Biology Chapter Notes
Biology Chapter Notes
Chapter Notes
Concept 1.1 Biologists explore life from the microscopic to the global scale
Life’s basic characteristic is a high degree of order.
Each level of biological organization has emergent properties.
Biological organization is based on a hierarchy of structural levels, each building on the
levels below.
At the lowest level are atoms that are ordered into complex biological molecules.
Biological molecules are organized into structures called organelles, the components of
cells.
Cells are the fundamental unit of structure and function of living things.
Some organisms consist of a single cell; others are multicellular aggregates of specialized
cells.
Whether multicellular or unicellular, all organisms must accomplish the same functions:
uptake and processing of nutrients, excretion of wastes, response to environmental
stimuli, and reproduction.
Multicellular organisms exhibit three major structural levels above the cell: similar
cells are grouped into tissues, several tissues coordinate to form organs, and several
organs form an organ system.
Biology Chapter Notes
For example, to coordinate locomotory movements, sensory information travels from
sense organs to the brain, where nervous tissues composed of billions of interconnected
neurons—supported by connective tissue—coordinate signals that travel via other neurons
to the individual muscle cells.
Organisms belong to populations, localized groups of organisms belonging to the same
species.
Populations of several species in the same area comprise a biological community.
Populations interact with their physical environment to form an ecosystem.
The biosphere consists of all the environments on Earth that are inhabited by life.
Organisms interact continuously with their environment.
Each organism interacts with its environment, which includes other organisms as well as
nonliving factors.
Both organism and environment are affected by the interactions between them.
The dynamics of any ecosystem include two major processes: the cycling of nutrients and
the flow of energy from sunlight to producers to consumers.
In most ecosystems, producers are plants and other photosynthetic organisms that
convert light energy to chemical energy.
Consumers are organisms that feed on producers and other consumers.
All the activities of life require organisms to perform work, and work requires a source of
energy.
The exchange of energy between an organism and its environment often involves the
transformation of energy from one form to another.
In all energy transformations, some energy is lost to the surroundings as heat.
In contrast to chemical nutrients, which recycle within an ecosystem, energy flows
through an ecosystem, usually entering as light and exiting as heat.
Cells are an organism’s basic unit of structure and function.
The cell is the lowest level of structure that is capable of performing all the activities of
life.
For example, the ability of cells to divide is the basis of all reproduction and the basis
of growth and repair of multicellular organisms.
Understanding how cells work is a major research focus of modern biology.
At some point, all cells contain deoxyribonucleic acid, or DNA, the heritable material that
directs the cell’s activities.
DNA is the substance of genes, the units of inheritance that transmit information from
parents to offspring.
Each of us began life as a single cell stocked with DNA inherited from our parents.
DNA in human cells is organized into chromosomes.
Each chromosome has one very long DNA molecule, with hundreds or thousands of
genes arranged along its length.
The DNA of chromosomes replicates as a cell prepares to divide.
Each of the two cellular offspring inherits a complete set of genes.
In each cell, the genes along the length of DNA molecules encode the information for
building the cell’s other molecules.
DNA thus directs the development and maintenance of the entire organism.
Biology Chapter Notes
Most genes program the cell’s production of proteins.
Each DNA molecule is made up of two long chains arranged in a double helix.
Each link of a chain is one of four nucleotides, encoding the cell’s information in
chemical letters.
The sequence of nucleotides along each gene codes for a specific protein with a unique
shape and function.
Almost all cellular activities involve the action of one or more proteins.
DNA provides the heritable blueprints, but proteins are the tools that actually build and
maintain the cell.
All forms of life employ essentially the same genetic code.
Because the genetic code is universal, it is possible to engineer cells to produce
proteins normally found only in some other organism.
The library of genetic instructions that an organism inherits is called its genome.
The chromosomes of each human cell contain about 3 billion nucleotides, including
genes coding for more than 70,000 kinds of proteins, each with a specific function.
Every cell is enclosed by a membrane that regulates the passage of material between a cell
and its surroundings.
Every cell uses DNA as its genetic material.
There are two basic types of cells: prokaryotic cells and eukaryotic cells.
The cells of the microorganisms called bacteria and archaea are prokaryotic.
All other forms of life have more complex eukaryotic cells.
Eukaryotic cells are subdivided by internal membranes into various organelles.
In most eukaryotic cells, the largest organelle is the nucleus, which contains the cell’s
DNA as chromosomes.
The other organelles are located in the cytoplasm, the entire region between the
nucleus and outer membrane of the cell.
Prokaryotic cells are much simpler and smaller than eukaryotic cells.
In a prokaryotic cell, DNA is not separated from the cytoplasm in a nucleus.
There are no membrane-enclosed organelles in the cytoplasm.
All cells, regardless of size, shape, or structural complexity, are highly ordered structures
that carry out complicated processes necessary for life.
Concept 1.2 Biological systems are much more than the sum of their parts
“The whole is greater than the sum of its parts.”
The combination of components can form a more complex organization called a system.
Examples of biological systems are cells, organisms, and ecosystems.
Consider the levels of life.
With each step upward in the hierarchy of biological order, novel properties emerge
that are not present at lower levels.
These emergent properties result from the arrangements and interactions between
components as complexity increases.
A cell is much more than a bag of molecules.
Biology Chapter Notes
Our thoughts and memories are emergent properties of a complex network of neurons.
This theme of emergent properties accents the importance of structural arrangement.
The emergent properties of life are not supernatural or unique to life but simply reflect a
hierarchy of structural organization.
The emergent properties of life are particularly challenging because of the unparalleled
complexity of living systems.
The complex organization of life presents a dilemma to scientists seeking to understand
biological processes.
We cannot fully explain a higher level of organization by breaking it down into its
component parts.
At the same time, it is futile to try to analyze something as complex as an organism or
cell without taking it apart.
Reductionism, reducing complex systems to simpler components, is a powerful strategy in
biology.
The Human Genome Project—the sequencing of the genome of humans and many
other species—is heralded as one of the greatest scientific achievements ever.
Research is now moving on to investigate the function of genes and the coordination of
the activity of gene products.
Biologists are beginning to complement reductionism with new strategies for
understanding the emergent properties of life—how all of the parts of biological systems
are functionally integrated.
The ultimate goal of systems biology is to model the dynamic behavior of whole
biological systems.
Accurate models allow biologists to predict how a change in one or more variables will
impact other components and the whole system.
Scientists investigating ecosystems pioneered this approach in the 1960s with elaborate
models diagramming the interactions of species and nonliving components in ecosystems.
Systems biology is now becoming increasingly important in cellular and molecular
biology, driven in part by the deluge of data from the sequencing of genomes and our
increased understanding of protein functions.
In 2003, a large research team published a network of protein interactions within a cell
of a fruit fly.
Three key research developments have led to the increased importance of systems
biology.
1. High-throughput technology. Systems biology depends on methods that can
analyze biological materials very quickly and produce enormous amounts of data.
An example is the automatic DNA-sequencing machines used by the Human
Genome Project.
2. Bioinformatics. The huge databases from high-throughput methods require
computing power, software, and mathematical models to process and integrate
information.
3. Interdisciplinary research teams. Systems biology teams may include engineers,
medical scientists, physicists, chemists, mathematicians, and computer scientists as
well as biologists.
Regulatory mechanisms ensure a dynamic balance in living systems.
Concept 1.3 Biologists explore life across its great diversity of species
Biology can be viewed as having two dimensions: a “vertical” dimension covering the
size scale from atoms to the biosphere and a “horizontal” dimension that stretches across
the diversity of life.
The latter includes not only present-day organisms, but also those that have existed
throughout life’s history.
Living things show diversity and unity.
Life is enormously diverse.
Biologists have identified and named about 1.8 million species.
This diversity includes 5,200 known species of prokaryotes, 100,000 fungi, 290,000
plants, 50,000 vertebrates, and 1,000,000 insects.
Thousands of newly identified species are added each year.
Estimates of the total species count range from 10 million to more than 200 million.
In the face of this complexity, humans are inclined to categorize diverse items into a
smaller number of groups.
Taxonomy is the branch of biology that names and classifies species into a hierarchical
order.
Until the past decade, biologists divided the diversity of life into five kingdoms.
New methods, including comparisons of DNA among organisms, have led to a
reassessment of the number and boundaries of the kingdoms.
Various classification schemes now include six, eight, or even dozens of kingdoms.
Coming from this debate has been the recognition that there are three even higher levels
of classifications, the domains.
The three domains are Bacteria, Archaea, and Eukarya.
The first two domains, domain Bacteria and domain Archaea, consist of prokaryotes.
All the eukaryotes are now grouped into various kingdoms of the domain Eukarya.
Biology Chapter Notes
The recent taxonomic trend has been to split the single-celled eukaryotes and their
close relatives into several kingdoms.
Domain Eukarya also includes the three kingdoms of multicellular eukaryotes: the
kingdoms Plantae, Fungi, and Animalia.
These kingdoms are distinguished partly by their modes of nutrition.
Most plants produce their own sugars and food by photosynthesis.
