at once, but one step at a time, backward.
Chapter 8 Energy and Metabolism 155
8.4 Metabolism is the chemical life of a cell.
Product
Enzyme 1
Enzyme 3
Enzyme 4
Substrate
Enzyme 2
FIGURE 8.15
A biochemical pathway. The original substrate is acted on by
enzyme 1, changing the substrate to a new form recognized by
enzyme 2. Each enzyme in the pathway acts on the product of the
previous stage.
How Biochemical Pathways Are
Regulated
For a biochemical pathway to operate effi-
ciently, its activity must be coordinated
and regulated by the cell. Not only is it
unnecessary to synthesize a compound
when plenty is already present, doing so
would waste energy and raw materials that
could be put to use elsewhere. It is, there-
fore, advantageous for a cell to temporar-
ily shut down biochemical pathways when
their products are not needed.
The regulation of simple biochemical
pathways often depends on an elegant
feedback mechanism: the end product of
the pathway binds to an allosteric site on
the enzyme that catalyzes the first reaction
in the pathway. In the hypothetical path-
way we just described, the enzyme catalyz-
ing the reaction C ?→ D would possess
an allosteric site for H, the end product of
the pathway. As the pathway churned out
its product and the amount of H in the
cell increased, it would become increas-
ingly likely that one of the H molecules
would encounter the allosteric site on the
C ?→ D enzyme. If the product H func-
tioned as an allosteric inhibitor of the en-
zyme, its binding to the enzyme would es-
sentially shut down the reaction C ?→
D. Shutting down this reaction, the first
reaction in the pathway, effectively shuts
down the whole pathway. Hence, as the cell produces in-
creasing quantities of the product H, it automatically in-
hibits its ability to produce more. This mode of regulation
is called feedback inhibition (figure 8.16).
A biochemical pathway is an organized series of
reactions, often regulated as a unit.
156 Part III Energetics
endergonic reaction A chemical reaction
to which energy from an outside source
must be added before the reaction proceeds;
the opposite of an exergonic reaction.
entropy A measure of the randomness or
disorder of a system. In cells, it is a measure
of how much energy has become so dis-
persed (usually as evenly distributed heat)
that it is no longer available to do work.
exergonic reaction. An energy-yielding
chemical reaction. Exergonic reactions tend
to proceed spontaneously, although activa-
tion energy is required to initiate them.
free energy Energy available to do work.
kilocalorie 1000 calories. A calorie is the
heat required to raise the temperature of
1 gram of water by 1°C.
metabolism The sum of all chemical
processes occurring within a living cell or
organism.
oxidation The loss of an electron by an
atom or molecule. It occurs simultaneously
with reduction of some other atom or mole-
cule because an electron that is lost by one
is gained by another.
reduction The gain of an electron by an
atom or molecule. Oxidation-reduction re-
actions are an important means of energy
transfer within living systems.
substrate A molecule on which an en-
zyme acts; the initial reactant in an enzyme-
catalyzed reaction.
activation energy The energy required to
destabilize chemical bonds and to initiate a
chemical reaction.
catalysis Acceleration of the rate of a
chemical reaction by lowering the activa-
tion energy.
coenzyme A nonprotein organic mole-
cule that plays an accessory role in enzyme-
catalyzed reactions, often by acting as a
donor or acceptor of electrons. NAD
+
is a
coenzyme.
A Vocabulary of
Metabolism
End
product
Initial
substrate
Intermediate
A
Intermediate
B
End product
End-product inhibition
+
(b)
Initial
substrate
Intermediate
A
Intermediate
B
End product
Enzyme 1
Enzyme 2
Enzyme 3
Enzyme 2
Enzyme 3
No end-product inhibition
(a)
Enzyme 1
FIGURE 8.16
Feedback inhibition. (a) A biochemical pathway with no feedback inhibition. (b) A
biochemical pathway in which the final end product becomes the allosteric effector for
the first enzyme in the pathway. In other words, the formation of the pathway’s final
end product stops the pathway.
The Evolution of Metabolism
Metabolism has changed a great deal as life on earth has
evolved. This has been particularly true of the reactions or-
ganisms use to capture energy from the sun to build or-
ganic molecules (anabolism), and then break down organic
molecules to obtain energy (catabolism). These processes,
the subject of the next two chapters, evolved in concert
with each other.
Degradation
The most primitive forms of life are thought to have ob-
tained chemical energy by degrading, or breaking down,
organic molecules that were abiotically produced.
The first major event in the evolution of metabolism was
the origin of the ability to harness chemical bond energy.
At an early stage, organisms began to store this energy in
the bonds of ATP, an energy carrier used by all organisms
today.
Glycolysis
The second major event in the evolution of metabolism
was glycolysis, the initial breakdown of glucose. As proteins
evolved diverse catalytic functions, it became possible to
capture a larger fraction of the chemical bond energy in or-
ganic molecules by breaking chemical bonds in a series of
steps. For example, the progressive breakdown of the six-
carbon sugar glucose into three-carbon molecules is per-
formed in a series of 10 steps that results in the net produc-
tion of two ATP molecules. The energy for the synthesis of
ATP is obtained by breaking chemical bonds and forming
new ones with less bond energy, the energy difference
being channeled into ATP production. This biochemical
pathway is called glycolysis.
Glycolysis undoubtedly evolved early in the history of life
on earth, since this biochemical pathway has been retained
by all living organisms. It is a chemical process that does not
appear to have changed for well over 3 billion years.
Anaerobic Photosynthesis
The third major event in the evolution of metabolism was
anaerobic photosynthesis. Early in the history of life, some
organisms evolved a different way of generating ATP,
called photosynthesis. Instead of obtaining energy for ATP
synthesis by reshuffling chemical bonds, as in glycolysis,
these organisms developed the ability to use light to pump
protons out of their cells, and to use the resulting proton
gradient to power the production of ATP, a process called
chemiosmosis.
Photosynthesis evolved in the absence of oxygen and
works well without it. Dissolved H
2
S, present in the oceans
beneath an atmosphere free of oxygen gas, served as a ready
source of hydrogen atoms for building organic molecules.
Free sulfur was produced as a by-product of this reaction.
Nitrogen Fixation
Nitrogen fixation was the fourth major step in the evolu-
tion of metabolism. Proteins and nucleic acids cannot be
synthesized from the products of photosynthesis because
both of these biologically critical molecules contain ni-
trogen. Obtaining nitrogen atoms from N
2
gas, a process
called nitrogen fixation, requires the breaking of an N≡N
triple bond. This important reaction evolved in the
hydrogen-rich atmosphere of the early earth, an atmos-
phere in which no oxygen was present. Oxygen acts as a
poison to nitrogen fixation, which today occurs only in
oxygen-free environments, or in oxygen-free compart-
ments within certain bacteria.
Oxygen-Forming Photosynthesis
The substitution of H
2
O for H
2
S in photosynthesis was the
fifth major event in the history of metabolism. Oxygen-
forming photosynthesis employs H
2
O rather than H
2
S as a
source of hydrogen atoms and their associated electrons.
Because it garners its hydrogen atoms from reduced oxygen
rather than from reduced sulfur, it generates oxygen gas
rather than free sulfur.
More than 2 billion years ago, small cells capable of car-
rying out this oxygen-forming photosynthesis, such as
cyanobacteria, became the dominant forms of life on earth.
Oxygen gas began to accumulate in the atmosphere. This
was the beginning of a great transition that changed condi-
tions on earth permanently. Our atmosphere is now 20.9%
oxygen, every molecule of which is derived from an
oxygen-forming photosynthetic reaction.
Aerobic Respiration
Aerobic respiration is the sixth and final event in the his-
tory of metabolism. This cellular process harvests energy
by stripping energetic electrons from organic molecules.
Aerobic respiration employs the same kind of proton
pumps as photosynthesis, and is thought to have evolved as
a modification of the basic photosynthetic machinery.
However, the hydrogens and their associated electrons are
not obtained from H
2
S or H
2
O, as in photosynthesis, but
rather from the breakdown of organic molecules.
Biologists think that the ability to carry out photosyn-
thesis without H
2
S first evolved among purple nonsulfur
bacteria, which obtain their hydrogens from organic com-
pounds instead. It was perhaps inevitable that among the
descendants of these respiring photosynthetic bacteria,
some would eventually do without photosynthesis en-
tirely, subsisting only on the energy and hydrogens de-
rived from the breakdown of organic molecules. The mi-
tochondria within all eukaryotic cells are thought to be
their descendants.
Six major innovations highlight the evolution of
metabolism as we know it today.
Chapter 8 Energy and Metabolism 157
? Energy Conversion
? Catalysis
? Thermodynamics
? Coupled Reactions
158 Part III Energetics
Chapter 8
Summary Questions Media Resources
8.1 The laws of thermodynamics describe how energy changes.
? Energy is the capacity to bring about change, to
provide motion against a force, or to do work.
? Kinetic energy is actively engaged in doing work,
while potential energy has the capacity to do so.
? An oxidation-reduction (redox) reaction is one in
which an electron is taken from one atom or
molecule (oxidation) and donated to another
(reduction).
? The First Law of Thermodynamics states that the
amount of energy in the universe is constant; energy
is neither lost nor created.
? The Second Law of Thermodynamics states that
disorder in the universe (entropy) tends to increase.
? Any chemical reaction whose products contain less
free energy than the original reactants can proceed
spontaneously. However, the difference in free
energy does not determine the rate of the reaction.
? The rate of a reaction depends on the amount of
activation energy required to break existing bonds.
? Catalysis is the process of lowering activation
energies by stressing chemical bonds.
1. What is the difference
between anabolism and
catabolism?
2. Define oxidation and
reduction. Why must these two
reactions always occur in
concert?
3. State the First and Second
Laws of Thermodynamics.
4. What is heat? What is
entropy? What is free energy?
5. What is the difference
between an exergonic and an
endergonic reaction? Which
type of reaction tends to proceed
spontaneously?
6. Define activation energy.
How does a catalyst affect the
final proportion of reactant
converted into product?
? Enzymes are the major catalysts of cells; they affect
the rate of a reaction but not the ultimate balance
between reactants and products.
? Cells contain many different enzymes, each of which
catalyzes a specific reaction.
? The specificity of an enzyme is due to its active site,
which fits only one or a few types of substrate
molecules.
7. How are the rates of
enzyme-catalyzed reactions
affected by temperature? What
is the molecular basis for the
effect on reaction rate?
8. What is the difference
between the active site and an
allosteric site on an enzyme?
8.2 Enzymes are biological catalysts.
? Cells obtain energy from photosynthesis and the
oxidation of organic molecules and use it to
manufacture ATP from ADP and phosphate.
? The energy stored in ATP is then used to drive
endergonic reactions.
9. What part of the ATP
molecule contains the bond that
is employed to provide energy
for most of the endergonic
reactions in cells?
8.3 ATP is the energy currency of life.
? Generally, the final reactions of a biochemical
pathway evolved first; preceding reactions in the
pathway were added later, one step at a time.
10. What is a biochemical
pathway? How does feedback
inhibition regulate the activity of
a biochemical pathway?
8.4 Metabolism is the chemical life of a cell.
? Exploration:
Thermodynamics
? Exploration: Kinetics
? Enzymes
? ATP
? Feedback Inhibition
http://www.mhhe.com/raven6e http://www.biocourse.com
159
9
How Cells Harvest Energy
Concept Outline
9.1 Cells harvest the energy in chemical bonds.
Using Chemical Energy to Drive Metabolism. The
energy in C—H, C—O, and other chemical bonds can be
captured and used to fuel the synthesis of ATP.
9.2 Cellular respiration oxidizes food molecules.
An Overview of Glucose Catabolism. The chemical
energy in sugar is harvested by both substrate-level
phosphorylation and by aerobic respiration.
Stage One: Glycolysis. The 10 reactions of glycolysis
capture energy from glucose by reshuffling the bonds.
Stage Two: The Oxidation of Pyruvate. Pyruvate, the
product of glycolysis, is oxidized to acetyl-CoA.
Stage Three: The Krebs Cycle. In a series of reactions,
electrons are stripped from acetyl-CoA.
Harvesting Energy by Extracting Electrons. The
respiration of glucose is a series of oxidation-reduction
reactions which involve stripping electrons from glucose and
using the energy of these electrons to power the synthesis of
ATP.
Stage Four: The Electron Transport Chain. The
electrons harvested from glucose pass through a chain of
membrane proteins that use the energy to pump protons,
driving the synthesis of ATP.
Summarizing Aerobic Respiration. The oxidation of
glucose by aerobic respiration in eukaryotes produces up to
three dozen ATP molecules, over half the energy in the
chemical bonds of glucose.
Regulating Aerobic Respiration. High levels of ATP
tend to shut down cellular respiration by feedback-inhibiting
key reactions.
9.3 Catabolism of proteins and fats can yield
considerable energy.
Glucose Is Not the Only Source of Energy. Proteins
and fats are dismantled and the products fed into cellular
respiration.
9.4 Cells can metabolize food without oxygen.
Fermentation. Fermentation allows continued metabolism
in the absence of oxygen by donating the electrons harvested
in glycolysis to organic molecules.
L
ife is driven by energy. All the activities organisms
carry out—the swimming of bacteria, the purring of a
cat, your reading of these words—use energy. In this chap-
ter, we will discuss the processes all cells use to derive
chemical energy from organic molecules and to convert
that energy to ATP. We will consider photosynthesis,
which uses light energy rather than chemical energy, in de-
tail in chapter 10. We examine the conversion of chemical
energy to ATP first because all organisms, both photosyn-
thesizers and the organisms that feed on them (like the
field mice in figure 9.1), are capable of harvesting energy
from chemical bonds. As you will see, though, this process
and photosynthesis have much in common.
FIGURE 9.1
Harvesting chemical energy. Organisms such as these harvest
mice depend on the energy stored in the chemical bonds of the
food they eat to power their life processes.
fireplace. In both instances, the reactants are carbohydrates
and oxygen, and the products are carbon dioxide, water,
and energy:
C
6
H
12
O
6
+ 6 O
2
?→ 6 CO
2
+ 6 H
2
O + energy (heat or ATP)
The change in free energy in this reaction is –720 kilo-
calories (–3012 kilojoules) per mole of glucose under the
conditions found within a cell (the traditional value of
–686 kilocalories, or –2870 kJ, per mole refers to stan-
dard conditions—room temperature, one atmosphere of
pressure, etc.). This change in free energy results largely
from the breaking of the six C—H bonds in the glucose
molecule. The negative sign indicates that the products
possess less free energy than the reactants. The same
amount of energy is released whether glucose is catabo-
lized or burned, but when it is burned most of the energy
is released as heat. This heat cannot be used to perform
work in cells. The key to a cell’s ability to harvest useful
energy from the catabolism of food molecules such as
glucose is its conversion of a portion of the energy into a
more useful form. Cells do this by using some of the en-
ergy to drive the production of ATP, a molecule that can
power cellular activities.
160 Part III Energetics
Using Chemical Energy to
Drive Metabolism
Plants, algae, and some bacteria harvest the en-
ergy of sunlight through photosynthesis, convert-
ing radiant energy into chemical energy. These
organisms, along with a few others that use
chemical energy in a similar way, are called au-
totrophs (“self-feeders”). All other organisms
live on the energy autotrophs produce and are
called heterotrophs (“fed by others”). At least
95% of the kinds of organisms on earth—all ani-
mals and fungi, and most protists and bacteria—
are heterotrophs.
Where is the chemical energy in food, and
how do heterotrophs harvest it to carry out the
many tasks of living (figure 9.2)? Most foods
contain a variety of carbohydrates, proteins,
and fats, all rich in energy-laden chemical
bonds. Carbohydrates and fats, for example,
possess many carbon-hydrogen (C—H), as well
as carbon-oxygen (C—O) bonds. The job of ex-
tracting energy from this complex organic mix-
ture is tackled in stages. First, enzymes break
the large molecules down into smaller ones, a
process called digestion. Then, other enzymes
dismantle these fragments a little at a time, har-
vesting energy from C—H and other chemical
bonds at each stage. This process is called ca-
tabolism.
While you obtain energy from many of the constituents
of food, it is traditional to focus first on the catabolism of
carbohydrates. We will follow the six-carbon sugar, glucose
(C
6
H
12
O
6
), as its chemical bonds are progressively har-
vested for energy. Later, we will come back and examine
the catabolism of proteins and fats.
Cellular Respiration
The energy in a chemical bond can be visualized as poten-
tial energy borne by the electrons that make up the cova-
lent bond. Cells harvest this energy by putting the elec-
trons to work, often to produce ATP, the energy currency
of the cell. Afterward, the energy-depleted electron (associ-
ated with a proton as a hydrogen atom) is donated to some
other molecule. When oxygen gas (O
2
) accepts the hydro-
gen atom, water forms, and the process is called aerobic
respiration. When an inorganic molecule other than oxy-
gen accepts the hydrogen, the process is called anaerobic
respiration. When an organic molecule accepts the hydro-
gen atom, the process is called fermentation.
Chemically, there is little difference between the catabo-
lism of carbohydrates in a cell and the burning of wood in a
9.1 Cells harvest the energy in chemical bonds.
FIGURE 9.2
Start every day with a good breakfast. The carbohydrates, proteins, and fats
in this fish contain energy that the bear’s cells can use to power their daily
activities.
The ATP Molecule
Adenosine triphosphate (ATP) is the energy currency of
the cell, the molecule that transfers the energy captured
during respiration to the many sites that use energy in the
cell. How is ATP able to transfer energy so readily? Recall
from chapter 8 that ATP is composed of a sugar (ribose)
bound to an organic base (adenine) and a chain of three
phosphate groups. As shown in figure 9.3, each phosphate
group is negatively charged. Because like charges repel
each other, the linked phosphate groups push against the
bond that holds them together. Like a cocked mousetrap,
the linked phosphates store the energy of their electrostatic
repulsion. Transferring a phosphate group to another mol-
ecule relaxes the electrostatic spring of ATP, at the same
time cocking the spring of the molecule that is phosphory-
lated. This molecule can then use the energy to undergo
some change that requires work.
How Cells Use ATP
Cells use ATP to do most of those activities that require
work. One of the most obvious is movement. Some bacte-
ria swim about, propelling themselves through the water by
rapidly spinning a long, tail-like flagellum, much as a ship
moves by spinning a propeller. During your development
as an embryo, many of your cells moved about, crawling
over one another to reach new positions. Movement also
occurs within cells. Tiny fibers within muscle cells pull
against one another when muscles contract. Mitochondria
pass a meter or more along the narrow nerve cells that con-
nect your feet with your spine. Chromosomes are pulled by
microtubules during cell division. All of these movements
by cells require the expenditure of ATP energy.
A second major way cells use ATP is to drive endergonic
reactions. Many of the synthetic activities of the cell are en-
dergonic, because building molecules takes energy. The
chemical bonds of the products of these reactions contain
more energy, or are more organized, than the reactants.
The reaction can’t proceed until that extra energy is sup-
plied to the reaction. It is ATP that provides this needed
energy.
How ATP Drives Endergonic Reactions
How does ATP drive an endergonic reaction? The en-
zyme that catalyzes the endergonic reaction has two bind-
ing sites on its surface, one for the reactant and another
for ATP. The ATP site splits the ATP molecule, liberat-
ing over 7 kcal (30 kJ) of chemical energy. This energy
pushes the reactant at the second site “uphill,” driving the
endergonic reaction. (In a similar way, you can make
water in a swimming pool leap straight up in the air, de-
spite the fact that gravity prevents water from rising spon-
taneously—just jump in the pool! The energy you add
going in more than compensates for the force of gravity
holding the water back.)