Most fungi are decomposers that absorb nutrients by breaking down dead organisms
and organic wastes.
Animals obtain food by ingesting other organisms.
Underlying the diversity of life is a striking unity, especially at the lower levels of
organization.
The universal genetic language of DNA unites prokaryotes and eukaryotes.
Among eukaryotes, unity is evident in many details of cell structure.
Above the cellular level, organisms are variously adapted to their ways of life.
How do we account for life’s dual nature of unity and diversity?
The process of evolution explains both the similarities and differences among living
things.
Living organisms and the world they live in are subject to the basic laws of physics and
chemistry.
Biology is a multidisciplinary science, drawing on insights from other sciences.
Life can be organized into a hierarchy of structural levels.
At each successive level, additional emergent properties appear.
Concept 2.1 Matter consists of chemical elements in pure form and in combinations
called compounds
Organisms are composed of matter.
Matter is anything that takes up space and has mass.
Matter is made up of elements.
An element is a substance that cannot be broken down into other substances by chemical
reactions.
There are 92 naturally occurring elements.
Each element has a unique symbol, usually the first one or two letters of the name.
Some of the symbols are derived from Latin or German names.
A compound is a substance consisting of two or more elements in a fixed ratio.
Table salt (sodium chloride or NaCl) is a compound with equal numbers of atoms of
the elements chlorine and sodium.
While pure sodium is a metal and chlorine is a gas, they combine to form an edible
compound. This change in characteristics when elements combine to form a compound
is an example of an emergent property.
25 chemical elements are essential to life.
About 25 of the 92 natural elements are known to be essential for life.
Four elements—carbon (C), oxygen (O), hydrogen (H), and nitrogen (N)—make up
96% of living matter.
Most of the remaining 4% of an organism’s weight consists of phosphorus (P), sulfur
(S), calcium (Ca), and potassium (K).
Trace elements are required by an organism but only in minute quantities.
Some trace elements, like iron (Fe), are required by all organisms.
Other trace elements are required by only some species.
For example, a daily intake of 0.15 milligrams of iodine is required for normal
activity of the human thyroid gland.
Because water is the substance that makes life possible on Earth, astronomers hope to find
evidence of water on newly discovered planets orbiting distant stars.
Life on Earth began in water and evolved there for 3 billion years before colonizing the
land.
Even terrestrial organisms are tied to water.
Most cells are surrounded by water.
Cells are about 70–95% water.
Water is a reactant in many of the chemical reactions of life.
Water is the only common substance that exists in the natural world in all three physical
states of matter: solid ice, liquid water, and water vapor.
Concept 3.2 Four emergent properties of water contribute to Earth’s fitness for life
Organisms depend on the cohesion of water molecules.
The hydrogen bonds joining water molecules are weak, about 1/20 as strong as covalent
bonds.
They form, break, and reform with great frequency. Each hydrogen bond lasts only a few
trillionths of a second.
At any instant, a substantial percentage of all water molecules are bonded to their
neighbors, creating a high level of structure.
Collectively, hydrogen bonds hold water together, a phenomenon called cohesion.
Cohesion among water molecules plays a key role in the transport of water and dissolved
nutrients against gravity in plants.
Biology Chapter Notes
Water molecules move from the roots to the leaves of a plant through water-
conducting vessels.
As water molecules evaporate from a leaf, other water molecules from vessels in the
leaf replace them.
Hydrogen bonds cause water molecules leaving the vessels to tug on molecules farther
down.
This upward pull is transmitted down to the roots.
Adhesion, clinging of one substance to another, contributes too, as water adheres to
the wall of the vessels.
Surface tension, a measure of the force necessary to stretch or break the surface of a
liquid, is related to cohesion.
Water has a greater surface tension than most other liquids because hydrogen bonds
among surface water molecules resist stretching or breaking the surface.
Water behaves as if covered by an invisible film.
Some animals can stand, walk, or run on water without breaking the surface.
Water moderates temperatures on Earth.
Water stabilizes air temperatures by absorbing heat from warmer air and releasing heat to
cooler air.
Water can absorb or release relatively large amounts of heat with only a slight change in
its own temperature.
Atoms and molecules have kinetic energy, the energy of motion, because they are always
moving.
The faster a molecule moves, the more kinetic energy it has.
Heat is a measure of the total quantity of kinetic energy due to molecular motion in a
body of matter.
Temperature measures the intensity of heat in a body of matter due to the average
kinetic energy of molecules.
As the average speed of molecules increases, a thermometer will record an increase in
temperature.
Heat and temperature are related, but not identical.
When two objects of different temperatures come together, heat passes from the warmer
object to the cooler object until the two are the same temperature.
Molecules in the cooler object speed up at the expense of kinetic energy of the warmer
object.
Ice cubes cool a glass of pop by absorbing heat from the pop as the ice melts.
In most biological settings, temperature is measured on the Celsius scale (°C).
At sea level, water freezes at 0°C and boils at 100°C.
Human body temperature is typically 37°C.
While there are several ways to measure heat energy, one convenient unit is the calorie
(cal).
One calorie is the amount of heat energy necessary to raise the temperature of one g of
water by 1°C.
A calorie is released when 1 g of water cools by 1°C.
In many biological processes, the kilocalorie (kcal) is more convenient.
Biology Chapter Notes
A kilocalorie is the amount of heat energy necessary to raise the temperature of 1000 g
of water by 1°C.
Another common energy unit, the joule (J), is equivalent to 0.239 cal.
Water stabilizes temperature because it has a high specific heat.
The specific heat of a substance is the amount of heat that must be absorbed or lost for 1
g of that substance to change its temperature by 1°C.
By definition, the specific heat of water is 1 cal per gram per degree Celsius or 1
cal/g/°C.
Water has a high specific heat compared to other substances.
For example, ethyl alcohol has a specific heat of 0.6 cal/g/°C.
The specific heat of iron is 1/10 that of water.
Water resists changes in temperature because of its high specific heat.
In other words, water absorbs or releases a relatively large quantity of heat for each
degree of temperature change.
Water’s high specific heat is due to hydrogen bonding.
Heat must be absorbed to break hydrogen bonds, and heat is released when hydrogen
bonds form.
Investment of one calorie of heat causes relatively little change to the temperature of
water because much of the energy is used to disrupt hydrogen bonds, not speed up the
movement of water molecules.
Water’s high specific heat has effects that range from the level of the whole Earth to the
level of individual organisms.
A large body of water can absorb a large amount of heat from the sun in daytime
during the summer and yet warm only a few degrees.
At night and during the winter, the warm water will warm cooler air.
Therefore, ocean temperatures and coastal land areas have more stable temperatures
than inland areas.
Living things are made primarily of water. Consequently, they resist changes in
temperature better than they would if composed of a liquid with a lower specific heat.
The transformation of a molecule from a liquid to a gas is called vaporization or
evaporation.
This occurs when the molecule moves fast enough to overcome the attraction of other
molecules in the liquid.
Even in a low-temperature liquid (with low average kinetic energy), some molecules
are moving fast enough to evaporate.
Heating a liquid increases the average kinetic energy and increases the rate of
evaporation.
Heat of vaporization is the quantity of heat that a liquid must absorb for 1 g of it to be
converted from liquid to gas.
Water has a relatively high heat of vaporization, requiring about 580 cal of heat to
evaporate 1 g of water at room temperature.
This is double the heat required to vaporize the same quantity of alcohol or ammonia.
This is because hydrogen bonds must be broken before a water molecule can evaporate
from the liquid.
Water’s high heat of vaporization moderates climate.
Concept 3.3 Dissociation of water molecules leads to acidic and basic conditions that
affect living organisms
Occasionally, a hydrogen atom participating in a hydrogen bond between two water
molecules shifts from one molecule to the other.
The hydrogen atom leaves its electron behind and is transferred as a single proton—a
hydrogen ion (H+).
−
The water molecule that lost the proton is now a hydroxide ion (OH ).
+
The water molecule with the extra proton is now a hydronium ion (H3O ).
A simplified way to view this process is to say that a water molecule dissociates into a
hydrogen ion and a hydroxide ion:
+ −
H2O <=> H + OH
This reaction is reversible.
At equilibrium, the concentration of water molecules greatly exceeds that of H+ and OH−.
In pure water, only one water molecule in every 554 million is dissociated.
+ − −7 °
At equilibrium, the concentration of H or OH is 10 M (at 25 C).
Although the dissociation of water is reversible and statistically rare, it is very important
in the chemistry of life.
Because hydrogen and hydroxide ions are very reactive, changes in their concentrations
can drastically affect the chemistry of a cell.
Adding certain solutes, called acids and bases, disrupts the equilibrium and modifies the
concentrations of hydrogen and hydroxide ions.
The pH scale is used to describe how acidic or basic a solution is.
Organisms are sensitive to changes in pH.
An acid is a substance that increases the hydrogen ion concentration in a solution.
When hydrochloric acid is added to water, hydrogen ions dissociate from chloride
ions: HCl -> H+ + Cl−
Addition of an acid makes a solution more acidic.