When the splitting of ATP molecules drives an energy-
requiring reaction in a cell, the two parts of the reaction—
ATP-splitting and endergonic—take place in concert. In
some cases, the two parts both occur on the surface of the
same enzyme; they are physically linked, or “coupled,” like
two legs walking. In other cases, a high-energy phosphate
from ATP attaches to the protein catalyzing the ender-
gonic process, activating it (figure 9.4). Coupling energy-
requiring reactions to the splitting of ATP in this way is
one of the key tools cells use to manage energy.
The catabolism of glucose into carbon dioxide and
water in living organisms releases about 720 kcal
(3012 kJ) of energy per mole of glucose. This energy is
captured in ATP, which stores the energy by linking
charged phosphate groups near one another. When the
phosphate bonds in ATP are hydrolyzed, energy is
released and available to do work.
Chapter 9 How Cells Harvest Energy 161
Triphosphate group
Sugar
Adenine
NH
2
OP CH
2
O
O
O
–
P
O
O
O
–
P
O
–
O
O
–
OH OH
O
N
N
N
N
FIGURE 9.3
Structure of the ATP molecule. ATP is composed of an organic
base and a chain of phosphates attached to opposite ends of a five-
carbon sugar. Notice that the charged regions of the phosphate
chain are close to one another. These like charges tend to repel
one another, giving the bonds that hold them together a
particularly high energy transfer potential.
ADP
Inactive Active
ATP
P
FIGURE 9.4
How ATP drives an endergonic reaction. In many cases, a
phosphate group split from ATP activates a protein, catalyzing an
endergonic process.
An Overview of Glucose
Catabolism
Cells are able to make ATP from the
catabolism of organic molecules in two
different ways.
1. Substrate-level phosphoryla-
tion. In the first, called
substrate- level phosphoryla-
tion, ATP is formed by transfer-
ring a phosphate group directly
to ADP from a phosphate-bear-
ing intermediate (figure 9.5).
During glycolysis, discussed
below, the chemical bonds of glu-
cose are shifted around in reac-
tions that provide the energy re-
quired to form ATP.
2. Aerobic respiration. In the
second, called aerobic respira-
tion, ATP forms as electrons are
harvested, transferred along the electron transport
chain, and eventually donated to oxygen gas. Eukary-
otes produce the majority of their ATP from glucose
in this way.
In most organisms, these two processes are combined.
To harvest energy to make ATP from the sugar glucose in
the presence of oxygen, the cell carries out a complex se-
ries of enzyme-catalyzed reactions that occur in four
stages: the first stage captures energy by substrate-level
phosphorylation through glycolysis, the following three
stages carry out aerobic respiration by oxidizing the end
product of glycolysis.
Glycolysis
Stage One: Glycolysis. The first stage of extracting en-
ergy from glucose is a 10-reaction biochemical pathway
called glycolysis that produces ATP by substrate-level
phosphorylation. The enzymes that catalyze the glycolytic
reactions are in the cytoplasm of the cell, not bound to any
membrane or organelle. Two ATP molecules are used up
early in the pathway, and four ATP molecules are formed
by substrate-level phosphorylation. This yields a net of
two ATP molecules for each molecule of glucose catabo-
lized. In addition, four electrons are harvested as NADH
that can be used to form ATP by aerobic respiration. Still,
the total yield of ATP is small. When the glycolytic
process is completed, the two molecules of pyruvate that
are formed still contain most of the energy the original
glucose molecule held.
Aerobic Respiration
Stage Two: Pyruvate Oxidation. In the second stage,
pyruvate, the end product from glycolysis, is converted into
carbon dioxide and a two-carbon molecule called acetyl-
CoA. For each molecule of pyruvate converted, one mole-
cule of NAD
+
is reduced to NADH.
Stage Three: The Krebs Cycle. The third stage intro-
duces this acetyl-CoA into a cycle of nine reactions called
the Krebs cycle, named after the British biochemist, Sir
Hans Krebs, who discovered it. (The Krebs cycle is also
called the citric acid cycle, for the citric acid, or citrate,
formed in its first step, and less commonly, the tricar-
boxylic acid cycle, because citrate has three carboxyl
groups.) In the Krebs cycle, two more ATP molecules are
extracted by substrate-level phosphorylation, and a large
number of electrons are removed by the reduction of
NAD
+
to NADH.
Stage Four: Electron Transport Chain. In the fourth
stage, the energetic electrons carried by NADH are em-
ployed to drive the synthesis of a large amount of ATP by
the electron transport chain.
Pyruvate oxidation, the reactions of the Krebs cycle, and
ATP production by electron transport chains occur within
many forms of bacteria and inside the mitochondria of all
eukaryotes. Recall from chapter 5 that mitochondria are
thought to have evolved from bacteria. Although plants and
algae can produce ATP by photosynthesis, they also pro-
duce ATP by aerobic respiration, just as animals and other
nonphotosynthetic eukaryotes do. Figure 9.6 provides an
overview of aerobic respiration.
162 Part III Energetics
9.2 Cellular respiration oxidizes food molecules.
P
PEP
Enzyme
ADP
Adenosine
P
P
Pyruvate
ATP
Adenosine
P
P
P
FIGURE 9.5
Substrate-level phosphorylation. Some molecules, such as phosphoenolpyruvate (PEP),
possess a high-energy phosphate bond similar to the bonds in ATP. When PEP’s
phosphate group is transferred enzymatically to ADP, the energy in the bond is conserved
and ATP is created.
Anaerobic Respiration
In the presence of oxygen, cells can respire aerobically,
using oxygen to accept the electrons harvested from food
molecules. In the absence of oxygen to accept the electrons,
some organisms can still respire anaerobically, using inor-
ganic molecules to accept the electrons. For example,
many bacteria use sulfur, nitrate, or other inorganic com-
pounds as the electron acceptor in place of oxygen.
Methanogens. Among the heterotrophs that practice
anaerobic respiration are primitive archaebacteria such as
the thermophiles discussed in chapter 4. Some of these,
called methanogens, use CO
2
as the electron acceptor, re-
ducing CO
2
to CH
4
(methane) with the hydrogens derived
from organic molecules produced by other organisms.
Sulfur Bacteria. Evidence of a second anaerobic respi-
ratory process among primitive bacteria is seen in a
group of rocks about 2.7 billion years old, known as the
Woman River iron formation. Organic material in these
rocks is enriched for the light isotope of sulfur,
32
S, rela-
tive to the heavier isotope
34
S. No known geochemical
process produces such enrichment, but biological sulfur
reduction does, in a process still carried out today by cer-
tain primitive bacteria. In this sulfate respiration, the
bacteria derive energy from the reduction of inorganic
sulfates (SO
4
) to H
2
S. The hydrogen atoms are obtained
from organic molecules other organisms produce. These
bacteria thus do the same thing methanogens do, but
they use SO
4
as the oxidizing (that is, electron-accepting)
agent in place of CO
2
.
The sulfate reducers set the stage for the evolution of
photosynthesis, creating an environment rich in H
2
S. As
discussed in chapter 8, the first form of photosynthesis ob-
tained hydrogens from H
2
S using the energy of sunlight.
In aerobic respiration, the cell harvests energy from
glucose molecules in a sequence of four major
pathways: glycolysis, pyruvate oxidation, the Krebs
cycle, and the electron transport chain. Oxygen is the
final electron acceptor. Anaerobic respiration donates
the harvested electrons to other inorganic compounds.
Chapter 9 How Cells Harvest Energy 163
NADH
NADH
H
2
O
CO
2
Extracellular fluid
Lactate
Mitochondrion
Plasma
membrane
ATP
ATP
ATP
E
le
ct
ro
n
tr
a
n
s
p
o
rt
s
y
s
te
m
O
2
NADH
Krebs
cycle
Acetyl-CoA
Pyruvate
Cytoplasm
GlycolysisGlucose
FIGURE 9.6
An overview of aerobic respiration.
Stage One: Glycolysis
The metabolism of primitive organisms focused on glu-
cose. Glucose molecules can be dismantled in many ways,
but primitive organisms evolved a glucose-catabolizing
process that releases enough free energy to drive the syn-
thesis of ATP in coupled reactions. This process, called
glycolysis, occurs in the cytoplasm and involves a se-
quence of 10 reactions that convert glucose into 2 three-
carbon molecules of pyruvate (figure 9.7). For each mole-
cule of glucose that passes through this transformation,
the cell nets two ATP molecules by substrate-level phos-
phorylation.
Priming
The first half of glycolysis consists of five sequential reac-
tions that convert one molecule of glucose into two mole-
cules of the three-carbon compound, glyceraldehyde 3-
phosphate (G3P). These reactions demand the expenditure
of ATP, so they are an energy-requiring process.
Step A: Glucose priming. Three reactions “prime”
glucose by changing it into a compound that can be
cleaved readily into 2 three-carbon phosphorylated mole-
cules. Two of these reactions require the cleavage of ATP,
so this step requires the cell to use two ATP molecules.
Step B: Cleavage and rearrangement. In the first of
the remaining pair of reactions, the six-carbon product
of step A is split into 2 three-carbon molecules. One is
G3P, and the other is then converted to G3P by the sec-
ond reaction (figure 9.8).
164 Part III Energetics
OVERVIEW OF GLYCOLYSIS
12 3
(Starting material)
6-carbon sugar diphosphate
6-carbon glucose
2
P P
6-carbon sugar diphosphate
P P
3-carbon sugar
phosphate
P
3-carbon sugar
phosphate
P
3-carbon sugar
phosphate
P
3-carbon sugar
phosphate
P
Priming reactions. Glycolysis begins
with the addition of energy. Two high-
energy phosphates from two molecules
of ATP are added to the six-carbon
molecule glucose, producing a six-carbon
molecule with two phosphates.
3-carbon
pyruvate
2
NADH
Cleavage reactions. Then, the six-carbon
molecule with two phosphates is split in
two, forming two three-carbon sugar
phosphates.
Energy-harvesting reactions. Finally, in
a series of reactions, each of the two three-
carbon sugar phosphates is converted to
pyruvate. In the process, an energy-rich
hydrogen is harvested as NADH, and two
ATP molecules are formed.
3-carbon
pyruvate
ATP
ATP 2
NADH
ATP
FIGURE 9.7
How glycolysis works.
Chapter 9 How Cells Harvest Energy 165
1. Phosphorylation of
glucose by ATP.
Glucose
Hexokinase
Phosphoglucoisomerase
Phosphofructokinase
Glyceraldehyde 3-
phosphate (G3P)
P
i
P
i
Dihydroxyacetone
phosphate
Glucose 6-phosphate
Fructose 6-phosphate
Fructose 1,6-bisphosphate
Isomerase
Triose phosphate
dehydrogenase
Aldolase
1,3-Bisphosphoglycerate
(BPG)
1,3-Bisphosphoglycerate
(BPG)
3-Phosphoglycerate
(3PG)
3-Phosphoglycerate
(3PG)
2-Phosphoglycerate
(2PG)
2-Phosphoglycerate
(2PG)
Phosphoenolpyruvate
(PEP)
Phosphoenolpyruvate
(PEP)
Pyruvate Pyruvate
2–3. Rearrangement,
followed by a second
ATP phosphorylation.
4–5. The six-carbon
molecule is split into
two three-carbon
molecules—one G3P,
another that is
converted into G3P in
another reaction.
6. Oxidation followed by
phosphorylation produces
two NADH molecules and
two molecules of BPG,
each with one high-energy
phosphate bond.
7. Removal of high-energy
phosphate by two ADP
molecules produces two
ATP molecules and leaves
two 3PG molecules.
8–9. Removal of water yields
two PEP molecules,
each with a high-energy
phosphate bond.
10. Removal of high-energy
phosphate by two ADP
molecules produces two
ATP molecules and two
pyruvate molecules.
1
2
3
6
Phosphoglycerokinase
7
Phosphoglyceromutase
8
EnolaseH
2
OH
2
O
9
Pyruvate kinase
10
4,5
ADP
ATP
ADP
ATP
ADP
ATP
ADP
ATP
ADP
ATP
ADP
ATP
NADH
NAD
+
NADH
NAD
+
O
CO
O
CH
2
OH
CH
2
P
CO
C
O
–
O
CH
3
PCO
C
O
–
O
CH
2
PCOH
C
O
–
O
CH
2
OH
CHOH
PO
C
O
–
O
CH
2
CHOH
PO
C
H
O
CH
2
POCH
2
P O CH
2
O
PO
CH
2
OH
CH
2
OH
CH
2
POCH
2
CHOH
P
P
O
O CO
CH
2
O
O
FIGURE 9.8
The glycolytic pathway. The first five reactions convert a molecule of glucose into two molecules of G3P. The second five reactions
convert G3P into pyruvate.
Substrate-Level Phosphorylation
In the second half of glycolysis, five more reactions convert
G3P into pyruvate in an energy-yielding process that gen-
erates ATP. Overall, then, glycolysis is a series of 10 en-
zyme-catalyzed reactions in which some ATP is invested in
order to produce more.
Step C: Oxidation. Two electrons and one proton are
transferred from G3P to NAD
+
, forming NADH. Note
that NAD
+
is an ion, and that both electrons in the new
covalent bond come from G3P.
Step D: ATP generation. Four reactions convert
G3P into another three-carbon molecule, pyruvate. This
process generates two ATP molecules (see figure 9.5).
Because each glucose molecule is split into two G3P
molecules, the overall reaction sequence yields two mol-
ecules of ATP, as well as two molecules of NADH and
two of pyruvate:
4 ATP (2 ATP for each of the 2 G3P molecules in step D)
– 2 ATP (used in the two reactions in step A )
2 ATP
Under the nonstandard conditions within a cell, each
ATP molecule produced represents the capture of about 12
kcal (50 kJ) of energy per mole of glucose, rather than the
7.3 traditionally quoted for standard conditions. This
means glycolysis harvests about 24 kcal/mole (100
kJ/mole). This is not a great deal of energy. The total en-
ergy content of the chemical bonds of glucose is 686 kcal
(2870 kJ) per mole, so glycolysis harvests only 3.5% of the
chemical energy of glucose.
Although far from ideal in terms of the amount of en-
ergy it releases, glycolysis does generate ATP. For more
than a billion years during the anaerobic first stages of
life on earth, it was the primary way heterotrophic organ-
isms generated ATP from organic molecules. Like many
biochemical pathways, glycolysis is believed to have
evolved backward, with the last steps in the process being
the most ancient. Thus, the second half of glycolysis, the
ATP-yielding breakdown of G3P, may have been the
original process early heterotrophs used to generate
ATP. The synthesis of G3P from glucose would have ap-
peared later, perhaps when alternative sources of G3P
were depleted.
All Cells Use Glycolysis
The glycolytic reaction sequence is thought to have been
among the earliest of all biochemical processes to evolve. It
uses no molecular oxygen and occurs readily in an anaero-
bic environment. All of its reactions occur free in the cyto-
plasm; none is associated with any organelle or membrane
structure. Every living creature is capable of carrying out
glycolysis. Most present-day organisms, however, can ex-
tract considerably more energy from glucose through aero-
bic respiration.
Why does glycolysis take place even now, since its en-
ergy yield in the absence of oxygen is comparatively so
paltry? The answer is that evolution is an incremental
process: change occurs by improving on past successes. In
catabolic metabolism, glycolysis satisfied the one essential
evolutionary criterion: it was an improvement. Cells that
could not carry out glycolysis were at a competitive disad-
vantage, and only cells capable of glycolysis survived the
early competition of life. Later improvements in catabolic
metabolism built on this success. Glycolysis was not dis-
carded during the course of evolution; rather, it served as
the starting point for the further extraction of chemical
energy. Metabolism evolved as one layer of reactions
added to another, just as successive layers of paint cover
the walls of an old building. Nearly every present-day or-
ganism carries out glycolysis as a metabolic memory of its
evolutionary past.
Closing the Metabolic Circle: The Regeneration
of NAD
+
Inspect for a moment the net reaction of the glycolytic se-
quence:
Glucose + 2 ADP + 2 P
i
+ 2 NAD
+
?→
2 Pyruvate + 2 ATP + 2 NADH + 2 H
+
+ 2 H
2
O
You can see that three changes occur in glycolysis: (1) glu-
cose is converted into two molecules of pyruvate; (2) two
molecules of ADP are converted into ATP via substrate
level phosphorylation; and (3) two molecules of NAD
+
are
reduced to NADH.
The Need to Recycle NADH
As long as food molecules that can be converted into glu-
cose are available, a cell can continually churn out ATP to
drive its activities. In doing so, however, it accumulates
NADH and depletes the pool of NAD
+
molecules. A cell
does not contain a large amount of NAD
+
, and for glycoly-
sis to continue, NADH must be recycled into NAD
+
. Some
other molecule than NAD
+
must ultimately accept the hy-
drogen atom taken from G3P and be reduced. Two mole-
cules can carry out this key task (figure 9.9):
1. Aerobic respiration. Oxygen is an excellent elec-
tron acceptor. Through a series of electron transfers,
the hydrogen atom taken from G3P can be donated
to oxygen, forming water. This is what happens in the
cells of eukaryotes in the presence of oxygen. Because
air is rich in oxygen, this process is also referred to as
aerobic metabolism.
166 Part III Energetics
2. Fermentation. When oxygen is unavailable, an or-
ganic molecule can accept the hydrogen atom instead.
Such fermentation plays an important role in the me-
tabolism of most organisms (figure 9.10), even those
capable of aerobic respiration.
The fate of the pyruvate that is produced by glycolysis
depends upon which of these two processes takes place.
The aerobic respiration path starts with the oxidation of
pyruvate to a molecule called acetyl-CoA, which is then
further oxidized in a series of reactions called the Krebs
cycle. The fermentation path, by contrast, involves the re-
duction of all or part of pyruvate. We will start by examin-
ing aerobic respiration, then look briefly at fermentation.
Glycolysis generates a small amount of ATP by
reshuffling the bonds of glucose molecules. In
glycolysis, two molecules of NAD
+
are reduced to
NADH. NAD
+
must be regenerated for glycolysis to
continue unabated.
Chapter 9 How Cells Harvest Energy 167
NADH
Pyruvate
With oxygen Without oxygen
Acetyl-CoA
Lactate
Ethanol
NAD
+
O
2
NAD
+
NADH
NAD
+
NADH
CO
2
Acetaldehyde
H
2
O
Krebs
cycle
FIGURE 9.9
What happens to pyruvate, the product of glycolysis? In the presence of oxygen, pyruvate is oxidized to acetyl-CoA, which enters the
Krebs cycle. In the absence of oxygen, pyruvate is instead reduced, accepting the electrons extracted during glycolysis and carried by
NADH. When pyruvate is reduced directly, as in muscle cells, the product is lactate. When CO
2
is first removed from pyruvate and the
product, acetaldehyde, is then reduced, as in yeast cells, the product is ethanol.
FIGURE 9.10
Fermentation. The conversion of pyruvate to ethanol takes place
naturally in grapes left to ferment on vines, as well as in
fermentation vats of crushed grapes. Yeasts carry out the process,
but when their conversion increases the ethanol concentration to
about 12%, the toxic effects of the alcohol kill the yeast cells.
What is left is wine.