Any substance that reduces the hydrogen ion concentration in a solution is a base.
Some bases reduce the H+ concentration directly by accepting hydrogen ions.
Biology Chapter Notes
Ammonia (NH3) acts as a base when the nitrogen’s unshared electron pair attracts a
hydrogen ion from the solution, creating an ammonium ion (NH4+).
+ +
NH3 + H <=> NH4
Other bases reduce H+ indirectly by dissociating to OH−, which then combines with H+ to
form water.
−
NaOH -> Na + OH
+
OH− + H+ -> H2O
Solutions with more OH− than H+ are basic solutions.
Solutions with more H+ than OH− are acidic solutions.
Solutions in which concentrations of OH− and H+ are equal are neutral solutions.
Some acids and bases (HCl and NaOH) are strong acids or bases.
These molecules dissociate completely in water.
Other acids and bases (NH3) are weak acids or bases.
For these molecules, the binding and release of hydrogen ions are reversible.
At equilibrium, there will be a fixed ratio of products to reactants.
Carbonic acid (H2CO3) is a weak acid:
H2CO3 <=> HCO3− + H+
At equilibrium, 1% of the H2CO3 molecules will be dissociated.
In any solution, the product of the H+ and OH− concentrations is constant at 10−14.
Brackets ([H+] and [OH−]) indicate the molar concentration of the enclosed substance.
+ − −14
[H ] [OH ] = 10
+ −7 − −7
In a neutral solution, [H ] = 10 M and [OH ] = 10 M
Adding acid to− a solution shifts the balance between H+ and OH− toward H+ and leads to a
decline in OH .
+ −5 − −9
If [H ] = 10 M, then [OH ] = 10 M
Hydroxide concentrations decline because some of the additional acid combines with
hydroxide to form water.
Adding a base does the opposite, increasing OH − concentration and lowering H+
concentration.
The H+ and OH− concentrations of solutions can vary by a factor of 100 trillion or more.
To express this variation more conveniently, the H+ and OH− concentrations are typically
expressed via the pH scale.
The pH scale, ranging from 1 to 14, compresses the range of concentrations by
employing logarithms.
+ + −pH
pH = − log [H ] or [H ] = 10
In a neutral solution, [H+] = 10−7 M, and the pH = 7.
Values for pH decline as [H+] increase.
While the pH scale is based on [H+], values for [OH−] can be easily calculated from the
product relationship.
The pH of a neutral solution is 7.
Acidic solutions have pH values less than 7, and basic solutions have pH values greater
than 7.
Most biological fluids have pH values in the range of 6 to 8.
Biology Chapter Notes
However, the human stomach has strongly acidic digestive juice with a pH of about 2.
Each pH unit represents a tenfold difference in H+ and OH− concentrations.
+ −
A small change in pH actually indicates a substantial change in H and OH
concentrations.
The chemical processes in the cell can be disrupted by changes to the H + and OH−
concentrations away from their normal values, usually near pH 7.
To maintain cellular pH values at a constant level, biological fluids have buffers.
Buffers resist changes to the pH of a solution when H+ or OH− is added to the solution.
Buffers accept hydrogen ions from the solution when they are in excess and donate
hydrogen ions when they have been depleted.
Buffers typically consist of a weak acid and its corresponding base.
One important buffer in human blood and other biological solutions is carbonic acid,
which dissociates to yield a bicarbonate ion and a hydrogen ion.
The chemical equilibrium between carbonic acid and bicarbonate acts as a pH
regulator. +The equilibrium shifts left or right as other metabolic processes add or
remove H from the solution.
Acid precipitation threatens the fitness of the environment.
Acid precipitation is a serious assault on water quality in some industrialized areas.
Uncontaminated rain has a slightly acidic pH of 5.6.
The acid is a product of the formation of carbonic acid from carbon dioxide and water.
Acid precipitation occurs when rain, snow, or fog has a pH that is more acidic than 5.6.
Acid precipitation is caused primarily by sulfur oxides and nitrogen oxides in the
atmosphere.
These molecules react with water to form strong acids that fall to the surface with rain
or snow.
The major source of these oxides is the burning of fossil fuels (coal, oil, and gas) in
factories and automobiles.
The presence of tall smokestacks allows this pollution to spread from its site of origin to
contaminate relatively pristine areas thousands of kilometers away.
In 2001, rain in the Adirondack Mountains of upstate New York had an average pH of
4.3.
The effects of acids in lakes and streams are more pronounced in the spring during
snowmelt.
As the surface snows melt and drain down through the snowfield, the meltwater
accumulates acid and brings it into lakes and streams all at once.
The pH of early meltwater may be as low as 3.
Acid precipitation has a great impact on the eggs and the early developmental stages of
aquatic organisms that are abundant in the spring.
Thus, strong acidity can alter the structure of molecules and impact ecological
communities.
Direct impacts of acid precipitation on forests and terrestrial life are more controversial.
However, acid precipitation can impact soils by affecting the solubility of soil minerals.
Acid precipitation can wash away key soil buffers and plant nutrients such as calcium
and magnesium ions.
Biology Chapter Notes
It can also increase the concentrations of compounds such as aluminum to toxic levels.
This has done major damage to forests in Europe and substantial damage of forests in
North America.
Progress has been made in reducing acid precipitation.
Concept 4.2 Carbon atoms can form diverse molecules by bonding to four other atoms
· With a total of 6 electrons, a carbon atom has 2 in the first electron shell and 4 in the second shell.
· Carbon has little tendency to form ionic bonds by losing or gaining 4 electrons to complete its valence
shell.
· Instead, carbon usually completes its valence shell by sharing electrons with other atoms in four covalent
bonds.
· This tetravalence by carbon makes large, complex molecules possible.
· When carbon forms covalent bonds with four other atoms, they are arranged at the corners of an
imaginary tetrahedron with bond angles of 109.5°.
· In molecules with multiple carbons, every carbon bonded to four other atoms has a tetrahedral shape.
· However, when two carbon atoms are joined by a double bond, all bonds around those carbons are in the
same plane and have a flat, three-dimensional structure.
· The three-dimensional shape of an organic molecule determines its function.
· The electron configuration of carbon makes it capable of forming covalent bonds with many different
elements.
· The valences of carbon and its partners can be viewed as the building code that governs the architecture
of organic molecules.
· In carbon dioxide, one carbon atom forms two double bonds with two different oxygen atoms.
· In the structural formula, O=C=O, each line represents a pair of shared electrons. This arrangement
completes the valence shells of all atoms in the molecule.
· While CO2 can be classified as either organic or inorganic, its importance to the living world is clear.
Biology Chapter Notes
· CO2 is the source of carbon for all organic molecules found in organisms. It is usually fixed into organic
molecules by the process of photosynthesis.
· Urea, CO(NH2)2, is another simple organic molecule in which each atom forms covalent bonds to
complete its valence shell.
Variation in carbon skeletons contributes to the diversity of organic molecules.
· Carbon chains form the skeletons of most organic molecules.
· The skeletons vary in length and may be straight, branched, or arranged in closed rings.
· The carbon skeletons may include double bonds.
· Atoms of other elements can be bonded to the atoms of the carbon skeleton.
· Hydrocarbons are organic molecules that consist of only carbon and hydrogen atoms.
· Hydrocarbons are the major component of petroleum, a fossil fuel that consists of the partially
decomposed remains of organisms that lived millions of years ago.
· Fats are biological molecules that have long hydrocarbon tails attached to a nonhydrocarbon component.
· Petroleum and fat are hydrophobic compounds that cannot dissolve in water because of their many
nonpolar carbon-to-hydrogen bonds.
· Isomers are compounds that have the same molecular formula but different structures and, therefore,
different chemical properties.
· For example, butane and isobutane have the same molecular formula, C4H10, but butane has a straight
skeleton and isobutane has a branched skeleton.
· The two butanes are structural isomers, molecules that have the same molecular formula but differ in
the covalent arrangement of atoms.
· Geometric isomers are compounds with the same covalent partnerships that differ in the spatial
arrangement of atoms around a carbon–carbon double bond.
· The double bond does not allow atoms to rotate freely around the bond axis.
· The biochemistry of vision involves a light-induced change in the structure of rhodopsin in the retina
from one geometric isomer to another.
· Enantiomers are molecules that are mirror images of each other.
· Enantiomers are possible when four different atoms or groups of atoms are bonded to a carbon.
· In this case, the four groups can be arranged in space in two different ways that are mirror images.
· They are like left-handed and right-handed versions of the molecule.
· Usually one is biologically active, while the other is inactive.
Concept 4.3 Functional groups are the parts of molecules involved in chemical reactions
· The components of organic molecules that are most commonly involved in chemical reactions are known
as functional groups.
· If we consider hydrocarbons to be the simplest organic molecules, we can view functional groups as
attachments that replace one or more of the hydrogen atoms bonded to the carbon skeleton of the
hydrocarbon.
· Each functional group behaves consistently from one organic molecule to another.
· The number and arrangement of functional groups help give each molecule its unique properties.
· As an example, the basic structure of testosterone (a male sex hormone) and estradiol (a female sex
hormone) is the same.
· Both are steroids with four fused carbon rings, but they differ in the functional groups attached to the
rings.
· These functional groups interact with different targets in the body.
· There are six functional groups that are most important to the chemistry of life: hydroxyl, carbonyl,
carboxyl, amino, sulfhydryl, and phosphate groups.
· All are hydrophilic and increase the solubility of organic compounds in water.
· In a hydroxyl group (—OH), a hydrogen atom forms a polar covalent bond with an oxygen atom, which
forms a polar covalent bond to the carbon skeleton.
· Because of these polar covalent bonds, hydroxyl groups increase the solubility of organic molecules.
· Organic compounds with hydroxyl groups are alcohols, and their names typically end in -ol.
· A carbonyl group (>CO) consists of an oxygen atom joined to the carbon skeleton by a double bond.
· If the carbonyl group is on the end of the skeleton, the compound is an aldehyde.
· If the carbonyl group is within the carbon skeleton, then the compound is a ketone.
· Isomers with aldehydes versus ketones have different properties.
· A carboxyl group (—COOH) consists of a carbon atom with a double bond to an oxygen atom and a
single bond to the oxygen of a hydroxyl group.
Biology Chapter Notes
· Compounds with carboxyl groups are carboxylic acids.
· A carboxyl group acts as an acid because the combined electronegativities of the two adjacent oxygen
atoms increase the dissociation of hydrogen as an ion (H+).
· An amino group (—NH2) consists of a nitrogen atom bonded to two hydrogen atoms and the carbon
skeleton.
· Organic compounds with amino groups are amines.
· The amino group acts as a base because the amino group can pick up a hydrogen ion (H+) from the
solution.
· Amino acids, the building blocks of proteins, have amino and carboxyl groups.
· A sulfhydryl group (—SH) consists of a sulfur atom bonded to a hydrogen atom and to the backbone.
· This group resembles a hydroxyl group in shape.
· Organic molecules with sulfhydryl groups are thiols.
· Two sulfhydryl groups can interact to help stabilize the structure of proteins.
· A phosphate group (—OPO32−) consists of a phosphorus atom bound to four oxygen atoms (three with
single bonds and one with a double bond).
· A phosphate group connects to the carbon backbone via one of its oxygen atoms.
· Phosphate groups are anions with two negative charges, as two protons have dissociated from the
oxygen atoms.
· One function of phosphate groups is to transfer energy between organic molecules.
· Adenosine triphosphate, or ATP, is the primary energy-transferring molecule in living cells.
These are the chemical elements of life.
· Living matter consists mainly of carbon, oxygen, hydrogen, and nitrogen, with smaller amounts of sulfur
and phosphorus.
· These elements are linked by strong covalent bonds.
· Carbon, with its four covalent bonds, is the basic building block in molecular architecture.
· The great diversity of organic molecules with their special properties emerges from the unique
arrangement of the carbon skeleton and the functional groups attached to the skeleton.
Within cells, small organic molecules are joined together to form larger molecules.
These large macromolecules may consist of thousands of covalently bonded atoms and
weigh more than 100,000 daltons.
The four major classes of macromolecules are carbohydrates, lipids, proteins, and nucleic
acids.
Concept 5.4 Proteins have many structures, resulting in a wide range of functions
Proteins account for more than 50% of the dry mass of most cells. They are instrumental
in almost everything that an organism does.
Protein functions include structural support, storage, transport, cellular signaling,
movement, and defense against foreign substances.
Most important, protein enzymes function as catalysts in cells, regulating metabolism
by selectively accelerating chemical reactions without being consumed.
Humans have tens of thousands of different proteins, each with a specific structure and
function.
Proteins are the most structurally complex molecules known.
Each type of protein has a complex three-dimensional shape or conformation.
All protein polymers are constructed from the same set of 20 amino acid monomers.
Polymers of proteins are called polypeptides.
A protein consists of one or more polypeptides folded and coiled into a specific
conformation.
Amino acids are the monomers from which proteins are constructed.
Amino acids are organic molecules with both carboxyl and amino groups.
At the center of an amino acid is an asymmetric carbon atom called the alpha carbon.
Four components are attached to the alpha carbon: a hydrogen atom, a carboxyl group, an
amino group, and a variable R group (or side chain).
Different R groups characterize the 20 different amino acids.
R groups may be as simple as a hydrogen atom (as in the amino acid glycine), or it may
be a carbon skeleton with various functional groups attached (as in glutamine).
The physical and chemical properties of the R group determine the unique characteristics
of a particular amino acid.
One group of amino acids has hydrophobic R groups.
Another group of amino acids has polar R groups that are hydrophilic.
A third group of amino acids includes those with functional groups that are charged
(ionized) at cellular pH.
Some acidic R groups are negative in charge due to the presence of a carboxyl
group.
Basic R groups have amino groups that are positive in charge.
Note that all amino acids have carboxyl and amino groups. The terms acidic and
basic in this context refer only to these groups in the R groups.
Biology Chapter Notes
Amino acids are joined together when a dehydration reaction removes a hydroxyl group
from the carboxyl end of one amino acid and a hydrogen from the amino group of
another.
The resulting covalent bond is called a peptide bond.
Repeating the process over and over creates a polypeptide chain.
At one end is an amino acid with a free amino group (the N-terminus) and at the other
is an amino acid with a free carboxyl group (the C-terminus).
Polypeptides range in size from a few monomers to thousands.
Each polypeptide has a unique linear sequence of amino acids.
The amino acid sequence of a polypeptide can be determined.
Frederick Sanger and his colleagues at Cambridge University determined the amino acid
sequence of insulin in the 1950s.
Sanger used protein-digesting enzymes and other catalysts to hydrolyze the insulin at
specific places.
The fragments were then separated by a technique called chromatography.
Hydrolysis by another agent broke the polypeptide at different sites, yielding a second
group of fragments.
Sanger used chemical methods to determine the sequence of amino acids in the small
fragments.
He then searched for overlapping regions among the pieces obtained by hydrolyzing
with the different agents.
After years of effort, Sanger was able to reconstruct the complete primary structure of
insulin.
Most of the steps in sequencing a polypeptide have since been automated.
Protein conformation determines protein function.
A functional protein consists of one or more polypeptides that have been twisted, folded,
and coiled into a unique shape.
It is the order of amino acids that determines what the three-dimensional conformation of
the protein will be.
A protein’s specific conformation determines its function.
When a cell synthesizes a polypeptide, the chain generally folds spontaneously to assume
the functional conformation for that protein.
The folding is reinforced by a variety of bonds between parts of the chain, which in turn
depend on the sequence of amino acids.
Many proteins are globular, while others are fibrous in shape.
In almost every case, the function of a protein depends on its ability to recognize and bind
to some other molecule.
For example, an antibody binds to a particular foreign substance.
An enzyme recognizes and binds to a specific substrate, facilitating a chemical
reaction.
Natural signal molecules called endorphins bind to specific receptor proteins on the
surface of brain cells in humans, producing euphoria and relieving pain.
Morphine, heroin, and other opiate drugs mimic endorphins because they are
similar in shape and can bind to the brain’s endorphin receptors.
Biology Chapter Notes
The function of a protein is an emergent property resulting from its specific molecular
order.
Three levels of structure—primary, secondary, and tertiary structures—organize the
folding within a single polypeptide.
Quaternary structure arises when two or more polypeptides join to form a protein.
The primary structure of a protein is its unique sequence of amino acids.
Lysozyme, an enzyme that attacks bacteria, consists of 129 amino acids.
The precise primary structure of a protein is determined by inherited genetic
information.
Even a slight change in primary structure can affect a protein’s conformation and ability
to function.
The substitution of one amino acid (valine) for the normal one (glutamic acid) at a
particular position in the primary structure of hemoglobin, the protein that carries
oxygen in red blood cells, can cause sickle-cell disease, an inherited blood disorder.
The abnormal hemoglobins crystallize, deforming the red blood cells into a sickle
shape and clogging capillaries.
Most proteins have segments of their polypeptide chains repeatedly coiled or folded.
These coils and folds are referred to as secondary structure and result from hydrogen
bonds between the repeating constituents of the polypeptide backbone.
The weakly positive hydrogen atom attached to the nitrogen atom has an affinity for
the oxygen atom of a nearby peptide bond.
Each hydrogen bond is weak, but the sum of many hydrogen bonds stabilizes the
structure of part of the protein.
Typical secondary structures are coils (an alpha helix) or folds (beta pleated sheets).
The structural properties of silk are due to beta pleated sheets.
The presence of so many hydrogen bonds makes each silk fiber stronger than a steel
strand of the same weight.
Tertiary structure is determined by interactions among various R groups.
These interactions include hydrogen bonds between polar and/or charged areas, ionic
bonds between charged R groups, and hydrophobic interactions and van der Waals
interactions among hydrophobic R groups.
While these three interactions are relatively weak, strong covalent bonds called
disulfide bridges that form between the sulfhydryl groups (SH) of two cysteine
monomers act to rivet parts of the protein together.