Stage Two: The Oxidation of Pyruvate
In the presence of oxygen, the oxidation of glucose that be-
gins in glycolysis continues where glycolysis leaves off—
with pyruvate. In eukaryotic organisms, the extraction of
additional energy from pyruvate takes place exclusively in-
side mitochondria. The cell harvests pyruvate’s consider-
able energy in two steps: first, by oxidizing pyruvate to
form acetyl-CoA, and then by oxidizing acetyl-CoA in the
Krebs cycle.
Producing Acetyl-CoA
Pyruvate is oxidized in a single “decarboxylation” reaction
that cleaves off one of pyruvate’s three carbons. This car-
bon then departs as CO
2
(figure 9.11, top). This reaction
produces a two-carbon fragment called an acetyl group, as
well as a pair of electrons and their associated hydrogen,
which reduce NAD
+
to NADH. The reaction is complex,
involving three intermediate stages, and is catalyzed
within mitochondria by a multienzyme complex. As chapter
8 noted, such a complex organizes a series of enzymatic
steps so that the chemical intermediates do not diffuse
away or undergo other reactions. Within the complex,
component polypeptides pass the substrates from one en-
zyme to the next, without releasing them. Pyruvate dehy-
drogenase, the complex of enzymes that removes CO
2
from
pyruvate, is one of the largest enzymes known: it contains
60 subunits! In the course of the reaction, the acetyl
group removed from pyruvate combines with a cofactor
called coenzyme A (CoA), forming a compound known as
acetyl-CoA:
Pyruvate + NAD
+
+ CoA ?→ Acetyl-CoA + NADH + CO
2
This reaction produces a molecule of NADH, which is
later used to produce ATP. Of far greater significance than
the reduction of NAD
+
to NADH, however, is the produc-
tion of acetyl-CoA (figure 9.11, bottom). Acetyl-CoA is im-
portant because so many different metabolic processes gen-
erate it. Not only does the oxidation of pyruvate, an
intermediate in carbohydrate catabolism, produce it, but
the metabolic breakdown of proteins, fats, and other lipids
also generate acetyl-CoA. Indeed, almost all molecules ca-
tabolized for energy are converted into acetyl-CoA. Acetyl-
CoA is then channeled into fat synthesis or into ATP pro-
duction, depending on the organism’s energy
requirements. Acetyl-CoA is a key point of focus for the
many catabolic processes of the eukaryotic cell.
Using Acetyl-CoA
Although the cell forms acetyl-CoA in many ways, only a
limited number of processes use acetyl-CoA. Most of it is
either directed toward energy storage (lipid synthesis, for
example) or oxidized in the Krebs cycle to produce ATP.
Which of these two options is taken depends on the level
of ATP in the cell. When ATP levels are high, the oxida-
tive pathway is inhibited, and acetyl-CoA is channeled
into fatty acid synthesis. This explains why many animals
(humans included) develop fat reserves when they con-
sume more food than their bodies require. Alternatively,
when ATP levels are low, the oxidative pathway is stimu-
lated, and acetyl-CoA flows into energy-producing oxida-
tive metabolism.
In the second energy-harvesting stage of glucose
catabolism, pyruvate is decarboxylated, yielding acetyl-
CoA, NADH, and CO
2
. This process occurs within the
mitochondrion.
168 Part III Energetics
Coenzyme A
Acetyl
group
Acetyl coenzyme A
ATPFat
CO
2
Protein Lipid
NADH
NAD
+
Pyruvate
Glycolysis
FIGURE 9.11
The oxidation of pyruvate. This complex reaction involves the
reduction of NAD
+
to NADH and is thus a significant source of
metabolic energy. Its product, acetyl-CoA, is the starting material
for the Krebs cycle. Almost all molecules that are catabolized for
energy are converted into acetyl-CoA, which is then channeled
into fat synthesis or into ATP production.
Stage Three: The Krebs Cycle
After glycolysis catabolizes glucose to produce pyruvate,
and pyruvate is oxidized to form acetyl-CoA, the third
stage of extracting energy from glucose begins. In this third
stage, acetyl-CoA is oxidized in a series of nine reactions
called the Krebs cycle. These reactions occur in the matrix
of mitochondria. In this cycle, the two-carbon acetyl group
of acetyl-CoA combines with a four-carbon molecule called
oxaloacetate (figure 9.12). The resulting six-carbon mole-
cule then goes through a sequence of electron-yielding oxi-
dation reactions, during which two CO
2
molecules split off,
restoring oxaloacetate. The oxaloacetate is then recycled to
bind to another acetyl group. In each turn of the cycle, a
new acetyl group replaces the two CO
2
molecules lost, and
more electrons are extracted to drive proton pumps that
generate ATP.
Overview of the Krebs Cycle
The nine reactions of the Krebs cycle occur in two steps:
Step A: Priming. Three reactions prepare the six-
carbon molecule for energy extraction. First, acetyl-
CoA joins the cycle, and then chemical groups are
rearranged.
Step B: Energy extraction. Four of the six reactions
in this step are oxidations in which electrons are re-
moved, and one generates an ATP equivalent directly by
substrate-level phosphorylation.
Chapter 9 How Cells Harvest Energy 169
OVERVIEW OF THE KREBS CYCLE
123
Finally, the resulting four-carbon molecule
is further oxidized (hydrogens removed
to form FADH
2
and NADH). This
regenerates the four-carbon starting
material, completing the cycle.
Then, the resulting six-carbon mole-
cule is oxidized (a hydrogen removed
to form NADH) and decarboxylated
(a carbon removed to form CO
2
). Next,
the five-carbon molecule is oxidized
and decarboxylated again, and a
coupled reaction generates ATP.
The Krebs cycle begins when a two-
carbon fragment is transferred from
acetyl-CoA to a four-carbon molecule
(the starting material).
5-carbon molecule4-carbon molecule
(Acetyl-CoA)
6-carbon molecule
4-carbon molecule
(Starting material)
CoA–
CoA
NADH
CO
2
NADH
CO
2
ATP
NADH
FADH
2
6-carbon molecule
4-carbon molecule
4-carbon molecule
(Starting material)
FIGURE 9.12
How the Krebs cycle works.
The Reactions of the Krebs Cycle
The Krebs cycle consists of nine sequential reactions that
cells use to extract energetic electrons and drive the synthe-
sis of ATP (figure 9.13). A two-carbon group from acetyl-
CoA enters the cycle at the beginning, and two CO
2
mole-
cules and several electrons are given off during the cycle.
Reaction 1: Condensation. The two-carbon group from
acetyl-CoA joins with a four-carbon molecule, oxaloac-
etate, to form a six-carbon molecule, citrate. This conden-
sation reaction is irreversible, committing the two-carbon
acetyl group to the Krebs cycle. The reaction is inhibited
when the cell’s ATP concentration is high and stimulated
when it is low. Hence, when the cell possesses ample
amounts of ATP, the Krebs cycle shuts down and acetyl-
CoA is channeled into fat synthesis.
Reactions 2 and 3: Isomerization. Before the oxidation
reactions can begin, the hydroxyl (—OH) group of citrate
must be repositioned. This is done in two steps: first, a
water molecule is removed from one carbon; then, water is
added to a different carbon. As a result, an —H group and
an —OH group change positions. The product is an iso-
mer of citrate called isocitrate.
Reaction 4: The First Oxidation. In the first energy-
yielding step of the cycle, isocitrate undergoes an oxidative
decarboxylation reaction. First, isocitrate is oxidized, yield-
ing a pair of electrons that reduce a molecule of NAD
+
to
NADH. Then the oxidized intermediate is decarboxylated;
the central carbon atom splits off to form CO
2
, yielding a
five-carbon molecule called α-ketoglutarate.
Reaction 5: The Second Oxidation. Next, α-ketoglutarate
is decarboxylated by a multienzyme complex similar to
pyruvate dehydrogenase. The succinyl group left after the
removal of CO
2
joins to coenzyme A, forming succinyl-
CoA. In the process, two electrons are extracted, and they
reduce another molecule of NAD
+
to NADH.
Reaction 6: Substrate-Level Phosphorylation. The
linkage between the four-carbon succinyl group and CoA is
a high-energy bond. In a coupled reaction similar to those
that take place in glycolysis, this bond is cleaved, and the
energy released drives the phosphorylation of guanosine
diphosphate (GDP), forming guanosine triphosphate
(GTP). GTP is readily converted into ATP, and the four-
carbon fragment that remains is called succinate.
Reaction 7: The Third Oxidation. Next, succinate is
oxidized to fumarate. The free energy change in this re-
action is not large enough to reduce NAD
+
. Instead,
flavin adenine dinucleotide (FAD) is the electron accep-
tor. Unlike NAD
+
, FAD is not free to diffuse within the
mitochondrion; it is an integral part of the inner mito-
chondrial membrane. Its reduced form, FADH
2
, con-
tributes electrons to the electron transport chain in the
membrane.
Reactions 8 and 9: Regeneration of Oxaloacetate. In
the final two reactions of the cycle, a water molecule is
added to fumarate, forming malate. Malate is then oxi-
dized, yielding a four-carbon molecule of oxaloacetate and
two electrons that reduce a molecule of NAD
+
to NADH.
Oxaloacetate, the molecule that began the cycle, is now
free to combine with another two-carbon acetyl group
from acetyl-CoA and reinitiate the cycle.
The Products of the Krebs Cycle
In the process of aerobic respiration, glucose is entirely
consumed. The six-carbon glucose molecule is first cleaved
into a pair of three-carbon pyruvate molecules during gly-
colysis. One of the carbons of each pyruvate is then lost as
CO
2
in the conversion of pyruvate to acetyl-CoA; two
other carbons are lost as CO
2
during the oxidations of the
Krebs cycle. All that is left to mark the passing of the glu-
cose molecule into six CO
2
molecules is its energy, some of
which is preserved in four ATP molecules and in the re-
duced state of 12 electron carriers. Ten of these carriers are
NADH molecules; the other two are FADH
2
.
The Krebs cycle generates two ATP molecules per
molecule of glucose, the same number generated by
glycolysis. More importantly, the Krebs cycle and the
oxidation of pyruvate harvest many energized electrons,
which can be directed to the electron transport chain to
drive the synthesis of much more ATP.
170 Part III Energetics
Chapter 9 How Cells Harvest Energy 171
Oxaloacetate (4C)
Krebs cycle
Mitochondrial
membrane
Oxidation of pyruvate
Pyruvate
Acetyl-CoA (2C)
CoA-SH
CoA-SH
Malate (4C)
Fumarate (4C)
Succinate (4C)
α-Ketoglutarate (5C)
Isocitrate (6C)
Citrate (6C)
Citrate
synthetase
Succinyl-CoA (4C)
The oxidation
of succinate
produces FADH
2
.
The dehydrogenation
of malate produces a
third NADH, and the
cycle returns to its
starting point.
A second oxidative
decarboxylation produces
a second NADH with the
release of a second CO
2
.
Oxidative
decarboxylation
produces NADH
with the release of
CO
2
.
The cycle begins
when a 2C unit from
acetyl-CoA reacts
with a 4C molecule
(oxaloacetate) to
produce citrate (6C).
H
2
O
NADH
NAD
+
NAD
+
NADH
NAD
+
NADH
NAD
+
NADH
FAD
FADH
2
CO
COO
–
COO
–
CH
2
C
S CoA
O
CH
3
COO
–
COO
–
CH
2
CH
2
CHHO
COO
–
COO
–
CH
2
CH
COO
–
COO
–
HC
COO
–
CO
S CoA
CH
2
CH
2
COO
–
COO
–
CO
CH
2
CH
2
COO
–
HC
CHHO
COO
–
COO
–
CH
2
COO
–
CHO
COO
–
COO
–
CH
2
CH
2
CO
2
CO
2
CO
2
1
Isocitrate
dehydrogenase
4
Malate
dehydrogenase
9
Succinate
dehydrogenase
7
Fumarase8
α-Ketoglutarate
dehydrogenase
Succinyl-CoA
synthetase
5
6
Aconitase
2
3
CoA-SH
GTP GDP + P
i
ATP
ADP
FIGURE 9.13
The Krebs cycle. This series of reactions takes place within the matrix of the mitochondrion. For the complete breakdown of a molecule
of glucose, the two molecules of acetyl-CoA produced by glycolysis and pyruvate oxidation will each have to make a trip around the Krebs
cycle. Follow the different carbons through the cycle, and notice the changes that occur in the carbon skeletons of the molecules as they
proceed through the cycle.
Harvesting Energy by Extracting
Electrons
To understand how cells direct some of the energy released
during glucose catabolism into ATP production, we need
to take a closer look at the electrons in the C—H bonds of
the glucose molecule. We stated in chapter 8 that when an
electron is removed from one atom and donated to an-
other, the electron’s potential energy of position is also
transferred. In this process, the atom that receives the elec-
tron is reduced. We spoke of reduction in an all-or-none
fashion, as if it involved the complete transfer of an elec-
tron from one atom to another. Often this is just what hap-
pens. However, sometimes a reduction simply changes the
degree of sharing within a covalent bond. Let us now revisit
that discussion and consider what happens when the trans-
fer of electrons is incomplete.
A Closer Look at Oxidation-Reduction
The catabolism of glucose is an oxidation-reduction reac-
tion. The covalent electrons in the C—H bonds of glucose
are shared approximately equally between the C and H
atoms because carbon and hydrogen nuclei have about the
same affinity for valence electrons (that is, they exhibit sim-
ilar electronegativity). However, when the carbon atoms of
glucose react with oxygen to form carbon dioxide, the elec-
trons in the new covalent bonds take a different position.
Instead of being shared equally, the electrons that were as-
sociated with the carbon atoms in glucose shift far toward
the oxygen atom in CO
2
because oxygen is very electroneg-
ative. Since these electrons are pulled farther from the car-
bon atoms, the carbon atoms of glucose have been oxidized
(loss of electrons) and the oxygen atoms reduced (gain of
electrons). Similarly, when the hydrogen atoms of glucose
combine with oxygen atoms to form water, the oxygen
atoms draw the shared electrons strongly toward them;
again, oxygen is reduced and glucose is oxidized. In this re-
action, oxygen is an oxidizing (electron-attracting) agent
because it oxidizes the atoms of glucose.
Releasing Energy
The key to understanding the oxidation of glucose is to
focus on the energy of the shared electrons. In a covalent
bond, energy must be added to remove an electron from an
atom, just as energy must be used to roll a boulder up a hill.
The more electronegative the atom, the steeper the energy
hill that must be climbed to pull an electron away from it.
However, energy is released when an electron is shifted
away from a less electronegative atom and closer to a more
electronegative atom, just as energy is released when a
boulder is allowed to roll down a hill. In the catabolism of
glucose, energy is released when glucose is oxidized, as
electrons relocate closer to oxygen (figure 9.14).
Glucose is an energy-rich food because it has an abun-
dance of C—H bonds. Viewed in terms of oxidation-
reduction, glucose possesses a wealth of electrons held far
from their atoms, all with the potential to move closer to-
ward oxygen. In oxidative respiration, energy is released
not simply because the hydrogen atoms of the C—H
bonds are transferred from glucose to oxygen, but be-
cause the positions of the valence electrons shift. This
shift releases energy that can be used to make ATP.
Harvesting the Energy in Stages
It is generally true that the larger the release of energy in
any single step, the more of that energy is released as heat
(random molecular motion) and the less there is available
to be channeled into more useful paths. In the combustion
of gasoline, the same amount of energy is released whether
all of the gasoline in a car’s gas tank explodes at once, or
whether the gasoline burns in a series of very small explo-
sions inside the cylinders. By releasing the energy in gaso-
line a little at a time, the harvesting efficiency is greater and
172 Part III Energetics
Electron
transport
chain
Low energy
High energy
Energy for
synthesis of
Electrons from food
Formation of water
ATP
e
–
e
–
FIGURE 9.14
How electron transport works. This diagram shows how ATP is
generated when electrons transfer from one energy level to
another. Rather than releasing a single explosive burst of energy,
electrons “fall” to lower and lower energy levels in steps, releasing
stored energy with each fall as they tumble to the lowest (most
electronegative) electron acceptor.
more of the energy can be used to push the pistons and
move the car.
The same principle applies to the oxidation of glucose
inside a cell. If all of the hydrogens were transferred to oxy-
gen in one explosive step, releasing all of the free energy at
once, the cell would recover very little of that energy in a
useful form. Instead, cells burn their fuel much as a car
does, a little at a time. The six hydrogens in the C—H
bonds of glucose are stripped off in stages in the series of
enzyme-catalyzed reactions collectively referred to as gly-
colysis and the Krebs cycle. We have had a great deal to say
about these reactions already in this chapter. Recall that the
hydrogens are removed by transferring them to a coenzyme
carrier, NAD
+
(figure 9.15). Discussed in chapter 8, NAD
+
is
a very versatile electron acceptor, shuttling energy-bearing
electrons throughout the cell. In harvesting the energy of
glucose, NAD
+
acts as the primary electron acceptor.
Following the Electrons
As you examine these reactions, try not to become confused
by the changes in electrical charge. Always follow the elec-
trons. Enzymes extract two hydrogens—that is, two elec-
trons and two protons—from glucose and transfer both
electrons and one of the protons to NAD
+
. The other pro-
ton is released as a hydrogen ion, H
+
, into the surrounding
solution. This transfer converts NAD
+
into NADH; that is,
two negative electrons and one positive proton are added to
one positively charged NAD
+
to form NADH, which is
electrically neutral.
Energy captured by NADH is not harvested all at once.
Instead of being transferred directly to oxygen, the two
electrons carried by NADH are passed along the electron
transport chain if oxygen is present. This chain consists of
a series of molecules, mostly proteins, embedded within the
inner membranes of mitochondria. NADH delivers elec-
trons to the top of the electron transport chain and oxygen
captures them at the bottom. The oxygen then joins with
hydrogen ions to form water. At each step in the chain, the
electrons move to a slightly more electronegative carrier,
and their positions shift slightly. Thus, the electrons move
down an energy gradient. The entire process releases a total
of 53 kcal/mole (222 kJ/mole) under standard conditions.
The transfer of electrons along this chain allows the energy
to be extracted gradually. In the next section, we will dis-
cuss how this energy is put to work to drive the production
of ATP.
The catabolism of glucose involves a series of
oxidation-reduction reactions that release energy by
repositioning electrons closer to oxygen atoms. Energy
is thus harvested from glucose molecules in gradual
steps, using NAD
+
as an electron carrier.
Chapter 9 How Cells Harvest Energy 173
N
+
OCH
2
H
P
O
O
O
H
O
OP
H
-
O
-
O
O
P
-
O
O
HO OH
O
CH
2
H
HO OH HO OH
O
CNH
2
+ 2H
Reduction
Oxidation
NAD
+
: oxidized form of nicotinamide
Adenine
N
N
N
N H
NH
2
N
OCH
2
H
P
O
-
O
O
O
HHO
O
H
HO OH
O
CH
2
H
O
CNH
2
+ H
+
NADH: reduced form of nicotinamide
Adenine
N
N
N
N H
NH
2
FIGURE 9.15
NAD
+
and NADH.
This dinucleotide
serves as an “electron
shuttle” during
cellular respiration.
NAD
+
accepts
electrons from
catabolized
macromolecules and
is reduced to NADH.
Stage Four: The Electron
Transport Chain
The NADH and FADH
2
molecules formed during the first
three stages of aerobic respiration each contain a pair of
electrons that were gained when NAD
+
and FAD were re-
duced. The NADH molecules carry their electrons to the
inner mitochondrial membrane, where they transfer the
electrons to a series of membrane-associated proteins col-
lectively called the electron transport chain.
Moving Electrons through the Electron
Transport Chain
The first of the proteins to receive the electrons is a com-
plex, membrane-embedded enzyme called NADH dehy-
drogenase. A carrier called ubiquinone then passes the
electrons to a protein-cytochrome complex called the bc
1
complex. This complex, along with others in the chain, oper-
ates as a proton pump, driving a proton out across the mem-
brane. Cytochromes are respiratory proteins that contain
heme groups, complex carbon rings with many alternating
single and double bonds and an iron atom in the center.
The electron is then carried by another carrier, cy-
tochrome c, to the cytochrome oxidase complex. This com-
plex uses four such electrons to reduce a molecule of oxy-
gen, each oxygen then combines with two hydrogen ions to
form water:
O
2
+ 4 H
+
+ 4 e
-
?→ 2 H
2
O
This series of membrane-associated electron carriers is
collectively called the electron transport chain (figure 9.16).
NADH contributes its electrons to the first protein of
the electron transport chain, NADH dehydrogenase.
FADH
2
, which is always attached to the inner mitochon-
drial membrane, feeds its electrons into the electron trans-
port chain later, to ubiquinone.
It is the availability of a plentiful electron acceptor
(often oxygen) that makes oxidative respiration possible. As
we’ll see in chapter 10, the electron transport chain used in
aerobic respiration is similar to, and may well have evolved
from, the chain employed in aerobic photosynthesis.
The electron transport chain is a series of five
membrane-associated proteins. Electrons delivered by
NADH and FADH
2
are passed from protein to protein
along the chain, like a baton in a relay race.
174 Part III Energetics
Intermembrane space
Mitochondrial matrix
Inner
mitochondrial
membrane
NAD
+
Q
C
NADH
H
2
O
2H
+
+ O
2
+ H
+
H
+
H
+
H
+
NADH
dehydrogenase
bc
1
complex
Cytochrome
oxidase complex
FADH
2
H11002
1
2
FIGURE 9.16
The electron transport chain. High-energy electrons harvested from catabolized molecules are transported (red arrows) by mobile
electron carriers (ubiquinone, marked Q, and cytochrome c, marked C) along a chain of membrane proteins. Three proteins use portions
of the electrons’ energy to pump protons (blue arrows) out of the matrix and into the intermembrane space. The electrons are finally
donated to oxygen to form water.
Building an Electrochemical Gradient
In eukaryotes, aerobic metabolism takes place within the
mitochondria present in virtually all cells. The internal
compartment, or matrix, of a mitochondrion contains the
enzymes that carry out the reactions of the Krebs cycle.
As the electrons harvested by oxidative respiration are
passed along the electron transport chain, the energy they
release transports protons out of the matrix and into the
outer compartment, sometimes called the intermembrane
space. Three transmembrane proteins in the inner mito-
chondrial membrane (see figure 9.16) actually accomplish
the transport. The flow of excited electrons induces a
change in the shape of these pump proteins, which causes
them to transport protons across the membrane. The
electrons contributed by NADH activate all three of these
proton pumps, while those contributed by FADH
2
acti-
vate only two.
Producing ATP: Chemiosmosis
As the proton concentration in the intermembrane space
rises above that in the matrix, the matrix becomes slightly
negative in charge. This internal negativity attracts the
positively charged protons and induces them to reenter the
matrix. The higher outer concentration tends to drive pro-
tons back in by diffusion; since membranes are relatively
impermeable to ions, most of the protons that reenter the
matrix pass through special proton
channels in the inner mitochondrial
membrane. When the protons pass
through, these channels synthesize
ATP from ADP + P
i
within the ma-
trix. The ATP is then transported by
facilitated diffusion out of the mito-
chondrion and into the cell’s cyto-
plasm. Because the chemical forma-
tion of ATP is driven by a diffusion
force similar to osmosis, this process
is referred to as chemiosmosis (fig-
ure 9.17).
Thus, the electron transport chain
uses electrons harvested in aerobic
respiration to pump a large number of
protons across the inner mitochon-
drial membrane. Their subsequent
reentry into the mitochondrial matrix
drives the synthesis of ATP by
chemiosmosis. Figure 9.18 summa-
rizes the overall process.
The electrons harvested from
glucose are pumped out of the
mitochondrial matrix by the
electron transport chain. The
return of the protons into the
matrix generates ATP.
Chapter 9 How Cells Harvest Energy 175
Intermembrane space
Mitochondrial matrix
Na-K
pump
ATP
ADP + P
i
NADH
H
+
H
+
H
+
H
+
H
+
H
+
NAD
+
Inner
mitochondrial
membrane
Proton pump ATP
synthase
FIGURE 9.17
Chemiosmosis. NADH transports high-energy electrons
harvested from the catabolism of macromolecules to “proton
pumps” that use the energy to pump protons out of the
mitochondrial matrix. As a result, the concentration of protons in
the intermembrane space rises, inducing protons to diffuse back
into the matrix. Many of the protons pass through special channels
that couple the reentry of protons to the production of ATP.
H
+
H
+
H
+
H
+
O
2
O
2
+
Q
C
ATP32
ATP2
Pyruvate from
cytoplasm
Electron
transport
system
Channel
protein
H
2
O
CO
2
Acetyl-CoA
Krebs
cycle
FADH
2
NADH
NADH
Intermembrane
space
Mitochondrial matrix
Inner
mitochondrial
membrane
1. Electrons are harvested
and carried to the
transport system.
2. Electrons provide
energy to pump
protons across the
membrane.
3. Oxygen joins with
protons to form water.
4. Protons diffuse back in,
driving the synthesis of
ATP.
2H
+
H11002
1
2
FIGURE 9.18
ATP generation during the Krebs cycle and electron transport chain. This process
begins with pyruvate, the product of glycolysis, and ends with the synthesis of ATP.
Summarizing Aerobic
Respiration
How much metabolic energy does a cell
actually gain from the electrons har-
vested from a molecule of glucose, using
the electron transport chain to produce
ATP by chemiosmosis?
Theoretical Yield
The chemiosmotic model suggests that
one ATP molecule is generated for
each proton pump activated by the
electron transport chain. Since the elec-
trons from NADH activate three
pumps and those from FADH
2
activate
two, we would expect each molecule of
NADH and FADH
2
to generate three
and two ATP molecules, respectively.
However, because eukaryotic cells carry
out glycolysis in their cytoplasm and
the Krebs cycle within their mitochon-
dria, they must transport the two mole-
cules of NADH produced during gly-
colysis across the mitochondrial
membranes, which requires one ATP
per molecule of NADH. Thus, the net
ATP production is decreased by two.
Therefore, the overall ATP production
resulting from aerobic respiration theo-
retically should be 4 (from substrate-
level phosphorylation during glycolysis)
+ 30 (3 from each of 10 molecules of
NADH) + 4 (2 from each of 2 mole-
cules of FADH
2
) – 2 (for transport of
glycolytic NADH) = 36 molecules of ATP (figure 9.19).
Actual Yield
The amount of ATP actually produced in a eukaryotic cell
during aerobic respiration is somewhat lower than 36, for
two reasons. First, the inner mitochondrial membrane is
somewhat “leaky” to protons, allowing some of them to
reenter the matrix without passing through ATP-generating
channels. Second, mitochondria often use the proton gra-
dient generated by chemiosmosis for purposes other than
ATP synthesis (such as transporting pyruvate into the ma-
trix). Consequently, the actual measured values of ATP
generated by NADH and FADH
2
are closer to 2.5 for
each NADH and 1.5 for each FADH
2
. With these correc-
tions, the overall harvest of ATP from a molecule of glu-
cose in a eukaryotic cell is closer to 4 (from substrate-level
phosphorylation) + 25 (2.5 from each of 10 molecules of
NADH) + 3 (1.5 from each of 2 molecules of FADH
2
) – 2
(transport of glycolytic NADH) = 30 molecules of ATP.
The catabolism of glucose by aerobic respiration, in
contrast to glycolysis, is quite efficient. Aerobic respiration
in a eukaryotic cell harvests about 7.3 × 30 ÷ 686 = 32% of
the energy available in glucose. (By comparison, a typical
car converts only about 25% of the energy in gasoline into
useful energy.) The efficiency of oxidative respiration at
harvesting energy establishes a natural limit on the maxi-
mum length of food chains.
The high efficiency of aerobic respiration was one of the
key factors that fostered the evolution of heterotrophs. With
this mechanism for producing ATP, it became feasible for
nonphotosynthetic organisms to derive metabolic energy ex-
clusively from the oxidative breakdown of other organisms.
As long as some organisms captured energy by photosynthe-
sis, others could exist solely by feeding on them.
Oxidative respiration produces approximately 30
molecules of ATP from each molecule of glucose in
eukaryotic cells. This represents more than half of the
energy in the chemical bonds of glucose.
176 Part III Energetics
Glycolysis2
26ATP
Pyruvate
Glucose
Acetyl-CoA
NADH
24ATPNADH
2 ATP
2 ATP
618ATPNADH
24ATP
Total net ATP yield = 36 ATP
FADH
2
Krebs
cycle
ATP
FIGURE 9.19
Theoretical ATP yield. The theoretical yield of ATP harvested from glucose by aerobic
respiration totals 36 molecules.
Regulating Aerobic Respiration
When cells possess plentiful amounts of ATP, the key reac-
tions of glycolysis, the Krebs cycle, and fatty acid break-
down are inhibited, slowing ATP production. The regula-
tion of these biochemical pathways by the level of ATP is
an example of feedback inhibition. Conversely, when ATP
levels in the cell are low, ADP levels are high; and ADP ac-
tivates enzymes in the pathways of carbohydrate catabolism
to stimulate the production of more ATP.
Control of glucose catabolism occurs at two key points
of the catabolic pathway (figure 9.20). The control point
in glycolysis is the enzyme phosphofructokinase, which
catalyzes reaction 3, the conversion of fructose phosphate
to fructose bisphosphate. This is the first reaction of gly-
colysis that is not readily reversible, committing the sub-
strate to the glycolytic sequence. High levels of ADP rela-
tive to ATP (implying a need to convert more ADP to
ATP) stimulate phosphofructokinase, committing more
sugar to the catabolic pathway; so do low levels of citrate
(implying the Krebs cycle is not running at full tilt and
needs more input). The main control point in the oxida-
tion of pyruvate occurs at the committing step in the
Krebs cycle with the enzyme pyruvate decarboxylase. It is
inhibited by high levels of NADH (implying no more is
needed).
Another control point in the Krebs cycle is the enzyme
citrate synthetase, which catalyzes the first reaction, the
conversion of oxaloacetate and acetyl-CoA into citrate.
High levels of ATP inhibit citrate synthetase (as well as
pyruvate decarboxylase and two other Krebs cycle en-
zymes), shutting down the catabolic pathway.
Relative levels of ADP and ATP regulate the catabolism
of glucose at key committing reactions.
Chapter 9 How Cells Harvest Energy 177
A Vocabulary of
ATP Generation
which the final electron acceptor is or-
ganic; it includes aerobic and anaerobic
respiration.
chemiosmosis The passage of high-
energy electrons along the electron trans-
port chain, which is coupled to the pump-
ing of protons across a membrane and the
return of protons to the original side of
the membrane through ATP-generating
channels.
fermentation Alternative ATP-producing
pathway performed by some cells in the ab-
sence of oxygen, in which the final electron
acceptor is an organic molecule.
maximum efficiency The maximum
number of ATP molecules generated by
oxidizing a substance, relative to the free
energy of that substance; in organisms, the
actual efficiency is usually less than the
maximum.
oxidation The loss of an electron. In cel-
lular respiration, high-energy electrons are
stripped from food molecules, oxidizing
them.
photosynthesis The chemiosmotic gen-
eration of ATP and complex organic mole-
cules powered by the energy derived from
light.
substrate-level phosphorylation The
generation of ATP by the direct transfer of
a phosphate group to ADP from another
phosphorylated molecule.
aerobic respiration The portion of cellu-
lar respiration that requires oxygen as an
electron acceptor; it includes pyruvate oxi-
dation, the Krebs cycle, and the electron
transport chain.
anaerobic respiration Cellular respiration
in which inorganic electron acceptors other
than oxygen are used; it includes glycolysis.
cellular respiration The oxidation of
organic molecules to produce ATP in
ATP
Krebs
cycle
Electron transport
chain and
chemiosmosis
NADH
Citrate
Glucose ADP
Fructose 6-phosphate
Fructose 1,6-bisphosphate
Pyruvate
Pyruvate decarboxylase
Acetyl-CoA
Phosphofructokinase
Inhibits
Activates
Activates
Inhibits
FIGURE 9.20
Control of glucose catabolism. The relative levels of ADP and
ATP control the catabolic pathway at two key points: the
committing reactions of glycolysis and the Krebs cycle.
Glucose Is Not the Only
Source of Energy
Thus far we have discussed oxidative
respiration of glucose, which organisms
obtain from the digestion of carbohy-
drates or from photosynthesis. Other
organic molecules than glucose, partic-
ularly proteins and fats, are also impor-
tant sources of energy (figure 9.21).
Cellular Respiration of Protein
Proteins are first broken down into their
individual amino acids. The nitrogen-
containing side group (the amino
group) is then removed from each
amino acid in a process called deamina-
tion. A series of reactions convert the
carbon chain that remains into a mole-
cule that takes part in glycolysis or the
Krebs cycle. For example, alanine is
converted into pyruvate, glutamate into
α-ketoglutarate (figure 9.22), and aspar-
tate into oxaloacetate. The reactions of
glycolysis and the Krebs cycle then ex-
tract the high-energy electrons from
these molecules and put them to work
making ATP.
178 Part III Energetics
9.3 Catabolism of proteins and fats can yield considerable energy.
Macromolecule
degradation
Cell
building blocks
Nucleic
acids
Proteins
Lipids
and fats
Polysaccharides
Nucleotides Amino acids Fatty acidsSugars
NH
3
H
2
OCO
2
Deamination H9252-oxidationGlycolysis
Oxidative
respiration
Ultimate
metabolic
products
Pyruvate
Acetyl-CoA
Krebs
cycle
FIGURE 9.21
How cells extract chemical energy. All eukaryotes and many prokaryotes extract energy
from organic molecules by oxidizing them. The first stage of this process, breaking down
macromolecules into their constituent parts, yields little energy. The second stage,
oxidative or aerobic respiration, extracts energy, primarily in the form of high-energy
electrons, and produces water and carbon dioxide.
H9251-Ketoglutarate
C
—
— —
O
O
H — C —
—
——
H — C —
HO
—
—
—
C
—
— —
HO
— C — H
H
H
Glutamate
H
2
N
C
—
— —
O
O
O
H — C —
—
—
——
H — C —
—
—
—
C
—
— —
HO
HO
C
H
H
NH
2
Urea
FIGURE 9.22
Deamination. After
proteins are broken down
into their amino acid
constituents, the amino
groups are removed from
the amino acids to form
molecules that participate in
glycolysis and the Krebs
cycle. For example, the
amino acid glutamate
becomes α-ketoglutarate, a
Krebs cycle molecule, when
it loses its amino group.
Cellular Respiration of Fat
Fats are broken down into fatty acids plus glycerol. The
tails of fatty acids typically have 16 or more —CH
2
links,
and the many hydrogen atoms in these long tails provide a
rich harvest of energy. Fatty acids are oxidized in the ma-
trix of the mitochondrion. Enzymes there remove the two-
carbon acetyl groups from the end of each fatty acid tail
until the entire fatty acid is converted into acetyl groups
(figure 9.23). Each acetyl group then combines with coen-
zyme A to form acetyl-CoA. This process is known as H9252-
oxidation.
How much ATP does the catabolism of fatty acids pro-
duce? Let’s compare a hypothetical six-carbon fatty acid
with the six-carbon glucose molecule, which we’ve said
yields about 30 molecules of ATP in a eukaryotic cell. Two
rounds of β-oxidation would convert the fatty acid into
three molecules of acetyl-CoA. Each round requires one
molecule of ATP to prime the process, but it also produces
one molecule of NADH and one of FADH
2
. These mole-
cules together yield four molecules of ATP (assuming 2.5
ATPs per NADH and 1.5 ATPs per FADH
2
). The oxida-
tion of each acetyl-CoA in the Krebs cycle ultimately pro-
duces an additional 10 molecules of ATP. Overall, then,
the ATP yield of a six-carbon fatty acid would be approxi-
mately 8 (from two rounds of β-oxidation) – 2 (for priming
those two rounds) + 30 (from oxidizing the three acetyl-
CoAs) = 36 molecules of ATP. Therefore, the respiration
of a six-carbon fatty acid yields 20% more ATP than the
respiration of glucose. Moreover, a fatty acid of that size
would weigh less than two-thirds as much as glucose, so a
gram of fatty acid contains more than twice as many kilo-
calories as a gram of glucose. That is why fat is a storage
molecule for excess energy in many types of animals. If ex-
cess energy were stored instead as carbohydrate, as it is in
plants, animal bodies would be much bulkier.
Proteins, fats, and other organic molecules are also
metabolized for energy. The amino acids of proteins are
first deaminated, while fats undergo a process called
β-oxidation.
Chapter 9 How Cells Harvest Energy 179
OH
O
Fatty acid C
H
H
C
H
H
C
Fatty acid C
H
H
C
H
H
C
O
FAD
FADH
2
ATP
AMP + PP
i
CoA
CoAFatty acid C
H
C
H
C
O
CoA
Fatty acid C
HO
H
C
H
H
C
O
CoA
Fatty acid C C
H
H
C
OO
CoA
NAD
+
NADH
H
2
O
Acetyl-CoA
Krebs
cycle
CoA
FIGURE 9.23
β-oxidation. Through a series of reactions known as β-oxidation, the
last two carbons in a fatty acid tail combine with coenzyme A to form
acetyl-CoA, which enters the Krebs cycle. The fatty acid, now two
carbons shorter, enters the pathway again and keeps reentering until
all its carbons have been used to form acetyl-CoA molecules. Each
round of β-oxidation uses one molecule of ATP and generates one
molecule each of FADH
2
and NADH, not including the molecules
generated from the Krebs cycle.
180 Part III Energetics
Even with oxidative metabolism, approx-
imately two-thirds of the available energy is
lost at each trophic level, and that puts a
limit on how long a food chain can be. Most
food chains, like the one illustrated in figure
9.A, involve only three or rarely four trophic
levels. Too much energy is lost at each
transfer to allow chains to be much longer
than that. For example, it would be impossi-
ble for a large human population to subsist
by eating lions captured from the Serengeti
Plain of Africa; the amount of grass available
there would not support enough zebras and
other herbivores to maintain the number of
lions needed to feed the human population.
Thus, the ecological complexity of our
world is fixed in a fundamental way by the
chemistry of oxidative respiration.
When organisms became able to extract
energy from organic molecules by oxidative
metabolism, this constraint became far less
severe, because the efficiency of oxidative
respiration is estimated to be about 52 to
63%. This increased efficiency results in
the transmission of much more energy
from one trophic level to another than does
glycolysis. (A trophic level is a step in the
movement of energy through an ecosys-
tem.) The efficiency of oxidative metabo-
lism has made possible the evolution of
food chains, in which autotrophs are con-
sumed by heterotrophs, which are con-
sumed by other heterotrophs, and so on.