Quaternary structure results from the aggregation of two or more polypeptide subunits.
Collagen is a fibrous protein of three polypeptides that are supercoiled like a rope.
This provides structural strength for collagen’s role in connective tissue.
Hemoglobin is a globular protein with quaternary structure.
It consists of four polypeptide subunits: two alpha and two beta chains.
Both types of subunits consist primarily of alpha-helical secondary structure.
Each subunit has a nonpeptide heme component with an iron atom that binds oxygen.
What are the key factors determining protein conformation?
A polypeptide chain of a given amino acid sequence can spontaneously arrange itself into
a 3D shape determined and maintained by the interactions responsible for secondary and
tertiary structure.
Biology Chapter Notes
The folding occurs as the protein is being synthesized within the cell.
However, protein conformation also depends on the physical and chemical conditions of
the protein’s environment.
Alterations in pH, salt concentration, temperature, or other factors can unravel or
denature a protein.
These forces disrupt the hydrogen bonds, ionic bonds, and disulfide bridges that
maintain the protein’s shape.
Most proteins become denatured if the are transferred to an organic solvent. The
polypeptide chain refolds so that its hydrophobic regions face outward, toward the
solvent.
Denaturation can also be caused by heat, which disrupts the weak interactions that
stabilize conformation.
This explains why extremely high fevers can be fatal. Proteins in the blood become
denatured by the high body temperatures.
Some proteins can return to their functional shape after denaturation, but others cannot,
especially in the crowded environment of the cell.
Biochemists now know the amino acid sequences of more than 875,000 proteins and the
3D shapes of about 7,000.
Nevertheless, it is still difficult to predict the conformation of a protein from its
primary structure alone.
Most proteins appear to undergo several intermediate stages before reaching their
“mature” configuration.
The folding of many proteins is assisted by chaperonins or chaperone proteins.
Chaperonins do not specify the final structure of a polypeptide but rather work to
segregate and protect the polypeptide while it folds spontaneously.
At present, scientists use X-ray crystallography to determine protein conformation.
This technique requires the formation of a crystal of the protein being studied.
The pattern of diffraction of an X-ray by the atoms of the crystal can be used to determine
the location of the atoms and to build a computer model of its structure.
Nuclear magnetic resonance (NMR) spectroscopy has recently been applied to this
problem.
This method does not require protein crystallization.
Concept 6.1 To study cells, biologists use microscopes and the tools of biochemistry
The discovery and early study of cells progressed with the invention of microscopes in
1590 and their improvement in the 17th century.
In a light microscope (LM), visible light passes through the specimen and then through
glass lenses.
The lenses refract light such that the image is magnified into the eye or onto a video
screen.
Microscopes vary in magnification and resolving power.
Magnification is the ratio of an object’s image to its real size.
Resolving power is a measure of image clarity.
It is the minimum distance two points can be separated and still be distinguished as
two separate points.
Resolution is limited by the shortest wavelength of the radiation used for imaging.
The minimum resolution of a light microscope is about 200 nanometers (nm), the size of a
small bacterium.
Light microscopes can magnify effectively to about 1,000 times the size of the actual
specimen.
At higher magnifications, the image blurs.
Techniques developed in the 20th century have enhanced contrast and enabled particular
cell components to be stained or labeled so they stand out.
While a light microscope can resolve individual cells, it cannot resolve much of the
internal anatomy, especially the organelles.
To resolve smaller structures, we use an electron microscope (EM), which focuses a
beam of electrons through the specimen or onto its surface.
Because resolution is inversely related to wavelength used, electron microscopes
(whose electron beams have shorter wavelengths than visible light) have finer
resolution.
Theoretically, the resolution of a modern EM could reach 0.002 nanometer (nm), but
the practical limit is closer to about 2 nm.
Transmission electron microscopes (TEMs) are used mainly to study the internal
ultrastructure of cells.
Biology Chapter Notes
A TEM aims an electron beam through a thin section of the specimen.
The image is focused and magnified by electromagnets.
To enhance contrast, the thin sections are stained with atoms of heavy metals.
Scanning electron microscopes (SEMs) are useful for studying surface structures.
The sample surface is covered with a thin film of gold.
The beam excites electrons on the surface of the sample.
These secondary electrons are collected and focused on a screen.
The result is an image of the topography of the specimen.
The SEM has great depth of field, resulting in an image that seems three-dimensional.
Electron microscopes reveal organelles that are impossible to resolve with the light
microscope.
However, electron microscopes can only be used on dead cells.
Light microscopes do not have as high a resolution, but they can be used to study live
cells.
Microscopes are major tools in cytology, the study of cell structures.
Cytology combined with biochemistry, the study of molecules and chemical processes in
metabolism, to produce modern cell biology.
Cell biologists can isolate organelles to study their functions.
The goal of cell fractionation is to separate the major organelles of the cells so their
individual functions can be studied.
This process is driven by an ultracentrifuge, a machine that can spin at up to 130,000
revolutions per minute and apply forces of more than 1 million times gravity (1,000,000
g).
Fractionation begins with homogenization, gently disrupting the cell.
The homogenate is spun in a centrifuge to separate heavier pieces into the pellet while
lighter particles remain in the supernatant.
As the process is repeated at higher speeds and for longer durations, smaller and
smaller organelles can be collected in subsequent pellets.
Cell fractionation prepares isolates of specific cell components.
This enables the functions of these organelles to be determined, especially by the
reactions or processes catalyzed by their proteins.
For example, one cellular fraction was enriched in enzymes that function in cellular
respiration.
Electron microscopy revealed that this fraction is rich in mitochondria.
This evidence helped cell biologists determine that mitochondria are the site of cellular
respiration.
Cytology and biochemistry complement each other in correlating cellular structure and
function.
Concept 6.2 Eukaryotic cells have internal membranes that compartmentalize their
functions
Prokaryotic and eukaryotic cells differ in size and complexity.
Biology Chapter Notes
All cells are surrounded by a plasma membrane.
The semifluid substance within the membrane is the cytosol, containing the organelles.
All cells contain chromosomes that have genes in the form of DNA.
All cells also have ribosomes, tiny organelles that make proteins using the instructions
contained in genes.
A major difference between prokaryotic and eukaryotic cells is the location of
chromosomes.
In a eukaryotic cell, chromosomes are contained in a membrane-enclosed organelle, the
nucleus.
In a prokaryotic cell, the DNA is concentrated in the nucleoid without a membrane
separating it from the rest of the cell.
In eukaryote cells, the chromosomes are contained within a membranous nuclear
envelope.
The region between the nucleus and the plasma membrane is the cytoplasm.
All the material within the plasma membrane of a prokaryotic cell is cytoplasm.
Within the cytoplasm of a eukaryotic cell are a variety of membrane-bound organelles of
specialized form and function.
These membrane-bound organelles are absent in prokaryotes.
Eukaryotic cells are generally much bigger than prokaryotic cells.
The logistics of carrying out metabolism set limits on cell size.
At the lower limit, the smallest bacteria, mycoplasmas, are between 0.1 to 1.0 micron.
Most bacteria are 1–10 microns in diameter.
Eukaryotic cells are typically 10–100 microns in diameter.
Metabolic requirements also set an upper limit to the size of a single cell.
As a cell increases in size, its volume increases faster than its surface area.
Smaller objects have a greater ratio of surface area to volume.
The plasma membrane functions as a selective barrier that allows the passage of oxygen,
nutrients, and wastes for the whole volume of the cell.
The volume of cytoplasm determines the need for this exchange.
Rates of chemical exchange across the plasma membrane may be inadequate to maintain a
cell with a very large cytoplasm.
The need for a surface sufficiently large to accommodate the volume explains the
microscopic size of most cells.
Larger organisms do not generally have larger cells than smaller organisms—simply
more cells.
Cells that exchange a lot of material with their surroundings, such as intestinal cells, may
have long, thin projections from the cell surface called microvilli. Microvilli increase
surface area without significantly increasing cell volume.
Internal membranes compartmentalize the functions of a eukaryotic cell.
A eukaryotic cell has extensive and elaborate internal membranes, which partition the cell
into compartments.
Concept 6.3 The eukaryotic cell’s genetic instructions are housed in the nucleus and
carried out by the ribosomes
The nucleus contains most of the genes in a eukaryotic cell.
Additional genes are located in mitochondria and chloroplasts.
The nucleus averages about 5 microns in diameter.
The nucleus is separated from the cytoplasm by a double membrane called the nuclear
envelope.
The two membranes of the nuclear envelope are separated by 20–40 nm.
The envelope is perforated by pores that are about 100 nm in diameter.
At the lip of each pore, the inner and outer membranes of the nuclear envelope are
fused to form a continuous membrane.
A protein structure called a pore complex lines each pore, regulating the passage of
certain large macromolecules and particles.
The nuclear side of the envelope is lined by the nuclear lamina, a network of protein
filaments that maintains the shape of the nucleus.
There is evidence that a framework of fibers called the nuclear matrix extends through
the nuclear interior.
Within the nucleus, the DNA and associated proteins are organized into discrete units
called chromosomes, structures that carry the genetic information.