You will read more about food chains in
chapter 28.
It has been estimated that a heterotroph
limited to glycolysis captures only 3.5% of
the energy in the food it consumes. Hence,
if such a heterotroph preserves 3.5% of the
energy in the autotrophs it consumes, then
any other heterotrophs that consume the
first hetertroph will capture through glycol-
ysis 3.5% of the energy in it, or 0.12% of
the energy available in the original au-
totrophs. A very large base of autotrophs
would thus be needed to support a small
number of heterotrophs.
Metabolic Efficiency
and the Length of
Food Chains
FIGURE 9.A
A food chain in the savannas, or open grasslands, of East Africa. Stage 1: Photosynthesizer. The grass under these palm trees grows
actively during the hot, rainy season, capturing the energy of the sun and storing it in molecules of glucose, which are then converted into
starch and stored in the grass. Stage 2: Herbivore. These large antelopes, known as wildebeests, consume the grass and transfer some of its
stored energy into their own bodies. Stage 3: Carnivore. The lion feeds on wildebeests and other animals, capturing part of their stored
energy and storing it in its own body. Stage 4: Scavenger. This hyena and the vultures occupy the same stage in the food chain as the lion.
They are also consuming the body of the dead wildebeest, which has been abandoned by the lion. Stage 5: Refuse utilizer. These butterflies,
mostly Precis octavia, are feeding on the material left in the hyena’s dung after the food the hyena consumed had passed through its
digestive tract. At each of these four levels, only about a third or less of the energy present is used by the recipient.
Stage 1: Photosynthesizers Stage 2: Herbivores
Stage 3: Carnivore Stage 4: Scavengers Stage 5: Refuse utilizers
Fermentation
In the absence of oxygen, aerobic metabolism cannot
occur, and cells must rely exclusively on glycolysis to pro-
duce ATP. Under these conditions, the hydrogen atoms
generated by glycolysis are donated to organic molecules in
a process called fermentation.
Bacteria carry out more than a dozen kinds of fermen-
tations, all using some form of organic molecule to ac-
cept the hydrogen atom from NADH and thus recycle
NAD
+
:
Organic molecule
?→
Reduced organic molecule
+ NADH + NAD
+
Often the reduced organic compound is an organic acid—
such as acetic acid, butyric acid, propionic acid, or lactic
acid—or an alcohol.
Ethanol Fermentation
Eukaryotic cells are capable of only a few types of fermen-
tation. In one type, which occurs in single-celled fungi
called yeast, the molecule that accepts hydrogen from
NADH is pyruvate, the end product of glycolysis itself.
Yeast enzymes remove a terminal CO
2
group from pyru-
vate through decarboxylation, producing a two-carbon
molecule called acetaldehyde. The CO
2
released causes
bread made with yeast to rise, while bread made without
yeast (unleavened bread) does not. The acetaldehyde ac-
cepts a hydrogen atom from NADH, producing NAD
+
and
ethanol (ethyl alcohol). This particular type of fermenta-
tion is of great interest to humans, since it is the source of
the ethanol in wine and beer (figure 9.24). Ethanol is a by-
product of fermentation that is actually toxic to yeast; as it
approaches a concentration of about 12%, it begins to kill
the yeast. That explains why naturally fermented wine con-
tains only about 12% ethanol.
Lactic Acid Fermentation
Most animal cells regenerate NAD
+
without decarboxyla-
tion. Muscle cells, for example, use an enzyme called lactate
dehydrogenase to transfer a hydrogen atom from NADH
back to the pyruvate that is produced by glycolysis. This
reaction converts pyruvate into lactic acid and regenerates
NAD
+
from NADH. It therefore closes the metabolic cir-
cle, allowing glycolysis to continue as long as glucose is
available. Circulating blood removes excess lactate (the ion-
ized form of lactic acid) from muscles, but when removal
cannot keep pace with production, the accumulating lactic
acid interferes with muscle function and contributes to
muscle fatigue.
In fermentation, which occurs in the absence of oxygen,
the electrons that result from the glycolytic breakdown
of glucose are donated to an organic molecule,
regenerating NAD
+
from NADH.
Chapter 9 How Cells Harvest Energy 181
9.4 Cells can metabolize food without oxygen.
Glucose
2 Pyruvate
Alcohol fermentation in yeast
G
L
Y
C
O
L
Y
S
I
S
2 ATP
2 ADP
CO
O
-
CO
CH
3
CO
O
–
CH
3
C
2 Ethanol
2 Acetaldehyde
2 Lactate
HOH
H
CH OH
CH
3
Glucose
2 Pyruvate
Lactic acid fermentation in muscle cells
G
L
Y
C
O
L
Y
S
I
S
2 ATP
2 ADP
CO
O
-
CO
CH
3
2 NAD
+
OC
H
CH
3
2 NADH
2 NAD
+
2 NADH
CO
2
FIGURE 9.24
How wine is made. Yeasts carry out the conversion of pyruvate to
ethanol. This takes place naturally in grapes left to ferment on
vines, as well as in fermentation vats containing crushed grapes.
When the ethanol concentration reaches about 12%, its toxic
effects kill the yeast; what remains is wine. Muscle cells convert
pyruvate into lactate, which is less toxic than ethanol. However,
lactate is still toxic enough to produce a painful sensation in
muscles during heavy exercise, when oxygen in the muscles is
depleted.
182 Part III Energetics
Chapter 9
Summary Questions Media Resources
9.1 Cells harvest the energy in chemical bonds.
? The reactions of cellular respiration are oxidation-
reduction (redox) reactions. Those that require a net
input of free energy are coupled to the cleavage of
ATP, which releases free energy.
? The mechanics of cellular respiration are often
dictated by electron behavior, which is in turn
influenced by the presence of electron acceptors.
Some atoms, such as oxygen, are very electronegative
and thus behave as good oxidizing agents.
1. What is the difference
between an autotroph and a
heterotroph? How does each
obtain energy?
2. What is the difference
between digestion and
catabolism? Which provides
more energy?
? In eukaryotic cells, the oxidative respiration of
pyruvate takes place within the matrix of
mitochondria.
? The electrons generated in the process are passed
along the electron transport chain, a sequence of
electron carriers in the inner mitochondrial
membrane.
? Some of the energy released by passage of electrons
along the electron transport chain is used to pump
protons out of the mitochondrial matrix. The reentry
of protons into the matrix is coupled to the
production of ATP. This process is called
chemiosmosis. The ATP then leaves the
mitochondrion by facilitated diffusion.
3. Where in a eukaryotic cell
does glycolysis occur? What is
the net production of ATP
during glycolysis, and why is it
different from the number of
ATP molecules synthesized
during glycolysis?
4. By what two mechanisms can
the NADH that results from
glycolysis be converted back into
NAD
+
?
5. What is the theoretical
maximum number of ATP
molecules produced during the
oxidation of a glucose molecule
by these processes? Why is the
actual number of ATP molecules
produced usually lower than the
theoretical maximum?
9.2 Cellular respiration oxidizes food molecules.
? The catabolism of fatty acids begins with H9252-oxidation
and provides more energy than the catabolism of
carbohydrates.
6. How is acetyl-CoA produced
during the aerobic oxidation of
carbohydrates, and what happens
to it? How is it produced during
the aerobic oxidation of fatty
acids, and what happens to it
then?
9.3 Catabolism of proteins and fats can yield considerable energy.
? Fermentation is an anaerobic process that uses an
organic molecule instead of oxygen as a final electron
acceptor.
? It occurs in bacteria as well as eukaryotic cells,
including yeast and the muscle cells of animals.
7. How do the amounts of ATP
produced by the aerobic
oxidation of glucose and fatty
acids compare? Which type of
substance contains more energy
on a per-weight basis?
9.4 Cells can metabolize food without oxygen.
http://www.mhhe.com/raven6e http://www.biocourse.com
? Exploration:
Oxidative Respiration
? Art Activity: Aerobic
Cellular Respiration
? Art Activity:
Organization of
Cristae
? Electron Transport
and ATP
? Glycolysis
? Krebs Cycle
? Electron Transport
? Fermentation
183
10
Photosynthesis
Concept Outline
10.1 What is photosynthesis?
The Chloroplast as a Photosynthetic Machine. The
highly organized system of membranes in chloroplasts is
essential to the functioning of photosynthesis.
10.2 Learning about photosynthesis: An experimental
journey.
The Role of Soil and Water. The added mass of a
growing plant comes mostly from photosynthesis. In plants,
water supplies the electrons used to reduce carbon dioxide.
Discovery of the Light-Independent Reactions.
Photosynthesis is a two-stage process. Only the first stage
directly requires light.
The Role of Light. The oxygen released during green
plant photosynthesis comes from water, and carbon atoms
from carbon dioxide are incorporated into organic molecules.
The Role of Reducing Power. Electrons released from
the splitting of water reduce NADP
+
; ATP and NADPH
are then used to reduce CO
2
and form simple sugars.
10.3 Pigments capture energy from sunlight.
The Biophysics of Light. The energy in sunlight occurs
in “packets” called photons, which are absorbed by pigments.
Chlorophylls and Carotenoids. Photosynthetic
pigments absorb light and harvest its energy.
Organizing Pigments into Photosystems. A
photosystem uses light energy to eject an energized electron.
How Photosystems Convert Light to Chemical Energy.
Some bacteria rely on a single photosystem to produce
ATP. Plants use two photosystems in series to generate
enough energy to reduce NADP
+
and generate ATP.
How the Two Photosystems of Plants Work Together.
Photosystems II and I drive the synthesis of the ATP and
NADPH needed to form organic molecules.
10.4 Cells use the energy and reducing power captured
by the light reactions to make organic molecules.
The Calvin Cycle. ATP and NADPH are used to build
organic molecules, a process reversed in mitochondria.
Reactions of the Calvin Cycle. Ribulose bisphosphate
binds CO
2
in the process of carbon fixation.
Photorespiration. The enzyme that catalyzes carbon
fixation also affects CO
2
release.
L
ife on earth would be impossible without photosyn-
thesis. Every oxygen atom in the air we breathe was
once part of a water molecule, liberated by photosynthesis.
The energy released by the burning of coal, firewood,
gasoline, and natural gas, and by our bodies’ burning of all
the food we eat—all, directly or indirectly, has been cap-
tured from sunlight by photosynthesis. It is vitally impor-
tant that we understand photosynthesis. Research may en-
able us to improve crop yields and land use, important
goals in an increasingly crowded world. In the previous
chapter we described how cells extract chemical energy
from food molecules and use that energy to power their
activities. In this chapter, we will examine photosynthesis,
the process by which organisms capture energy from sun-
light and use it to build food molecules rich in chemical
energy (figure 10.1).
FIGURE 10.1
Capturing energy. These sunflower plants, growing vigorously
in the August sun, are capturing light energy for conversion into
chemical energy through photosynthesis.
184 Part III Energetics
10.1 What is photosynthesis?
Cuticle
Epidermis
Mesophyll
Vascular
bundle
Stoma
Bundle
sheath
Chloroplasts
Vacuole
Nucleus
Cell wall
Outer
membrane
Inner
membrane
Stroma
Granum
Thylakoid
The Chloroplast as a
Photosynthetic Machine
Life is powered by sunshine. The energy used by most liv-
ing cells comes ultimately from the sun, captured by plants,
algae, and bacteria through the process of photosynthesis.
The diversity of life is only possible because our planet is
awash in energy streaming earthward from the sun. Each
day, the radiant energy that reaches the earth equals about
1 million Hiroshima-sized atomic bombs. Photosynthesis
captures about 1% of this huge supply of energy, using it to
provide the energy that drives all life.
The Photosynthetic Process: A Summary
Photosynthesis occurs in many kinds of bacteria and algae,
and in the leaves and sometimes the stems of green plants.
Figure 10.2 describes the levels of organization in a plant
leaf. Recall from chapter 5 that the cells of plant leaves
contain organelles called chloroplasts that actually carry
out the photosynthetic process. No other structure in a
plant cell is able to carry out photosynthesis. Photosynthe-
FIGURE 10.2
Journey into a leaf. A plant leaf possesses a thick layer of cells (the mesophyll) rich in chloroplasts. The flattened thylakoids in the
chloroplast are stacked into columns called grana (singular, granum). The light reactions take place on the thylakoid
sis takes place in three stages: (1) capturing energy from
sunlight; (2) using the energy to make ATP and reducing
power in the form of a compound called NADPH; and
(3) using the ATP and NADPH to power the synthesis of
organic molecules from CO
2
in the air (carbon fixation).
The first two stages take place in the presence of light
and are commonly called the light reactions. The third
stage, the formation of organic molecules from atmos-
pheric CO
2
, is called the Calvin cycle. As long as ATP and
NADPH are available, the Calvin cycle may occur in the
absence of light.
The following simple equation summarizes the overall
process of photosynthesis:
6 CO
2
+ 12 H
2
O + light —→ C
6
H
12
O
6
+ 6 H
2
O + 6 O
2
carbon water glucose water oxygen
dioxide
Inside the Chloroplast
The internal membranes of chloroplasts are organized into
sacs called thylakoids, and often numerous thylakoids are
stacked on one another in columns called grana. The thy-
lakoid membranes house the photosynthetic pigments for
capturing light energy and the machinery to make ATP.
Surrounding the thylakoid membrane system is a semiliq-
uid substance called stroma. The stroma houses the en-
zymes needed to assemble carbon molecules. In the mem-
branes of thylakoids, photosynthetic pigments are clustered
together to form a photosystem.
Each pigment molecule within the photosystem is capa-
ble of capturing photons, which are packets of energy. A lat-
tice of proteins holds the pigments in close contact with
one another. When light of a proper wavelength strikes a
pigment molecule in the photosystem, the resulting excita-
tion passes from one chlorophyll molecule to another. The
excited electron is not transferred physically—it is the en-
ergy that passes from one molecule to another. A crude
analogy to this form of energy transfer is the initial “break”
in a game of pool. If the cue ball squarely hits the point of
the triangular array of 15 pool balls, the two balls at the far
corners of the triangle fly off, but none of the central balls
move. The energy passes through the central balls to the
most distant ones.
Eventually the energy arrives at a key chlorophyll mole-
cule that is touching a membrane-bound protein. The en-
ergy is transferred as an excited electron to that protein,
which passes it on to a series of other membrane proteins
that put the energy to work making ATP and NADPH and
building organic molecules. The photosystem thus acts as a
large antenna, gathering the light harvested by many indi-
vidual pigment molecules.
The reactions of photosynthesis take place within
thylakoid membranes within chloroplasts in leaf cells.
Chapter 10 Photosynthesis 185
Sunlight
Light reactions
Organic
molecules
CO
2
H
2
O
O
2
Photosystem
ADP
NADPH NADP
+
Stroma
Thylakoid
Thylakoid
Stroma
Granum
ATP
Calvin
cycle
FIGURE 10.2 (continued)
membrane and generate the ATP and NADPH that fuel the Calvin cycle. The fluid interior matrix of a chloroplast, the stroma, contains the
enzymes that carry out the Calvin cycle.
The Role of Soil and Water
The story of how we learned about photosynthesis is one of
the most interesting in science and serves as a good intro-
duction to this complex process. The story starts over 300
years ago, with a simple but carefully designed experiment
by a Belgian doctor, Jan Baptista van Helmont
(1577–1644). From the time of the Greeks, plants were
thought to obtain their food from the soil, literally sucking
it up with their roots; van Helmont thought of a simple
way to test the idea. He planted a small willow tree in a pot
of soil after weighing the tree and the soil. The tree grew in
the pot for several years, during which time van Helmont
added only water. At the end of five years, the tree was
much larger: its weight had increased by 74.4 kilograms.
However, all of this added mass could not have come from the
soil, because the soil in the pot weighed only 57 grams less
than it had five years earlier! With this experiment, van
Helmont demonstrated that the substance of the plant was
not produced only from the soil. He incorrectly concluded
that mainly the water he had been adding accounted for the
plant’s increased mass.
A hundred years passed before the story became clearer.
The key clue was provided by the English scientist Joseph
Priestly, in his pioneering studies of the properties of air.
On the 17th of August, 1771, Priestly “accidentally hit
upon a method of restoring air that had been injured by the
burning of candles.” He “put a [living] sprig of mint into
air in which a wax candle had burnt out and found that, on
the 27th of the same month, another candle could be
burned in this same air.” Somehow, the vegetation seemed
to have restored the air! Priestly found that while a mouse
could not breathe candle-exhausted air, air “restored” by
vegetation was not “at all inconvenient to a mouse.” The
key clue was that living vegetation adds something to the air.
How does vegetation “restore” air? Twenty-five years
later, Dutch physician Jan Ingenhousz solved the puzzle.
Working over several years, Ingenhousz reproduced and
significantly extended Priestly’s results, demonstrating that
air was restored only in the presence of sunlight, and only
by a plant’s green leaves, not by its roots. He proposed that
the green parts of the plant carry out a process (which we
now call photosynthesis) that uses sunlight to split carbon
dioxide (CO
2
) into carbon and oxygen. He suggested that
the oxygen was released as O
2
gas into the air, while the
carbon atom combined with water to form carbohydrates.
His proposal was a good guess, even though the later step
was subsequently modified. Chemists later found that the
proportions of carbon, oxygen, and hydrogen atoms in car-
bohydrates are indeed about one atom of carbon per mole-
cule of water (as the term carbohydrate indicates). A Swiss
botanist found in 1804 that water was a necessary reactant.
By the end of that century the overall reaction for photo-
synthesis could be written as:
CO
2
+ H
2
O + light energy —→ (CH
2
O) + O
2
It turns out, however, that there’s more to it than that.
When researchers began to examine the process in more
detail in the last century, the role of light proved to be un-
expectedly complex.
Van Helmont showed that soil did not add mass to a
growing plant. Priestly and Ingenhousz and others then
worked out the basic chemical reaction.
Discovery of the Light-Independent
Reactions
Ingenhousz’s early equation for photosynthesis includes
one factor we have not discussed: light energy. What role
does light play in photosynthesis? At the beginning of the
previous century, the English plant physiologist F. F.
Blackman began to address the question of the role of light
in photosynthesis. In 1905, he came to the startling conclu-
sion that photosynthesis is in fact a two-stage process, only
one of which uses light directly.
Blackman measured the effects of different light inten-
sities, CO
2
concentrations, and temperatures on photo-
synthesis. As long as light intensity was relatively low, he
found photosynthesis could be accelerated by increasing
the amount of light, but not by increasing the tempera-
ture or CO
2
concentration (figure 10.3). At high light in-
tensities, however, an increase in temperature or CO
2
concentration greatly accelerated photosynthesis. Black-
man concluded that photosynthesis consists of an initial
set of what he called “light” reactions, that are largely in-
dependent of temperature, and a second set of “dark” re-
actions, that seemed to be independent of light but lim-
ited by CO
2
. Do not be confused by Blackman’s
labels—the so-called “dark” reactions occur in the light
(in fact, they require the products of the light reactions);
their name simply indicates that light is not directly in-
volved in those reactions.
Blackman found that increased temperature increases
the rate of the dark carbon-reducing reactions, but only up
to about 35°C. Higher temperatures caused the rate to fall
off rapidly. Because 35°C is the temperature at which many
plant enzymes begin to be denatured (the hydrogen bonds
that hold an enzyme in its particular catalytic shape begin
to be disrupted), Blackman concluded that enzymes must
carry out the dark reactions.
Blackman showed that capturing photosynthetic energy
requires sunlight, while building organic molecules
does not.
186 Part III Energetics
10.2 Learning about photosynthesis: An experimental journey.
The Role of Light
The role of light in the so-called light and
dark reactions was worked out in the 1930s
by C. B. van Niel, then a graduate student at
Stanford University studying photosynthesis
in bacteria. One of the types of bacteria he
was studying, the purple sulfur bacteria, does
not release oxygen during photosynthesis;
instead, they convert hydrogen sulfide (H
2
S)
into globules of pure elemental sulfur that
accumulate inside themselves. The process
that van Niel observed was
CO
2
+ 2 H
2
S + light energy → (CH
2
O) + H
2
O + 2 S
The striking parallel between this equation
and Ingenhousz’s equation led van Niel to
propose that the generalized process of
photosynthesis is in fact
CO
2
+ 2 H
2
A + light energy → (CH
2
O) + H
2
O + 2 A
In this equation, the substance H
2
A serves
as an electron donor. In photosynthesis
performed by green plants, H
2
A is water,
while among purple sulfur bacteria, H
2
A is
hydrogen sulfide. The product, A, comes
from the splitting of H
2
A. Therefore, the
O
2
produced during green plant photosyn-
thesis results from splitting water, not car-
bon dioxide.
When isotopes came into common use in biology in the
early 1950s, it became possible to test van Niel’s revolu-
tionary proposal. Investigators examined photosynthesis in
green plants supplied with
18
O water; they found that the
18
O label ended up in oxygen gas rather than in carbohy-
drate, just as van Niel had predicted:
CO
2
+ 2 H
2
18
O + light energy —→ (CH
2
O) + H
2
O +
18
O
2
In algae and green plants, the carbohydrate typically pro-
duced by photosynthesis is the sugar glucose, which has six
carbons. The complete balanced equation for photosynthe-
sis in these organisms thus becomes
6 CO
2
+ 12 H
2
O + light energy —→ C
6
H
12
O
6
+ 6 O
2
+ 6 H
2
O.
We now know that the first stage of photosynthesis, the
light reactions, uses the energy of light to reduce NADP
(an electron carrier molecule) to NADPH and to manufac-
ture ATP. The NADPH and ATP from the first stage of
photosynthesis are then used in the second stage, the
Calvin cycle, to reduce the carbon in carbon dioxide and
form a simple sugar whose carbon skeleton can be used to
synthesize other organic molecules.
Van Niel discovered that photosynthesis splits water
molecules, incorporating the carbon atoms of carbon
dioxide gas and the hydrogen atoms of water into
organic molecules and leaving oxygen gas.
The Role of Reducing Power
In his pioneering work on the light reactions, van Niel had
further proposed that the reducing power (H
+
) generated by
the splitting of water was used to convert CO
2
into organic
matter in a process he called carbon fixation. Was he right?
In the 1950s Robin Hill demonstrated that van Niel was
indeed right, and that light energy could be used to generate
reducing power. Chloroplasts isolated from leaf cells were
able to reduce a dye and release oxygen in response to light.
Later experiments showed that the electrons released from
water were transferred to NADP
+
. Arnon and coworkers
showed that illuminated chloroplasts deprived of CO
2
accu-
mulate ATP. If CO
2
is then introduced, neither ATP nor
NADPH accumulate, and the CO
2
is assimilated into organic
molecules. These experiments are important for three rea-
sons. First, they firmly demonstrate that photosynthesis oc-
curs only within chloroplasts. Second, they show that the
light-dependent reactions use light energy to reduce NADP
+
and to manufacture ATP. Thirdly, they confirm that the
ATP and NADPH from this early stage of photosynthesis are
then used in the later light-independent reactions to reduce
carbon dioxide, forming simple sugars.
Hill showed that plants can use light energy to generate
reducing power. The incorporation of carbon dioxide
into organic molecules in the light-independent
reactions is called carbon fixation.
Chapter 10 Photosynthesis 187
Maximum rate
Excess CO
2
; 35°C
Temperature limited
Excess CO
2
; 20°C
CO
2
limited
Light intensity (foot-candles)
Rate of photosynthesis
Light limited
500
(b)
1000 1500 2000 2500
Insufficient CO
2
(0.01%); 20°C
FIGURE 10.3
Discovery of the dark reactions. (a) Blackman measured photosynthesis rates under
differing light intensities, CO
2
concentrations, and temperatures. (b) As this graph
shows, light is the limiting factor at low light intensities, while temperature and CO
2
concentration are the limiting factors at higher light intensities.
(a)
The Biophysics of Light
Where is the energy in light? What is
there in sunlight that a plant can use to
reduce carbon dioxide? This is the
mystery of photosynthesis, the one fac-
tor fundamentally different from
processes such as respiration. To an-
swer these questions, we will need to
consider the physical nature of light it-
self. James Clerk Maxwell had theo-
rized that light was an electromagnetic
wave—that is, that light moved
through the air as oscillating electric
and magnetic fields. Proof of this came
in a curious experiment carried out in a
laboratory in Germany in 1887. A
young physicist, Heinrich Hertz, was
attempting to verify a highly mathe-
matical theory that predicted the exis-
tence of electromagnetic waves. To see
whether such waves existed, Hertz de-
signed a clever experiment. On one
side of a room he constructed a powerful spark generator
that consisted of two large, shiny metal spheres standing
near each other on tall, slender rods. When a very high sta-
tic electrical charge was built up on one sphere, sparks
would jump across to the other sphere.
After constructing this device, Hertz set out to investigate
whether the sparking would create invisible electromagnetic
waves, so-called radio waves, as predicted by the mathemati-
cal theory. On the other side of the room, he placed the
world’s first radio receiver, a thin metal hoop on an insulat-
ing stand. There was a small gap at the bottom of the hoop,
so that the hoop did not quite form a complete circle. When
Hertz turned on the spark generator across the room, he saw
tiny sparks passing across the gap in the hoop! This was the
first demonstration of radio waves. But Hertz noted another
curious phenomenon. When UV light was shining across
the gap on the hoop, the sparks were produced more readily.
This unexpected facilitation, called the photoelectric effect,
puzzled investigators for many years.
The photoelectric effect was finally explained using a
concept proposed by Max Planck in 1901. Planck devel-
oped an equation that predicted the blackbody radiation
curve based upon the assumption that light and other forms
of radiation behaved as units of energy called photons. In
1905 Albert Einstein explained the photoelectric effect uti-
lizing the photon concept. Ultraviolet light has photons of
sufficient energy that when they fell on the loop, electrons
were ejected from the metal surface. The photons had
transferred their energy to the electrons, literally blasting
them from the ends of the hoop and thus facilitating the
passage of the electric spark induced by the radio waves.
Visible wavelengths of light were unable to remove the
electrons because their photons did not have enough en-
ergy to free the electrons from the metal surface at the ends
of the hoop.
The Energy in Photons
Photons do not all possess the same amount of energy (fig-
ure 10.4). Instead, the energy content of a photon is in-
versely proportional to the wavelength of the light: short-
wavelength light contains photons of higher energy than
long-wavelength light. X rays, which contain a great deal of
energy, have very short wavelengths—much shorter than visi-
ble light, making them ideal for high-resolution microscopes.
Hertz had noted that the strength of the photoelectric
effect depends on the wavelength of light; short wave-
lengths are much more effective than long ones in produc-
ing the photoelectric effect. Einstein’s theory of the photo-
electric effect provides an explanation: sunlight contains
photons of many different energy levels, only some of
which our eyes perceive as visible light. The highest energy
photons, at the short-wavelength end of the electromag-
netic spectrum (see figure 10.4), are gamma rays, with
wavelengths of less than 1 nanometer; the lowest energy
photons, with wavelengths of up to thousands of meters,
are radio waves. Within the visible portion of the spectrum,
violet light has the shortest wavelength and the most ener-
getic photons, and red light has the longest wavelength and
the least energetic photons.
188 Part III Energetics
10.3 Pigments capture energy from sunlight.
1 nm
400 nm
0.001 nm 10 nm 1000 nm
Increasing wavelength
Visible light
Increasing energy
0.01 cm 1 cm 1 m
Radio wavesInfraredX raysGamma rays
UV
light
100 m
430 nm 500 nm 560 nm 600 nm 650 nm 740 nm
FIGURE 10.4
The electromagnetic spectrum. Light is a form of electromagnetic energy conveniently
thought of as a wave. The shorter the wavelength of light, the greater its energy. Visible
light represents only a small part of the electromagnetic spectrum between 400 and 740
nanometers.
Ultraviolet Light
The sunlight that reaches the earth’s surface
contains a significant amount of ultraviolet
(UV) light, which, because of its shorter
wavelength, possesses considerably more en-
ergy than visible light. UV light is thought to
have been an important source of energy on
the primitive earth when life originated. To-
day’s atmosphere contains ozone (derived
from oxygen gas), which absorbs most of the
UV photons in sunlight, but a considerable
amount of UV light still manages to pene-
trate the atmosphere. This UV light is a po-
tent force in disrupting the bonds of DNA,
causing mutations that can lead to skin can-
cer. As we will describe in a later chapter,
loss of atmospheric ozone due to human ac-
tivities threatens to cause an enormous jump
in the incidence of human skin cancers
throughout the world.
Absorption Spectra and Pigments
How does a molecule “capture” the energy
of light? A photon can be envisioned as a
very fast-moving packet of energy. When it
strikes a molecule, its energy is either lost as
heat or absorbed by the electrons of the mol-
ecule, boosting those electrons into higher
energy levels. Whether or not the photon’s
energy is absorbed depends on how much
energy it carries (defined by its wavelength)
and on the chemical nature of the molecule it
hits. As we saw in chapter 2, electrons occupy
discrete energy levels in their orbits around
atomic nuclei. To boost an electron into a different energy
level requires just the right amount of energy, just as reach-
ing the next rung on a ladder requires you to raise your
foot just the right distance. A specific atom can, therefore,
absorb only certain photons of light—namely, those that
correspond to the atom’s available electron energy levels.
As a result, each molecule has a characteristic absorption
spectrum, the range and efficiency of photons it is capable
of absorbing.
Molecules that are good absorbers of light in the visible
range are called pigments. Organisms have evolved a vari-
ety of different pigments, but there are only two general
types used in green plant photosynthesis: carotenoids and
chlorophylls. Chlorophylls absorb photons within narrow
energy ranges. Two kinds of chlorophyll in plants, chloro-
phylls a and b, preferentially absorb violet-blue and red
light (figure 10.5). Neither of these pigments absorbs pho-
tons with wavelengths between about 500 and 600
nanometers, and light of these wavelengths is, therefore,
reflected by plants. When these photons are subsequently
absorbed by the pigment in our eyes, we perceive them as
green.
Chlorophyll a is the main photosynthetic pigment and is
the only pigment that can act directly to convert light en-
ergy to chemical energy. However, chlorophyll b, acting as
an accessory or secondary light-absorbing pigment, com-
plements and adds to the light absorption of chlorophyll a.
Chlorophyll b has an absorption spectrum shifted toward
the green wavelengths. Therefore, chlorophyll b can absorb
photons chlorophyll a cannot. Chlorophyll b therefore
greatly increases the proportion of the photons in sunlight
that plants can harvest. An important group of accessory
pigments, the carotenoids, assist in photosynthesis by cap-
turing energy from light of wavelengths that are not effi-
ciently absorbed by either chlorophyll.
In photosynthesis, photons of light are absorbed by
pigments; the wavelength of light absorbed depends
upon the specific pigment.
Chapter 10 Photosynthesis 189
Chlorophyll b Chlorophyll a
Relative light absorption
Wavelength (nm)
400 450 500 550 600 650 700
Carotenoids
FIGURE 10.5
The absorption spectrum of chlorophyll. The peaks represent wavelengths of
sunlight that the two common forms of photosynthetic pigment, chlorophyll a (solid
line) and chlorophyll b (dashed line), strongly absorb. These pigments absorb
predominately violet-blue and red light in two narrow bands of the spectrum and
reflect the green light in the middle of the spectrum. Carotenoids (not shown here)
absorb mostly blue and green light and reflect orange and yellow light.
Chlorophylls and Carotenoids
Chlorophylls absorb photons by means of an excitation
process analogous to the photoelectric effect. These pigments
contain a complex ring structure, called a porphyrin ring,
with alternating single and double bonds. At the center of the
ring is a magnesium atom. Photons absorbed by the pigment
molecule excite electrons in the ring, which are then chan-
neled away through the alternating carbon-bond system. Sev-
eral small side groups attached to the outside of the ring alter
the absorption properties of the molecule in different kinds of
chlorophyll (figure 10.6). The precise absorption spectrum is
also influenced by the local microenvironment created by the
association of chlorophyll with specific proteins.
Once Ingenhousz demonstrated that only the green parts
of plants can “restore” air, researchers suspected chlorophyll
was the primary pigment that plants employ to absorb light
in photosynthesis. Experiments conducted in the 1800s
clearly verified this suspicion. One such experiment, per-
formed by T. W. Englemann in 1882 (figure 10.7), serves as
a particularly elegant example, simple in design and clear in
outcome. Englemann set out to characterize the action
spectrum of photosynthesis, that is, the relative effective-
ness of different wavelengths of light in promoting photo-
synthesis. He carried out the entire experiment utilizing a
single slide mounted on a microscope. To obtain different
wavelengths of light, he placed a prism under his micro-
scope, splitting the light that illuminated the slide into a
spectrum of colors. He then arranged a filament of green
algal cells across the spectrum, so that different parts of the
filament were illuminated with different wavelengths, and
allowed the algae to carry out photosynthesis. To assess
how fast photosynthesis was proceeding, Englemann chose
to monitor the rate of oxygen production. Lacking a mass
spectrometer and other modern instruments, he added
aerotactic (oxygen-seeking) bacteria to the slide; he knew
they would gather along the filament at locations where
oxygen was being produced. He found that the bacteria ac-
cumulated in areas illuminated by red and violet light, the
two colors most strongly absorbed by chlorophyll.
All plants, algae, and cyanobacteria use chlorophyll a as
their primary pigments. It is reasonable to ask why these
photosynthetic organisms do not use a pigment like retinal
(the pigment in our eyes), which has a broad absorption
spectrum that covers the range of 500 to 600 nanometers.
The most likely hypothesis involves photoefficiency. Al-
though retinal absorbs a broad range of wavelengths, it
does so with relatively low efficiency. Chlorophyll, in con-
trast, absorbs in only two narrow bands, but does so with
high efficiency. Therefore, plants and most other photo-
synthetic organisms achieve far higher overall photon cap-
ture rates with chlorophyll than with other pigments.
190 Part III Energetics
Thylakoid
membrane
Thylakoid
Granum
Chlorophyll a: R = —CH
3
Chlorophyll b: R = —CHO
CO
2
CH
3
CH
2
CH
2
C
O
CH
2
CH
CCH
3
CH
2
CH
2
CH
2
CHCH
3
CH
2
CH
2
CH
2
CHCH
3
CH
2
CH
2
CH
2
CHCH
3
CH
3
CH
2
CH
3
CH
3
H
Mg
O
O
NN
NN
CH
3
CH
3
H
H
H
H
Porphyrin head
HRCHH
2
C
Hydrocarbon
tail
Chlorophyll molecules
embedded in a
protein complex
in the thylakoid
membrane
FIGURE 10.6
Chlorophyll.
Chlorophyll
molecules consist
of a porphyrin
head and a
hydrocarbon tail
that anchors the
pigment molecule
to hydrophobic
regions of proteins
embedded within
the membranes of
thylakoids. The
only difference
between the two
chlorophyll
molecules is the
substitution of a
—CHO
(aldehyde) group
in chlorophyll b
for a —CH
3
(methyl) group in
chlorophyll a.
Carotenoids consist of carbon rings linked to chains
with alternating single and double bonds. They can absorb
photons with a wide range of energies, although they are
not always highly efficient in transferring this energy.
Carotenoids assist in photosynthesis by capturing energy
from light of wavelengths that are not efficiently absorbed
by chlorophylls (figure 10.8; see figure 10.5).
A typical carotenoid is β-carotene, whose two carbon
rings are connected by a chain of 18 carbon atoms with al-
ternating single and double bonds. Splitting a molecule of
β-carotene into equal halves produces two molecules of vit-
amin A. Oxidation of vitamin A produces retinal, the pig-
ment used in vertebrate vision. This explains why carrots,
which are rich in β-carotene, enhance vision.
A pigment is a molecule that absorbs light. The
wavelengths absorbed by a particular pigment depend
on the available energy levels to which light-excited
electrons can be boosted in the pigment.
Chapter 10 Photosynthesis 191
Absorbance
Filament of
green alga
Oxygen-seeking bacteria
T.W. Englemann revealed the action spectrum of photosynthesis in the filamentous alga Spirogyra in 1882. Englemann used the rate of
oxygen production to measure the rate of photosynthesis. As his oxygen indicator, he chose bacteria that are attracted by oxygen. In place
of the mirror and diaphragm usually used to illuminate objects under view in his microscope, he substituted a "microspectral apparatus,"
which, as its name implies, produced a tiny spectrum of colors that it projected upon the slide under the microscope. Then he arranged a
filament of algal cells parallel to the spread of the spectrum. The oxygen-seeking bacteria congregated mostly in the areas where the violet
and red wavelengths fell upon the algal filament.
FIGURE 10.7
Constructing an action spectrum for photosynthesis. As you can see, the action spectrum for photosynthesis that Englemann revealed
in his experiment parallels the absorption spectrum of chlorophyll (see figure 10.5).
Oak leaf in summer
Oak leaf in autumn
FIGURE 10.8
Fall colors are produced by carotenoids and other accessory pigments. During the spring and summer, chlorophyll in leaves masks the
presence of carotenoids and other accessory pigments. When cool fall temperatures cause leaves to cease manufacturing chlorophyll, the
chlorophyll is no longer present to reflect green light, and the leaves reflect the orange and yellow light that carotenoids and other
pigments do not absorb.
Organizing Pigments into
Photosystems
The light reactions of photosynthesis occur in membranes.
In bacteria like those studied by van Niel, the plasma mem-
brane itself is the photosynthetic membrane. In plants and
algae, by contrast, photosynthesis is carried out by or-
ganelles that are the evolutionary descendants of photosyn-
thetic bacteria, chloroplasts—the photosynthetic mem-
branes exist within the chloroplasts. The light reactions
take place in four stages:
1. Primary photoevent. A photon of light is captured
by a pigment. The result of this primary photoevent
is the excitation of an electron within the pigment.
2. Charge separation. This excitation energy is trans-
ferred to a specialized chlorophyll pigment termed a
reaction center, which reacts by transferring an ener-
getic electron to an acceptor molecule, thus initiating
electron transport.
3. Electron transport. The excited electron is shut-
tled along a series of electron-carrier molecules em-
bedded within the photosynthetic membrane. Several
of them react by transporting protons across the
membrane, generating a gradient of proton concen-
tration. Its arrival at the pump induces the transport
of a proton across the membrane. The electron is
then passed to an acceptor.
4. Chemiosmosis. The protons that accumulate on
one side of the membrane now flow back across the
membrane through specific protein complexes where
chemiosmotic synthesis of ATP takes place, just as it
does in aerobic respiration.