Each chromosome is made up of fibrous material called chromatin, a complex of proteins
and DNA.
Stained chromatin appears through light microscopes and electron microscopes as a
diffuse mass.
As the cell prepares to divide, the chromatin fibers coil up and condense, becoming thick
enough to be recognized as the familiar chromosomes.
Each eukaryotic species has a characteristic number of chromosomes.
A typical human cell has 46 chromosomes.
A human sex cell (egg or sperm) has only 23 chromosomes.
In the nucleus is a region of densely stained fibers and granules adjoining chromatin, the
nucleolus.
Biology Chapter Notes
In the nucleolus, ribosomal RNA (rRNA) is synthesized and assembled with proteins
from the cytoplasm to form ribosomal subunits.
The subunits pass through the nuclear pores to the cytoplasm, where they combine to
form ribosomes.
The nucleus directs protein synthesis by synthesizing messenger RNA (mRNA).
The mRNA travels to the cytoplasm through the nuclear pores and combines with
ribosomes to translate its genetic message into the primary structure of a specific
polypeptide.
Ribosomes build a cell’s proteins.
Ribosomes, containing rRNA and protein, are the organelles that carry out protein
synthesis.
Cell types that synthesize large quantities of proteins (e.g., pancreas cells) have large
numbers of ribosomes and prominent nucleoli.
Some ribosomes, free ribosomes, are suspended in the cytosol and synthesize proteins that
function within the cytosol.
Other ribosomes, bound ribosomes, are attached to the outside of the endoplasmic
reticulum or nuclear envelope.
These synthesize proteins that are either included in membranes or exported from the
cell.
Ribosomes can shift between roles depending on the polypeptides they are synthesizing.
Concept 6.4 The endomembrane system regulates protein traffic and performs
metabolic functions in the cell
Many of the internal membranes in a eukaryotic cell are part of the endomembrane
system.
These membranes are either directly continuous or connected via transfer of vesicles, sacs
of membrane.
In spite of these connections, these membranes are diverse in function and structure.
The thickness, molecular composition and types of chemical reactions carried out by
proteins in a given membrane may be modified several times during a membrane’s
life.
The endomembrane system includes the nuclear envelope, endoplasmic reticulum, Golgi
apparatus, lysosomes, vacuoles, and the plasma membrane.
The endoplasmic reticulum manufactures membranes and performs many other
biosynthetic functions.
The endoplasmic reticulum (ER) accounts for half the membranes in a eukaryotic cell.
The ER includes membranous tubules and internal, fluid-filled spaces called cisternae.
The ER membrane is continuous with the nuclear envelope, and the cisternal space of the
ER is continuous with the space between the two membranes of the nuclear envelope.
There are two connected regions of ER that differ in structure and function.
Smooth ER looks smooth because it lacks ribosomes.
Rough ER looks rough because ribosomes (bound ribosomes) are attached to the
outside, including the outside of the nuclear envelope.
Biology Chapter Notes
The smooth ER is rich in enzymes and plays a role in a variety of metabolic processes.
Enzymes of smooth ER synthesize lipids, including oils, phospholipids, and steroids.
These include the sex hormones of vertebrates and adrenal steroids.
In the smooth ER of the liver, enzymes help detoxify poisons and drugs such as
alcohol and barbiturates.
Frequent use of these drugs leads to the proliferation of smooth ER in liver cells,
increasing the rate of detoxification.
This increases tolerance to the target and other drugs, so higher doses are required
to achieve the same effect.
Smooth ER stores calcium ions.
Muscle cells have a specialized smooth ER that pumps calcium ions from the
cytosol and stores them in its cisternal space.
When a nerve impulse stimulates a muscle cell, calcium ions rush from the ER into
the cytosol, triggering contraction.
Enzymes then pump the calcium back, readying the cell for the next stimulation.
Rough ER is especially abundant in cells that secrete proteins.
As a polypeptide is synthesized on a ribosome attached to rough ER, it is threaded into
the cisternal space through a pore formed by a protein complex in the ER membrane.
As it enters the cisternal space, the new protein folds into its native conformation.
Most secretory polypeptides are glycoproteins, proteins to which a carbohydrate is
attached.
Secretory proteins are packaged in transport vesicles that carry them to their next
stage.
Rough ER is also a membrane factory.
Membrane-bound proteins are synthesized directly into the membrane.
Enzymes in the rough ER also synthesize phospholipids from precursors in the cytosol.
As the ER membrane expands, membrane can be transferred as transport vesicles to
other components of the endomembrane system.
The Golgi apparatus is the shipping and receiving center for cell products.
Many transport vesicles from the ER travel to the Golgi apparatus for modification of
their contents.
The Golgi is a center of manufacturing, warehousing, sorting, and shipping.
The Golgi apparatus is especially extensive in cells specialized for secretion.
The Golgi apparatus consists of flattened membranous sacs—cisternae—looking like a
stack of pita bread.
The membrane of each cisterna separates its internal space from the cytosol.
One side of the Golgi, the cis side, is located near the ER. The cis face receives
material by fusing with transport vesicles from the ER.
The other side, the trans side, buds off vesicles that travel to other sites.
During their transit from the cis to the trans side, products from the ER are usually
modified.
The Golgi can also manufacture its own macromolecules, including pectin and other
noncellulose polysaccharides.
The Golgi apparatus is a very dynamic structure.
Biology Chapter Notes
According to the cisternal maturation model, the cisternae of the Golgi progress from
the cis to the trans face, carrying and modifying their protein cargo as they move.
Finally, the Golgi sorts and packages materials into transport vesicles.
Molecular identification tags are added to products to aid in sorting.
Products are tagged with identifiers such as phosphate groups. These act like ZIP
codes on mailing labels to identify the product’s final destination.
Lysosomes are digestive compartments.
A lysosome is a membrane-bound sac of hydrolytic enzymes that an animal cell uses to
digest macromolecules.
Lysosomal enzymes can hydrolyze proteins, fats, polysaccharides, and nucleic acids.
These enzymes work best at pH 5.
Proteins in the lysosomal membrane pump hydrogen ions from the cytosol into the
lumen of the lysosomes.
Rupture of one or a few lysosomes has little impact on a cell because the lysosomal
enzymes are not very active at the neutral pH of the cytosol.
However, massive rupture of many lysosomes can destroy a cell by autodigestion.
Lysosomal enzymes and membrane are synthesized by rough ER and then transferred to
the Golgi apparatus for further modification.
Proteins on the inner surface of the lysosomal membrane are spared by digestion by their
three-dimensional conformations, which protect vulnerable bonds from hydrolysis.
Lysosomes carry out intracellular digestion in a variety of circumstances.
Amoebas eat by engulfing smaller organisms by phagocytosis.
The food vacuole formed by phagocytosis fuses with a lysosome, whose enzymes
digest the food.
As the polymers are digested, monomers pass to the cytosol to become nutrients for the
cell.
Lysosomes can play a role in recycling of the cell’s organelles and macromolecules.
This recycling, or autophagy, renews the cell.
During autophagy, a damaged organelle or region of cytosol becomes surrounded by
membrane.
A lysosome fuses with the resulting vesicle, digesting the macromolecules and
returning the organic monomers to the cytosol for reuse.
The lysosomes play a critical role in the programmed destruction of cells in multicellular
organisms.
This process plays an important role in development.
The hands of human embryos are webbed until lysosomes digest the cells in the tissue
between the fingers.
This important process is called programmed cell death, or apoptosis.
Vacuoles have diverse functions in cell maintenance.
Vesicles and vacuoles (larger versions) are membrane-bound sacs with varied functions.
Food vacuoles are formed by phagocytosis and fuse with lysosomes.
Contractile vacuoles, found in freshwater protists, pump excess water out of the cell
to maintain the appropriate concentration of salts.
A large central vacuole is found in many mature plant cells.
Biology Chapter Notes
The membrane surrounding the central vacuole, the tonoplast, is selective in its
transport of solutes into the central vacuole.
The functions of the central vacuole include stockpiling proteins or inorganic ions,
disposing of metabolic byproducts, holding pigments, and storing defensive
compounds that defend the plant against herbivores.
Because of the large vacuole, the cytosol occupies only a thin layer between the
plasma membrane and the tonoplast. The presence of a large vacuole increases
surface area to volume ratio for the cell.
Concept 6.5 Mitochondria and chloroplasts change energy from one form to another
Mitochondria and chloroplasts are the organelles that convert energy to forms that cells
can use for work.
Mitochondria are the sites of cellular respiration, generating ATP from the catabolism of
sugars, fats, and other fuels in the presence of oxygen.
Chloroplasts, found in plants and algae, are the sites of photosynthesis.
They convert solar energy to chemical energy and synthesize new organic compounds
such as sugars from CO2 and H2O.
Mitochondria and chloroplasts are not part of the endomembrane system.
In contrast to organelles of the endomembrane system, each mitochondrion or
chloroplast has two membranes separating the innermost space from the cytosol.