Discovery of Photosystems
One way to study how pigments absorb light is to measure
the dependence of the output of photosynthesis on the in-
tensity of illumination—that is, how much photosynthesis
is produced by how much light. When experiments of this
sort are done on plants, they show that the output of pho-
tosynthesis increases linearly at low intensities but lessens
at higher intensities, finally saturating at high-intensity
light (figure 10.9). Saturation occurs because all of the
light-absorbing capacity of the plant is in use; additional
light doesn’t increase the output because there is nothing
to absorb the added photons.
It is tempting to think that at saturation, all of a plant’s
pigment molecules are in use. In 1932 plant physiologists
Robert Emerson and William Arnold set out to test this
hypothesis in an organism where they could measure both
the number of chlorophyll molecules and the output of
photosynthesis. In their experiment, they measured the
oxygen yield of photosynthesis when Chlorella (unicellular
green algae) were exposed to very brief light flashes lasting
only a few microseconds. Assuming the hypothesis of pig-
ment saturation to be correct, they expected to find that as
they increased the intensity of the flashes, the yield per
flash would increase, until each chlorophyll molecule ab-
sorbed a photon, which would then be used in the light re-
actions, producing a molecule of O
2
.
Unexpectedly, this is not what happened. Instead, satu-
ration was achieved much earlier, with only one molecule
of O
2
per 2500 chlorophyll molecules! This led Emerson
and Arnold to conclude that light is absorbed not by inde-
pendent pigment molecules, but rather by clusters of
chlorophyll and accessory pigment molecules which have
come to be called photosystems. Light is absorbed by any one
of the hundreds of pigment molecules in a photosystem,
which transfer their excitation energy to one with a lower
energy level than the others. This reaction center of the
photosystem acts as an energy sink, trapping the excitation
energy. It was the saturation of these reaction centers, not
individual molecules, that was observed by Emerson and
Arnold.
Architecture of a Photosystem
In chloroplasts and all but the most primitive bacteria, light
is captured by such photosystems. Each photosystem is a
network of chlorophyll a molecules, accessory pigments,
and associated proteins held within a protein matrix on the
surface of the photosynthetic membrane. Like a magnify-
ing glass focusing light on a precise point, a photosystem
channels the excitation energy gathered by any one of its
pigment molecules to a specific molecule, the reaction cen-
ter chlorophyll. This molecule then passes the energy out
192 Part III Energetics
Expected
Observed
Saturation when all
photosystems are
in use
Intensity of light flashes
Output ( O
2
yield per flash)
Saturation when all
chlorophyll molecules
are in use
FIGURE 10.9
Emerson and Arnold’s experiment. When photosynthetic
saturation is achieved, further increases in intensity cause no
increase in output.
of the photosystem so it can be put to work driving the syn-
thesis of ATP and organic molecules.
A photosystem thus consists of two closely linked
components: (1) an antenna complex of hundreds of pig-
ment molecules that gather photons and feed the cap-
tured light energy to the reaction center; and (2) a reac-
tion center, consisting of one or more chlorophyll a
molecules in a matrix of protein, that passes the energy
out of the photosystem.
The Antenna Complex. The antenna complex captures
photons from sunlight (figure 10.10). In chloroplasts, the
antenna complex is a web of chlorophyll molecules linked
together and held tightly on the thylakoid membrane by a
matrix of proteins. Varying amounts of carotenoid acces-
sory pigments may also be present. The protein matrix
serves as a sort of scaffold, holding individual pigment mol-
ecules in orientations that are optimal for energy transfer.
The excitation energy resulting from the absorption of a
photon passes from one pigment molecule to an adjacent
molecule on its way to the reaction center. After the trans-
fer, the excited electron in each molecule returns to the
low-energy level it had before the photon was absorbed.
Consequently, it is energy, not the excited electrons them-
selves, that passes from one pigment molecule to the next.
The antenna complex funnels the energy from many elec-
trons to the reaction center.
The Reaction Center. The reaction center is a trans-
membrane protein-pigment complex. In the reaction cen-
ter of purple photosynthetic bacteria, which is simpler than
in chloroplasts but better understood, a pair of chlorophyll
a molecules acts as a trap for photon energy, passing an ex-
cited electron to an acceptor precisely positioned as its
neighbor. Note that here the excited electron itself is trans-
ferred, not just the energy as we saw in pigment-pigment
transfers. This allows the photon excitation to move away
from the chlorophylls and is the key conversion of light to
chemical energy.
Figure 10.11 shows the transfer of energy from the reac-
tion center to the primary electron acceptor. By energizing
an electron of the reaction center chlorophyll, light creates
a strong electron donor where none existed before. The
chlorophyll transfers the energized electron to the primary
acceptor, a molecule of quinone, reducing the quinone and
converting it to a strong electron donor. A weak electron
donor then donates a low-energy electron to the chloro-
phyll, restoring it to its original condition. In plant chloro-
plasts, water serves as the electron donor.
Photosystems contain pigments that capture photon
energy from light. The pigments transfer the energy to
reaction centers. There, the energy excites electrons,
which are channeled away to do chemical work.
Chapter 10 Photosynthesis 193
Electron
donor
Electron
acceptor
Photon
Reaction
center
chlorophyll
Chlorophyll
molecules
Photosystem
FIGURE 10.10
How the antenna complex works. When light of the proper
wavelength strikes any pigment molecule within a photosystem,
the light is absorbed by that pigment molecule. The excitation
energy is then transferred from one molecule to another within
the cluster of pigment molecules until it encounters the reaction
center chlorophyll a. When excitation energy reaches the reaction
center chlorophyll, electron transfer is initiated.
Electron
donor
Electron
acceptor
Acceptor
reduced
Chlorophyll
oxidized
Acceptor
reduced
Donor
oxidized
Excited
chlorophyll
molecule
+ – + –
FIGURE 10.11
Converting light to chemical energy. The reaction center
chlorophyll donates a light-energized electron to the primary
electron acceptor, reducing it. The oxidized chlorophyll then fills
its electron “hole” by oxidizing a donor molecule.
How Photosystems Convert Light
to Chemical Energy
Bacteria Use a Single Photosystem
Photosynthetic pigment arrays are thought to have evolved
more than 3 billion years ago in bacteria similar to the sul-
fur bacteria studied by van Niel.
1. Electron is joined with a proton to make hydrogen.
In these bacteria, the absorption of a photon of light at a
peak absorption of 870 nanometers (near infrared, not visi-
ble to the human eye) by the photosystem results in the
transmission of an energetic electron along an electron
transport chain, eventually combining with a proton to
form a hydrogen atom. In the sulfur bacteria, the proton is
extracted from hydrogen sulfide, leaving elemental sulfur as
a by-product. In bacteria that evolved later, as well as in
plants and algae, the proton comes from water, producing
oxygen as a by-product.
2. Electron is recycled to chlorophyll. The ejection of
an electron from the bacterial reaction center leaves it
short one electron. Before the photosystem of the sulfur
bacteria can function again, an electron must be re-
turned. These bacteria channel the electron back to the
pigment through an electron transport system similar to
the one described in chapter 9; the electron’s passage drives
a proton pump that promotes the chemiosmotic synthesis
of ATP. One molecule of ATP is produced for every
three electrons that follow this path. Viewed overall (fig-
ure 10.12), the path of the electron is thus a circle.
Chemists therefore call the electron transfer process
leading to ATP formation cyclic photophosphorylation.
Note, however, that the electron that left the P
870
reac-
tion center was a high-energy electron, boosted by the
absorption of a photon of light, while the electron that
returns has only as much energy as it had before the pho-
ton was absorbed. The difference in the energy of that
electron is the photosynthetic payoff, the energy that drives
the proton pump.
For more than a billion years, cyclic photophosphory-
lation was the only form of photosynthetic light reaction
that organisms used. However, its major limitation is
that it is geared only toward energy production, not to-
ward biosynthesis. Most photosynthetic organisms incor-
porate atmospheric carbon dioxide into carbohydrates.
Because the carbohydrate molecules are more reduced
(have more hydrogen atoms) than carbon dioxide, a
source of reducing power (that is, hydrogens) must be
provided. Cyclic photophosphorylation does not do this.
The hydrogen atoms extracted from H
2
S are used as a
source of protons, and are not available to join to carbon.
Thus bacteria that are restricted to this process must
scavenge hydrogens from other sources, an inefficient
undertaking.
Why Plants Use Two Photosystems
After the sulfur bacteria appeared, other kinds of bacteria
evolved an improved version of the photosystem that over-
came the limitation of cyclic photophosphorylation in a
neat and simple way: a second, more powerful photosystem
using another arrangement of chlorophyll a was combined
with the original.
In this second photosystem, called photosystem II,
molecules of chlorophyll a are arranged with a different
geometry, so that more shorter wavelength, higher energy
photons are absorbed than in the ancestral photosystem,
which is called photosystem I. As in the ancestral photo-
system, energy is transmitted from one pigment molecule
to another within the antenna complex of these photosys-
tems until it reaches the reaction center, a particular pig-
ment molecule positioned near a strong membrane-bound
electron acceptor. In photosystem II, the absorption peak
(that is, the wavelength of light most strongly absorbed) of
the pigments is approximately 680 nanometers; therefore,
the reaction center pigment is called P
680
. The absorption
peak of photosystem I pigments in plants is 700 nanome-
ters, so its reaction center pigment is called P
700
. Working
together, the two photosystems carry out a noncyclic elec-
tron transfer.
When the rate of photosynthesis is measured using two
light beams of different wavelengths (one red and the
194 Part III Energetics
Electron
acceptor
Photon
Energy of electrons
Plastocyanin
Photosystem
ATP
ADP
pC
Ferredoxin
b
6
-f
complex
b
6
-f complex
Fd
e
–
e
–
e
–
Excited
reaction
center
Reaction
center
P
870
FIGURE 10.12
The path of an electron in purple sulfur bacteria. When a
light-energized electron is ejected from the photosystem reaction
center (P
870
), it passes in a circle, eventually returning to the
photosystem from which it was ejected.
other far-red), the rate was greater than the sum of the
rates using individual beams of red and far-red light (fig-
ure 10.13). This surprising result, called the enhancement
effect, can be explained by a mechanism involving two
photosystems acting in series (that is, one after the other),
one of which absorbs preferentially in the red, the other
in the far-red.
The use of two photosystems solves the problem of ob-
taining reducing power in a simple and direct way, by har-
nessing the energy of two photosystems. The scheme
shown in figure 10.14, called a Z diagram, illustrates the
two electron-energizing steps, one catalyzed by each pho-
tosystem. The electrons originate from water, which holds
onto its electrons very tightly (redox potential = +820 mV),
and end up in NADPH, which holds its electrons much
more loosely (redox potential = –320 mV).
In sulfur bacteria, excited electrons ejected from the
reaction center travel a circular path, driving a proton
pump and then returning to their original photosystem.
Plants employ two photosystems in series, which
generates power to reduce NADP
+
to NADPH with
enough left over to make ATP.
Chapter 10 Photosynthesis 195
Far-red
light on
Both
lights on
Red light
on
Off Off
Time
Relative rate of photosynthesis
Off
FIGURE 10.13
The “enhancement effect.” The rate of photosynthesis when
red and far-red light are provided together is greater than the sum
of the rates when each wavelength is provided individually. This
result baffled researchers in the 1950s. Today it provides the key
evidence that photosynthesis is carried out by two photochemical
systems with slightly different wavelength optima.
Proton gradient
formed for ATP
synthesis
Water-splitting
enzyme
Photon
Photon
Energy of electrons
Plastocyanin
Ferredoxin
Plastoquinone
Photosystem II
NADP
+
+ H
+
b
6
-f
complex
NADP
reductase
b
6
-f complex Photosystem I
NADP reductase
e
–
e
–
H
2
O
2H
+
+ H11002O
2
e
–
H
+
e
–
e
–
pC
Q
P
700
P
680
NADPH
Fd
1
2
Reaction
center
Excited
reaction
center
Reaction
center
Excited
reaction
center
FIGURE 10.14
A Z diagram of photosystems I and II. Two photosystems work sequentially. First, a photon of light ejects a high-energy electron from
photosystem II; that electron is used to pump a proton across the membrane, contributing chemiosmotically to the production of a
molecule of ATP. The ejected electron then passes along a chain of cytochromes to photosystem I. When photosystem I absorbs a photon
of light, it ejects a high-energy electron used to drive the formation of NADPH.
How the Two
Photosystems of Plants
Work Together
Plants use the two photosystems dis-
cussed earlier in series, first one and
then the other, to produce both ATP
and NADPH. This two-stage process
is called noncyclic photophosphory-
lation, because the path of the elec-
trons is not a circle—the electrons
ejected from the photosystems do not
return to it, but rather end up in
NADPH. The photosystems are re-
plenished instead with electrons ob-
tained by splitting water. Photosystem
II acts first. High-energy electrons
generated by photosystem II are used
to synthesize ATP and then passed to
photosystem I to drive the production
of NADPH. For every pair of elec-
trons obtained from water, one mole-
cule of NADPH and slightly more
than one molecule of ATP are pro-
duced.
Photosystem II
The reaction center of photosystem II,
called P
680
, closely resembles the reac-
tion center of purple bacteria. It con-
sists of more than 10 transmembrane
protein subunits. The light-harvesting
antenna complex consists of some 250
molecules of chlorophyll a and acces-
sory pigments bound to several protein
chains. In photosystem II, the oxygen
atoms of two water molecules bind to a
cluster of manganese atoms which are
embedded within an enzyme and
bound to the reaction center. In a way that is poorly under-
stood, this enzyme splits water, removing electrons one at a
time to fill the holes left in the reaction center by departure
of light-energized electrons. As soon as four electrons have
been removed from the two water molecules, O
2
is released.
The Path to Photosystem I
The primary electron acceptor for the light-energized elec-
trons leaving photosystem II is a quinone molecule, as it
was in the bacterial photosystem described earlier. The re-
duced quinone which results (plastoquinone, symbolized Q)
is a strong electron donor; it passes the excited electron to a
proton pump called the b
6
-f complex embedded within the
thylakoid membrane (figure 10.15). This complex closely
resembles the bc
1
complex in the respiratory electron trans-
port chain of mitochondria discussed in chapter 9. Arrival
of the energetic electron causes the b
6
-f complex to pump a
proton into the thylakoid space. A small copper-containing
protein called plastocyanin (symbolized pC) then carries the
electron to photosystem I.
Making ATP: Chemiosmosis
Each thylakoid is a closed compartment into which pro-
tons are pumped from the stroma by the b
6
-f complex.
The splitting of water also produces added protons that
contribute to the gradient. The thylakoid membrane is
impermeable to protons, so protons cross back out almost
exclusively via the channels provided by ATP synthases.
These channels protrude like knobs on the external sur-
face of the thylakoid membrane. As protons pass out of
196 Part III Energetics
NADPH
Photosystem II Photosystem Ib
6
-f complex
Photon
Stroma
Thylakoid
space
H
2
O
2H
+
H
+
H
+
+
NADP
+
NADP
reductase
Thylakoid
membrane
Antenna
complex
Plastoquinone
Water-splitting
enzyme
Photon
Proton
gradient
Plastocyanin Ferredoxin
Fd
pC
Q
H11002O
2
1
2
FIGURE 10.15
The photosynthetic electron transport system. When a photon of light strikes a pigment
molecule in photosystem II, it excites an electron. This electron is coupled to a proton
stripped from water by an enzyme and is passed along a chain of membrane-bound
cytochrome electron carriers (red arrow). When water is split, oxygen is released from the
cell, and the hydrogen ions remain in the thylakoid space. At the proton pump
(b
6
-f complex), the energy supplied by the photon is used to transport a proton across the
membrane into the thylakoid. The concentration of hydrogen ions within the thylakoid
thus increases further. When photosystem I absorbs another photon of light, its pigment
passes a second high-energy electron to a reduction complex, which generates NADPH.
the thylakoid through the ATP syn-
thase channel, ADP is phosphorylated
to ATP and released into the stroma,
the fluid matrix inside the chloroplast
(figure 10.16). The stroma contains
the enzymes that catalyze the reac-
tions of carbon fixation.
Photosystem I
The reaction center of photosystem I,
called P
700
, is a transmembrane complex
consisting of at least 13 protein sub-
units. Energy is fed to it by an antenna
complex consisting of 130 chlorophyll a
and accessory pigment molecules. Pho-
tosystem I accepts an electron from
plastocyanin into the hole created by
the exit of a light-energized electron.
This arriving electron has by no means
lost all of its light-excited energy; al-
most half remains. Thus, the absorp-
tion of a photon of light energy by
photosystem I boosts the electron leav-
ing the reaction center to a very high
energy level. Unlike photosystem II
and the bacterial photosystem, photo-
system I does not rely on quinones as
electron acceptors. Instead, it passes
electrons to an iron-sulfur protein
called ferredoxin (Fd).
Making NADPH
Photosystem I passes electrons to
ferredoxin on the stromal side of the
membrane (outside the thylakoid). The
reduced ferredoxin carries a very-high-
potential electron. Two of them, from
two molecules of reduced ferredoxin, are then donated to a
molecule of NADP
+
to form NADPH. The reaction is cat-
alyzed by the membrane-bound enzyme NADP reductase.
Because the reaction occurs on the stromal side of the
membrane and involves the uptake of a proton in forming
NADPH, it contributes further to the proton gradient es-
tablished during photosynthetic electron transport.
Making More ATP
The passage of an electron from water to NADPH in the
noncyclic photophosphorylation described previously gen-
erates one molecule of NADPH and slightly more than one
molecule of ATP. However, as you will learn later in this
chapter, building organic molecules takes more energy than
that—it takes one-and-a-half ATP molecules per NADPH
molecule to fix carbon. To produce the extra ATP, many
plant species are capable of short-circuiting photosystem I,
switching photosynthesis into a cyclic photophosphorylation
mode, so that the light-excited electron leaving photosystem
I is used to make ATP instead of NADPH. The energetic
electron is simply passed back to the b
6
-f complex rather
than passing on to NADP
+
. The b
6
-f complex pumps out a
proton, adding to the proton gradient driving the chemios-
motic synthesis of ATP. The relative proportions of cyclic
and noncyclic photophosphorylation in these plants deter-
mines the relative amounts of ATP and NADPH available
for building organic molecules.
The electrons that photosynthesis strips from water
molecules provide the energy to form ATP and
NADPH. The residual oxygen atoms of the water
molecules combine to form oxygen gas.
Chapter 10 Photosynthesis 197
Photosystem II b
6
-f complex ATP synthase
H
+
H
+
H
+
H
+
H
+
H
+
H
+
H
+
Photon
Stroma
Thylakoid
space
ATPADP
Q
Chloroplast
Plant cell
H11002O
2
1
2
H
2
O
2
FIGURE 10.16
Chemiosmosis in a chloroplast. The b
6
-f complex embedded in the thylakoid membrane
pumps protons into the interior of the thylakoid. ATP is produced on the outside surface of
the membrane (stroma side), as protons diffuse back out of the thylakoid through ATP
synthase channels.
The Calvin Cycle
Photosynthesis is a way of making organic molecules from
carbon dioxide (CO
2
). These organic molecules contain
many C—H bonds and are highly reduced compared with
CO
2
. To build organic molecules, cells use raw materials
provided by the light reactions:
1. Energy. ATP (provided by cyclic and noncyclic pho-
tophosphorylation) drives the endergonic reactions.