Their membrane proteins are not made by the ER, but rather by free ribosomes in the
cytosol and by ribosomes within the organelles themselves.
Both organelles have small quantities of DNA that direct the synthesis of the polypeptides
produced by these internal ribosomes.
Mitochondria and chloroplasts grow and reproduce as semiautonomous organelles.
Almost all eukaryotic cells have mitochondria.
There may be one very large mitochondrion or hundreds to thousands of individual
mitochondria.
The number of mitochondria is correlated with aerobic metabolic activity.
A typical mitochondrion is 1–10 microns long.
Mitochondria are quite dynamic: moving, changing shape, and dividing.
Mitochondria have a smooth outer membrane and a convoluted inner membrane with
infoldings called cristae.
The inner membrane divides the mitochondrion into two internal compartments.
The first is the intermembrane space, a narrow region between the inner and outer
membranes.
The inner membrane encloses the mitochondrial matrix, a fluid-filled space with
DNA, ribosomes, and enzymes.
Some of the metabolic steps of cellular respiration are catalyzed by enzymes in the
matrix.
The cristae present a large surface area for the enzymes that synthesize ATP.
The chloroplast is one of several members of a generalized class of plant structures called
plastids.
Amyloplasts are colorless plastids that store starch in roots and tubers.
Chromoplasts store pigments for fruits and flowers.
Biology Chapter Notes
Chloroplasts contain the green pigment chlorophyll as well as enzymes and other
molecules that function in the photosynthetic production of sugar.
Chloroplasts measure about 2 microns × 5 microns and are found in leaves and other
green organs of plants and algae.
The contents of the chloroplast are separated from the cytosol by an envelope consisting
of two membranes separated by a narrow intermembrane space.
Inside the innermost membrane is a fluid-filled space, the stroma, in which float
membranous sacs, the thylakoids.
The stroma contains DNA, ribosomes, and enzymes.
The thylakoids are flattened sacs that play a critical role in converting light to chemical
energy. In some regions, thylakoids are stacked like poker chips into grana.
The membranes of the chloroplast divide the chloroplast into three compartments: the
intermembrane space, the stroma, and the thylakoid space.
Like mitochondria, chloroplasts are dynamic structures.
Their shape is plastic, and they can reproduce themselves by pinching in two.
Mitochondria and chloroplasts are mobile and move around the cell along tracks of the
cytoskeleton.
Peroxisomes generate and degrade H2O2 in performing various metabolic functions.
Peroxisomes contain enzymes that transfer hydrogen from various substrates to oxygen.
An intermediate product of this process is hydrogen peroxide (H2O2), a poison.
The peroxisome contains an enzyme that converts H2O2 to water.
Some peroxisomes break fatty acids down to smaller molecules that are transported to
mitochondria as fuel for cellular respiration.
Peroxisomes in the liver detoxify alcohol and other harmful compounds.
Specialized peroxisomes, glyoxysomes, convert the fatty acids in seeds to sugars,
which the seedling can use as a source of energy and carbon until it is capable of
photosynthesis.
Peroxisomes are bound by a single membrane.
They form not from the endomembrane system, but by incorporation of proteins and
lipids from the cytosol.
They split in two when they reach a certain size.
Concept 6.6 The cytoskeleton is a network of fibers that organizes structures and
activities in the cell
The cytoskeleton is a network of fibers extending throughout the cytoplasm.
The cytoskeleton organizes the structures and activities of the cell.
The cytoskeleton provides support, motility, and regulation.
The cytoskeleton provides mechanical support and maintains cell shape.
The cytoskeleton provides anchorage for many organelles and cytosolic enzymes.
The cytoskeleton is dynamic and can be dismantled in one part and reassembled in
another to change the shape of the cell.
Concept 6.7 Extracellular components and connections between cells help coordinate
cellular activities
Plant cells are encased by cell walls.
The cell wall, found in prokaryotes, fungi, and some protists, has multiple functions.
In plants, the cell wall protects the cell, maintains its shape, and prevents excessive uptake
of water.
It also supports the plant against the force of gravity.
The thickness and chemical composition of cell walls differs from species to species and
among cell types within a plant.
The basic design consists of microfibrils of cellulose embedded in a matrix of proteins
and other polysaccharides. This is the basic design of steel-reinforced concrete or
fiberglass.
A mature cell wall consists of a primary cell wall, a middle lamella with sticky
polysaccharides that holds cells together, and layers of secondary cell wall.
Plant cell walls are perforated by channels between adjacent cells called plasmodesmata.
The extracellular matrix (ECM) of animal cells functions in support, adhesion,
movement, and regulation.
Though lacking cell walls, animal cells do have an elaborate extracellular matrix
(ECM).
The primary constituents of the extracellular matrix are glycoproteins, especially collagen
fibers, embedded in a network of glycoprotein proteoglycans.
In many cells, fibronectins in the ECM connect to integrins, intrinsic membrane proteins
that span the membrane and bind on their cytoplasmic side to proteins attached to
microfilaments of the cytoskeleton.
The interconnections from the ECM to the cytoskeleton via the fibronectin-integrin
link permit the integration of changes inside and outside the cell.
The ECM can regulate cell behavior.
Embryonic cells migrate along specific pathways by matching the orientation of their
microfilaments to the “grain” of fibers in the extracellular matrix.
The extracellular matrix can influence the activity of genes in the nucleus via a
combination of chemical and mechanical signaling pathways.
Biology Chapter Notes
This may coordinate the behavior of all the cells within a tissue.
Intercellular junctions help integrate cells into higher levels of structure and function.
Neighboring cells in tissues, organs, or organ systems often adhere, interact, and
communicate through direct physical contact.
Plant cells are perforated with plasmodesmata, channels allowing cytosol to pass
between cells.
Water and small solutes can pass freely from cell to cell.
In certain circumstances, proteins and RNA can be exchanged.
Animals have 3 main types of intercellular links: tight junctions, desmosomes, and gap
junctions.
In tight junctions, membranes of adjacent cells are fused, forming continuous belts
around cells.
This prevents leakage of extracellular fluid.
Desmosomes (or anchoring junctions) fasten cells together into strong sheets, much like
rivets.
Intermediate filaments of keratin reinforce desmosomes.
Gap junctions (or communicating junctions) provide cytoplasmic channels between
adjacent cells.
Special membrane proteins surround these pores.
Ions, sugars, amino acids, and other small molecules can pass.
In embryos, gap junctions facilitate chemical communication during development.
A cell is a living unit greater than the sum of its parts.
While the cell has many structures with specific functions, all these structures must work
together.
For example, macrophages use actin filaments to move and extend pseudopodia to
capture their bacterial prey.
Food vacuoles are digested by lysosomes, a product of the endomembrane system of
ER and Golgi.
The enzymes of the lysosomes and proteins of the cytoskeleton are synthesized on the
ribosomes.
The information for the proteins comes from genetic messages sent by DNA in the
nucleus.
All of these processes require energy in the form of ATP, most of which is supplied by the
mitochondria.
A cell is a living unit greater than the sum of its parts.
The plasma membrane separates the living cell from its nonliving surroundings.
This thin barrier, 8 nm thick, controls traffic into and out of the cell.
Like all biological membranes, the plasma membrane is selectively permeable, allowing
some substances to cross more easily than others.
Concept 7.1 Cellular membranes are fluid mosaics of lipids and proteins
The main macromolecules in membranes are lipids and proteins, but carbohydrates are
also important.
The most abundant lipids are phospholipids.
Phospholipids and most other membrane constituents are amphipathic molecules.
Amphipathic molecules have both hydrophobic regions and hydrophilic regions.
The arrangement of phospholipids and proteins in biological membranes is described by
the fluid mosaic model.
Membrane models have evolved to fit new data.
Models of membranes were developed long before membranes were first seen with
electron microscopes in the 1950s.
In 1915, membranes isolated from red blood cells were chemically analyzed and found
to be composed of lipids and proteins.
In 1925, E. Gorter and F. Grendel reasoned that cell membranes must be a
phospholipid bilayer two molecules thick.
The molecules in the bilayer are arranged such that the hydrophobic fatty acid tails are
sheltered from water while the hydrophilic phosphate groups interact with water.
Actual membranes adhere more strongly to water than do artificial membranes
composed only of phospholipids.
One suggestion was that proteins on the surface of the membrane increased adhesion.
In 1935, H. Davson and J. Danielli proposed a sandwich model in which the
phospholipid bilayer lies between two layers of globular proteins.
Early images from electron microscopes seemed to support the Davson-Danielli
model, and until the 1960s, it was widely accepted as the structure of the plasma
membrane and internal membranes.
Further investigation revealed two problems.
First, not all membranes were alike. Membranes differ in thickness, appearance
when stained, and percentage of proteins.
Membranes with different functions differ in chemical composition and
structure.
Second, measurements showed that membrane proteins are not very soluble in
water.
Biology Chapter Notes
Membrane proteins are amphipathic, with hydrophobic and hydrophilic regions.
If membrane proteins were at the membrane surface, their hydrophobic regions
would be in contact with water.