2. Reducing power. NADPH (provided by photosys-
tem I) provides a source of hydrogens and the energetic
electrons needed to bind them to carbon atoms. Much
of the light energy captured in photosynthesis ends up
invested in the energy-rich C—H bonds of sugars.
Carbon Fixation
The key step in the Calvin cycle—the event that makes the
reduction of CO
2
possible—is the attachment of CO
2
to a
very special organic molecule. Photosynthetic cells produce
this molecule by reassembling the bonds of two intermedi-
ates in glycolysis, fructose 6-phosphate and glyceraldehyde
3-phosphate, to form the energy-rich five-carbon sugar,
ribulose1,5-bisphosphate (RuBP), and a four-carbon sugar.
CO
2
binds to RuBP in the key process called carbon
fixation, forming two three-carbon molecules of phospho-
glycerate (PGA) (figure 10.17). The enzyme that carries
out this reaction, ribulose bisphosphate carboxylase/oxygenase
(usually abbreviated rubisco) is a very large four-subunit
enzyme present in the chloroplast stroma. This enzyme
works very sluggishly, processing only about three mole-
cules of RuBP per second (a typical enzyme processes
about 1000 substrate molecules per second). Because it
works so slowly, many molecules of rubisco are needed.
In a typical leaf, over 50% of all the protein is rubisco. It
is thought to be the most abundant protein on earth.
Discovering the Calvin Cycle
Nearly 100 years ago, Blackman concluded that, because of
its temperature dependence, photosynthesis might involve
enzyme-catalyzed reactions. These reactions form a cycle
of enzyme-catalyzed steps similar to the Krebs cycle. This
cycle of reactions is called the Calvin cycle, after its dis-
coverer, Melvin Calvin of the University of California,
Berkeley. Because the cycle begins when CO
2
binds RuBP
to form PGA, and PGA contains three carbon atoms, this
process is also called C
3
photosynthesis.
198 Part III Energetics
10.4 Cells use the energy and reducing power captured by the light
reactions to make organic molecules.
H
2
C
Rubisco
2 molecules of
phosphoglycerate
(PGA)
CO
O
HC OH
HC OH
P
H
2
C O
Ribulose
1,5-bisphosphate
(RuBP)
P H
2
C
C
O
–
O
O
HC OH
P
CO
2
+ H
2
O
H
2
C
C
O
–
O
O
HC OH
P
FIGURE 10.17
The key step in the Calvin cycle.
Melvin Calvin and his coworkers at the
University of California worked out the
first step of what later became known as
the Calvin cycle. They exposed
photosynthesizing algae to radioactive
carbon dioxide (
14
CO
2
). By following the
fate of a radioactive carbon atom, they
found that it first binds to a molecule of
ribulose 1,5-bisphosphate (RuBP), then
immediately splits, forming two
molecules of phosphoglycerate (PGA).
One of these PGAs contains the
radioactive carbon atom. In 1948,
workers isolated the enzyme responsible
for this remarkable carbon-fixing
reaction: rubisco.
The Energy Cycle
The energy-capturing metabolisms of the chloroplasts
studied in this chapter and the mitochondria studied in the
previous chapter are intimately related. Photosynthesis uses
the products of respiration as starting substrates, and respi-
ration uses the products of photosynthesis as its starting
substrates (figure 10.18). The Calvin cycle even uses part of
the ancient glycolytic pathway, run in reverse, to produce
glucose. And, the principal proteins involved in electron
transport in plants are related to those in mitochondria,
and in many cases are actually the same.
Photosynthesis is but one aspect of plant biology, al-
though it is an important one. In chapters 37 through 43,
we will examine plants in more detail. We have treated
photosynthesis here, in a section devoted to cell biology,
because photosynthesis arose long before plants did, and all
organisms depend directly or indirectly on photosynthesis
for the energy that powers their lives.
Chloroplasts put ATP and NADPH to work building
carbon-based molecules, a process that essentially
reverses the breakdown of such molecules that occurs
in mitochondria. Taken together, chloroplasts and
mitochondria carry out a cycle in which energy enters
from the sun and leaves as heat and work.
Chapter 10 Photosynthesis 199
ATP
Photo-
system
I
NADP
+
ADP
ATP
Calvin
cycle
ATP
ATP
NADPH
Sunlight
Glucose
Chloroplast Mitochondrion
Heat
Pyruvate
Krebs
cycle
Electron
transport
system
NAD
+
NADH
Photo-
system
II
CO
2
H
2
O
O
2
FIGURE 10.18
Chloroplasts and mitochondria: Completing an energy cycle. Water and oxygen gas cycle between chloroplasts and mitochondria
within a plant cell, as do glucose and CO
2
. Cells with chloroplasts require an outside source of CO
2
and water and generate glucose and
oxygen. Cells without chloroplasts, such as animal cells, require an outside source of glucose and oxygen and generate CO
2
and water.
Reactions of the Calvin Cycle
In a series of reactions (figure 10.19), three molecules of
CO
2
are fixed by rubisco to produce six molecules of PGA
(containing 6 × 3 = 18 carbon atoms in all, three from CO
2
and 15 from RuBP). The 18 carbon atoms then undergo a
cycle of reactions that regenerates the three molecules of
RuBP used in the initial step (containing 3 × 5 = 15 carbon
atoms). This leaves one molecule of glyceraldehyde 3-
phosphate (three carbon atoms) as the net gain.
The net equation of the Calvin cycle is:
3 CO
2
+ 9 ATP + 6 NADPH + water —→
glyceraldehyde 3-phosphate + 8 P
i
+ 9 ADP + 6 NADP
+
With three full turns of the cycle, three molecules of
carbon dioxide enter, a molecule of glyceraldehyde 3-
phosphate (G3P) is produced, and three molecules of
RuBP are regenerated (figure 10.20).
We now know that light is required indirectly for differ-
ent segments of the CO
2
reduction reactions. Five of the
Calvin cycle enzymes—including rubisco—are light acti-
vated; that is, they become functional or operate more effi-
ciently in the presence of light. Light also promotes trans-
port of three-carbon intermediates across chloroplast
membranes that are required for Calvin cycle reactions. And
finally, light promotes the influx of Mg
++
into the chloro-
plast stroma, which further activates the enzyme rubisco.
200 Part III Energetics
THE CALVIN CYCLE
123
3
3
CO
2
6
3-phosphoglycerate
P
The Calvin cycle begins when a carbon
atom from a CO
2
molecule is added to a
five-carbon molecule (the starting
material). The resulting six-carbon
molecule is unstable and immediately
splits into three-carbon molecules.
Then, through a series of reactions,
energy from ATP and hydrogens from
NADPH (the products of the light
reactions) are added to the three-
carbon molecules. The now-reduced
three-carbon molecules either combine
to make glucose or are used to make
other molecules.
Most of the reduced three-carbon
molecules are used to regenerate the
five-carbon starting material, thus
completing the cycle.
5
Glyceraldehyde
3-phosphate
3
RuBP
3
P
6
6
6
Glyceraldehyde
3-phosphate
Glucose
NADPH
3-phosphoglycerate
ATP
ATP
P
6
Glyceraldehyde
3-phosphate
P
1
P
(Starting material)
RuBP
(Starting material)
FIGURE 10.19
How the Calvin cycle works.
Output of the Calvin Cycle
The glyceraldehyde 3-phosphate that is the product of the
Calvin cycle is a three-carbon sugar that is a key intermedi-
ate in glycolysis. Much of it is exported from the chloro-
plast to the cytoplasm of the cell, where the reversal of sev-
eral reactions in glycolysis allows it to be converted to
fructose 6-phosphate and glucose 1-phosphate, and from
that to sucrose, a major transport sugar in plants (sucrose,
common table sugar, is a disaccharide made of fructose and
glucose).
In times of intensive photosynthesis, glyceraldehyde 3-
phosphate levels in the stroma of the chloroplast rise. As a
consequence, some glyceraldehyde 3-phosphate in the
chloroplast is converted to glucose 1-phosphate, in an anal-
ogous set of reactions to those done in the cytoplasm, by
reversing several reactions similar to those of glycolysis.
The glucose 1-phosphate is then combined into an insolu-
ble polymer, forming long chains of starch stored as bulky
starch grains in chloroplasts.
Plants incorporate carbon dioxide into sugars by means of
a cycle of reactions called the Calvin cycle, which is driven
by the ATP and NADPH produced in the light reactions
which are consumed as CO
2
is reduced to G3P.
Chapter 10 Photosynthesis 201
3 molecules of
1 molecule of
Glyceraldehyde 3-phosphate (3C) (G3P)
Glucose and
other sugars
3 molecules of
6 molecules of
6 molecules of
6 molecules of
5 molecules of
Ribulose 1,5-bisphosphate (RuBP) (5C)
3-phosphoglycerate (3C) (PGA)
Glyceraldehyde 3-phosphate (3C)
Glyceraldehyde 3-phosphate (3C) (G3P)
PGA
kinase
Rubisco
Reforming
RuBP
Reverse of
glycolysis
G3P dehydrogenase
Carbon fixation
Stroma of chloroplast
Carbon
dioxide (CO
2
)
6 NADPH
3 ATP
3 ADP
6 ATP
6 ADP
6 NADP
+
P
i
6
P
i
2
1,3-bisphosphoglycerate (3C)
FIGURE 10.20
The Calvin cycle. For every three molecules of CO
2
that enter the cycle, one molecule of the three-carbon compound, glyceraldehyde 3-
phosphate (G3P), is produced. Notice that the process requires energy stored in ATP and NADPH, which are generated by the light
reactions. This process occurs in the stroma of the chloroplast.
Photorespiration
Evolution does not necessarily result in optimum solutions.
Rather, it favors workable solutions that can be derived
from others that already exist. Photosynthesis is no excep-
tion. Rubisco, the enzyme that catalyzes the key carbon-
fixing reaction of photosynthesis, provides a decidedly sub-
optimal solution. This enzyme has a second enzyme activity
that interferes with the Calvin cycle, oxidizing ribulose 1,5-
bisphosphate. In this process, called photorespiration, O
2
is incorporated into ribulose 1,5-bisphosphate, which un-
dergoes additional reactions that actually release CO
2
.
Hence, photorespiration releases CO
2
—essentially undoing
the Calvin cycle which reduces CO
2
to carbohydrate.
The carboxylation and oxidation of ribulose 1,5-bispho-
sphate are catalyzed at the same active site on rubisco, and
compete with each other. Under normal conditions at
25°C, the rate of the carboxylation reaction is four times
that of the oxidation reaction, meaning that 20% of photo-
synthetically fixed carbon is lost to photorespiration. This
loss rises substantially as temperature increases, because the
rate of the oxidation reaction increases with temperature
far faster than the carboxylation reaction rate.
Plants that fix carbon using only C
3
photosynthesis (the
Calvin cycle) are called C
3
plants. In C
3
photosynthesis,
ribulose 1,5-bisphosphate is carboxylated to form a three-
carbon compound via the activity of rubisco. Other plants
use C
4
photosynthesis, in which phosphoenolpyruvate, or
PEP, is carboxylated to form a four-carbon compound
using the enzyme PEP carboxylase. This enzyme has no
oxidation activity, and thus no photorespiration. Further-
more, PEP carboxylase has a much greater affinity for CO
2
than does rubisco. In the C
4
pathway, the four-carbon
compound undergoes further modification, only to be de-
carboxylated. The CO
2
which is released is then captured
by rubisco and drawn into the Calvin cycle. Because an or-
ganic compound is donating the CO
2
, the effective concen-
tration of CO
2
relative to O
2
is increased, and photorespi-
ration is minimized.
The loss of fixed carbon as a result of photorespiration is
not trivial. C
3
plants lose between 25 and 50% of their
photosynthetically fixed carbon in this way. The rate de-
pends largely upon the temperature. In tropical climates,
especially those in which the temperature is often above
28°C, the problem is severe, and it has a major impact on
tropical agriculture.
The C
4
Pathway
Plants that adapted to these warmer environments have
evolved two principal ways that use the C
4
pathway to
deal with this problem. In one approach, plants conduct
C
4
photosynthesis in the mesophyll cells and the Calvin
cycle in the bundle sheath cells. This creates high local
levels of CO
2
to favor the carboxylation reaction of ru-
bisco. These plants are called C
4
plants and include corn,
sugarcane, sorghum, and a number of other grasses. In the
C
4
pathway, the three-carbon metabolite phospho-
enolpyruvate is carboxylated to form the four-carbon
molecule oxaloacetate, which is the first product of CO
2
fixation (figure 10.21). In C
4
plants, oxaloacetate is in turn
converted into the intermediate malate, which is trans-
ported to an adjacent bundle-sheath cell. Inside the bundle-
sheath cell, malate is decarboxylated to produce pyruvate,
releasing CO
2
. Because bundle-sheath cells are imperme-
able to CO
2
, the CO
2
is retained within them in high con-
centrations. Pyruvate returns to the mesophyll cell, where
two of the high-energy bonds in an ATP molecule are
split to convert the pyruvate back into phosphoenolpyru-
vate, thus completing the cycle.
The enzymes that carry out the Calvin cycle in a C
4
plant are located within the bundle-sheath cells, where the
increased CO
2
concentration decreases photorespiration.
Because each CO
2
molecule is transported into the bundle-
sheath cells at a cost of two high-energy ATP bonds, and
because six carbons must be fixed to form a molecule of
glucose, 12 additional molecules of ATP are required to
form a molecule of glucose. In C
4
photosynthesis, the ener-
getic cost of forming glucose is almost twice that of C
3
photosynthesis: 30 molecules of ATP versus 18. Neverthe-
less, C
4
photosynthesis is advantageous in a hot climate:
photorespiration would otherwise remove more than half
of the carbon fixed.
202 Part III Energetics
CO
2
Phosphoenol-
pyruvate (PEP)
Oxaloacetate
Pyruvate Malate
Bundle-
sheath
cell
Mesophyll
cell
PP
i
+ AMP
ATP
P
i
+
Calvin
cycle
Glucose
CO
2
MalatePyruvate
FIGURE 10.21
Carbon fixation in C
4
plants. This process is called the C
4
pathway because the starting material, oxaloacetate, is a molecule
containing four carbons.
The Crassulacean Acid Pathway
A second strategy to decrease photorespiration in hot re-
gions has been adopted by many succulent (water-storing)
plants such as cacti, pineapples, and some members of
about two dozen other plant groups. This mode of initial
carbon fixation is called crassulacean acid metabolism
(CAM), after the plant family Crassulaceae (the
stonecrops or hens-and-chicks), in which it was first dis-
covered. In these plants, the stomata (singular, stoma),
specialized openings in the leaves of all plants through
which CO
2
enters and water vapor is lost, open during the
night and close during the day. This pattern of stomatal
opening and closing is the reverse of that in most plants.
CAM plants open stomata at night and initially fix CO
2
into organic compounds using the C
4
pathway. These or-
ganic compounds accumulate throughout the night and
are decarboxylated during the day to yield high levels of
CO
2
. In the day, these high levels of CO
2
drive the Calvin
cycle and minimize photorespiration. Like C
4
plants,
CAM plants use both C
4
and C
3
pathways. They differ
from C
4
plants in that they use the C
4
pathway at night
and the C
3
pathway during the day within the same cells. In
C
4
plants, the two pathways take place in different cells
(figure 10.22).
Photorespiration results in decreased yields of
photosynthesis. C
4
and CAM plants circumvent this
problem through modifications of leaf architecture and
photosynthetic chemistry that locally increase CO
2
concentrations. C
4
plants isolate CO
2
production
spatially, CAM plants temporally.
Chapter 10 Photosynthesis 203
C
4
plants
Mesophyll
cell
Bundle-
sheath
cell
CO
2
Calvin
cycle
Glucose
C
4
pathway
CO
2
CAM plants
Mesophyll
cell
Night
Day
CO
2
Calvin
cycle
C
4
pathway
Glucose
CO
2
FIGURE 10.22
A comparison of C
4
and CAM plants. Both C
4
and CAM plants utilize the C
4
and the C
3
pathways. In C
4
plants, the pathways are
separated spatially: the C
4
pathway takes place in the mesophyll cells and the C
3
pathway in the bundle-sheath cells. In CAM plants, the
two pathways are separated temporally: the C
4
pathway is utilized at night and the C
3
pathway during the day.
204 Part III Energetics
Chapter 10
Summary Questions Media Resources
10.1 What is photosynthesis?
? Light is used by plants, algae, and some bacteria, in a
process called photosynthesis, to convert atmospheric
carbon (CO
2
) into carbohydrate.
1. Where do the oxygen atoms
in the O2 produced during
photosynthesis come from?
? A series of simple experiments demonstrated that
plants capture energy from light and use it to convert
the carbon atoms of CO
2
and the hydrogen atoms of
water into organic molecules.
2. How did van Helmont
determine that plants do not
obtain their food from the soil?
10.2 Learning about photosynthesis: An experimental journey.
? Light consists of energy packets called photons; the
shorter the wavelength of light, the more its energy.
When photons are absorbed by a pigment, electrons
in the pigment are boosted to a higher energy level.
? Photosynthesis channels photon excitation energy
into a single pigment molecule. In bacteria, that
molecule then donates an electron to an electron
transport chain, which drives a proton pump and
ultimately returns the electron to the pigment.
? Plants employ two photosystems. Light is first
absorbed by photosystem II and passed to
photosystem I, driving a proton pump and bringing
about the chemiosmotic synthesis of ATP.
? When the electron arrives at photosystem I, another
photon of light is absorbed, and energized electrons
are channeled to a primary electron acceptor, which
reduces NADP
+
to NADPH. Use of NADPH rather
than NADH allows plants and algae to keep the
processes of photosynthesis and oxidative respiration
separate from each other.
3. How is the energy of light
captured by a pigment molecule?
Why does light reflected by the
pigment chlorophyll appear
green?
4. What is the function of the
reaction center chlorophyll?
What is the function of the
primary electron acceptor?
5. Explain how photosynthesis in
the sulfur bacteria is a cyclic
process. What is its energy yield
in terms of ATP molecules
synthesized per electron?
6. How do the two photosystems
in plants and algae work? Which
stage generates ATP and which
generates NADPH?
10.3 Pigments capture energy from sunlight.
? The ATP and reducing power produced by the light
reactions are used to fix carbon in a series of reactions
called the Calvin cycle.
? RuBP carboxylase, the enzyme that fixes carbon in
the Calvin cycle, also carries out an oxidative reaction
that uses the products of photosynthesis, a process
called photorespiration.
? Many tropical plants inhibit photorespiration by
expending ATP to increase the intracellular
concentration of CO
2
. This process, called the C
4
pathway, nearly doubles the energetic cost of
synthesizing glucose.
7. In a C
3
plant, where do the
light reactions occur? Where
does the Calvin cycle occur?
8. What is photorespiration?
What advantage do C
4
plants
have over C
3
plants with respect
to photorespiration? What
disadvantage do C
4
plants have
that limits their distribution
primarily to warm regions of the
earth?
10.4 Cells use the energy and reducing power captured by the light reactions to make organic molecules.
http://www.mhhe.com/raven6e http://www.biocourse.com
? Art Activity:
Chloroplast Structure
? Chloroplast
? Energy Conversion
? Photosynthesis
? Light Dependent
Reactions
? Light Independent
Reactions
? Light and
Pigmentation
? The Calvin Cycle
? Photorespiration
? Art Activity:
Electromagnetic
Spectrum