In 1972, S. J. Singer and G. Nicolson presented a revised model that proposed that the
membrane proteins are dispersed and individually inserted into the phospholipid bilayer.
In this fluid mosaic model, the hydrophilic regions of proteins and phospholipids are in
maximum contact with water, and the hydrophobic regions are in a nonaqueous
environment within the membrane.
A specialized preparation technique, freeze-fracture, splits a membrane along the middle
of the phospholipid bilayer.
When a freeze-fracture preparation is viewed with an electron microscope, protein
particles are interspersed in a smooth matrix, supporting the fluid mosaic model.
Membranes are fluid.
Membrane molecules are held in place by relatively weak hydrophobic interactions.
Most of the lipids and some proteins drift laterally in the plane of the membrane, but
rarely flip-flop from one phospholipid layer to the other.
The lateral movements of phospholipids are rapid, about 2 microns per second. A
phospholipid can travel the length of a typical bacterial cell in 1 second.
Many larger membrane proteins drift within the phospholipid bilayer, although they move
more slowly than the phospholipids.
Some proteins move in a very directed manner, perhaps guided or driven by motor
proteins attached to the cytoskeleton.
Other proteins never move and are anchored to the cytoskeleton.
Membrane fluidity is influenced by temperature. As temperatures cool, membranes switch
from a fluid state to a solid state as the phospholipids pack more closely.
Membrane fluidity is also influenced by its components. Membranes rich in unsaturated
fatty acids are more fluid that those dominated by saturated fatty acids because the kinks
in the unsaturated fatty acid tails at the locations of the double bonds prevent tight
packing.
The steroid cholesterol is wedged between phospholipid molecules in the plasma
membrane of animal cells.
At warm temperatures (such as 37°C), cholesterol restrains the movement of
phospholipids and reduces fluidity.
At cool temperatures, it maintains fluidity by preventing tight packing.
Thus, cholesterol acts as a “temperature buffer” for the membrane, resisting changes in
membrane fluidity as temperature changes.
To work properly with active enzymes and appropriate permeability, membranes must be
about as fluid as salad oil.
Cells can alter the lipid composition of membranes to compensate for changes in fluidity
caused by changing temperatures.
For example, cold-adapted organisms such as winter wheat increase the percentage of
unsaturated phospholipids in their membranes in the autumn.
This prevents membranes from solidifying during winter.
Membranes are mosaics of structure and function.
Biology Chapter Notes
A membrane is a collage of different proteins embedded in the fluid matrix of the lipid
bilayer.
Proteins determine most of the membrane’s specific functions.
The plasma membrane and the membranes of the various organelles each have unique
collections of proteins.
There are two major populations of membrane proteins.
Peripheral proteins are not embedded in the lipid bilayer at all.
Instead, they are loosely bound to the surface of the protein, often connected to
integral proteins.
Integral proteins penetrate the hydrophobic core of the lipid bilayer, often completely
spanning the membrane (as transmembrane proteins).
The hydrophobic regions embedded in the membrane’s core consist of stretches of
nonpolar amino acids, often coiled into alpha helices.
Where integral proteins are in contact with the aqueous environment, they have
hydrophilic regions of amino acids.
On the cytoplasmic side of the membrane, some membrane proteins connect to the
cytoskeleton.
On the exterior side of the membrane, some membrane proteins attach to the fibers of
the extracellular matrix.
The proteins of the plasma membrane have six major functions:
4. Transport of specific solutes into or out of cells.
5. Enzymatic activity, sometimes catalyzing one of a number of steps of a metabolic
pathway.
6. Signal transduction, relaying hormonal messages to the cell.
7. Cell-cell recognition, allowing other proteins to attach two adjacent cells together.
8. Intercellular joining of adjacent cells with gap or tight junctions.
9. Attachment to the cytoskeleton and extracellular matrix, maintaining cell shape
and stabilizing the location of certain membrane proteins.
Membrane carbohydrates are important for cell-cell recognition.
The plasma membrane plays the key role in cell-cell recognition.
Cell-cell recognition, the ability of a cell to distinguish one type of neighboring cell
from another, is crucial to the functioning of an organism.
This attribute is important in the sorting and organization of cells into tissues and
organs during development.
It is also the basis for rejection of foreign cells by the immune system.
Cells recognize other cells by binding to surface molecules, often carbohydrates, on
the plasma membrane.
Membrane carbohydrates are usually branched oligosaccharides with fewer than 15 sugar
units.
They may be covalently bonded to lipids, forming glycolipids, or more commonly to
proteins, forming glycoproteins.
The oligosaccharides on the external side of the plasma membrane vary from species to
species, from individual to individual, and even from cell type to cell type within the same
individual.
This variation distinguishes each cell type.
Biology Chapter Notes
The four human blood groups (A, B, AB, and O) differ in the external carbohydrates
on red blood cells.
Membranes have distinctive inside and outside faces.
Membranes have distinct inside and outside faces. The two layers may differ in lipid
composition. Each protein in the membrane has a directional orientation in the membrane.
The asymmetrical orientation of proteins, lipids and associated carbohydrates begins
during the synthesis of membrane in the ER and Golgi apparatus.
Membrane lipids and proteins are synthesized in the endoplasmic reticulum.
Carbohydrates are added to proteins in the ER, and the resulting glycoproteins are further
modified in the Golgi apparatus. Glycolipids are also produced in the Golgi apparatus.
When a vesicle fuses with the plasma membrane, the outside layer of the vesicle becomes
continuous with the inside layer of the plasma membrane. In that way, molecules that
originate on the inside face of the ER end up on the outside face of the plasma membrane.
Concept 7.4 Active transport uses energy to move solutes against their gradients
Some transport proteins can move solutes across membranes against their concentration
gradient, from the side where they are less concentrated to the side where they are more
concentrated.
This active transport requires the cell to expend metabolic energy.
Active transport enables a cell to maintain its internal concentrations of small molecules
that would otherwise diffuse across the membrane.
Active transport is performed by specific proteins embedded in the membranes.
ATP supplies the energy for most active transport.
Concept 7.5 Bulk transport across the plasma membrane occurs by exocytosis and
endocytosis
Small molecules and water enter or leave the cell through the lipid bilayer or by transport
proteins.
Large molecules, such as polysaccharides and proteins, cross the membrane via vesicles.
During exocytosis, a transport vesicle budded from the Golgi apparatus is moved by the
cytoskeleton to the plasma membrane.
When the two membranes come in contact, the bilayers fuse and spill the contents to the
outside.
Many secretory cells use exocytosis to export their products.
During endocytosis, a cell brings in macromolecules and particulate matter by forming
new vesicles from the plasma membrane.
Endocytosis is a reversal of exocytosis, although different proteins are involved in the two
processes.
A small area of the plasma membrane sinks inward to form a pocket.
As the pocket deepens, it pinches in to form a vesicle containing the material that had
been outside the cell.
There are three types of endocytosis: phagocytosis (“cellular eating”), pinocytosis
(“cellular drinking”), and receptor-mediated endocytosis.
In phagocytosis, the cell engulfs a particle by extending pseudopodia around it and
packaging it in a large vacuole.
The contents of the vacuole are digested when the vacuole fuses with a lysosome.
In pinocytosis, a cell creates a vesicle around a droplet of extracellular fluid. All included
solutes are taken into the cell in this nonspecific process.
Receptor-mediated endocytosis allows greater specificity, transporting only certain
substances.
This process is triggered when extracellular substances, or ligands, bind to special
receptors on the membrane surface. The receptor proteins are clustered in regions of the
Concept 8.1 An organism’s metabolism transforms matter and energy, subject to the
laws of thermodynamics
The totality of an organism’s chemical reactions is called metabolism.
Metabolism is an emergent property of life that arises from interactions between
molecules within the orderly environment of the cell.
The chemistry of life is organized into metabolic pathways.
Metabolic pathways begin with a specific molecule, which is then altered in a series of
defined steps to form a specific product.
A specific enzyme catalyzes each step of the pathway.
Catabolic pathways release energy by breaking down complex molecules to simpler
compounds.
A major pathway of catabolism is cellular respiration, in which the sugar glucose is
broken down in the presence of oxygen to carbon dioxide and water.
Anabolic pathways consume energy to build complicated molecules from simpler
compounds. They are also called biosynthetic pathways.
The synthesis of protein from amino acids is an example of anabolism.
The energy released by catabolic pathways can be stored and then used to drive anabolic
pathways.
Energy is fundamental to all metabolic processes, and therefore an understanding of
energy is key to understanding how the living cell works.
Bioenergetics is the study of how organisms manage their energy resources.
Organisms transform energy.
Energy is the capacity to do work.
Energy exists in various forms, and cells transform energy from one type into another.
Kinetic energy is the energy associated with the relative motion of objects.
Objects in motion can perform work by imparting motion to other matter.
Photons of light can be captured and their energy harnessed to power photosynthesis in
green plants.
Heat or thermal energy is kinetic energy associated with the random movement of
atoms or molecules.
Potential energy is the energy that matter possesses because of its location or structure.
Chemical energy is a form of potential energy stored in molecules because of the
arrangement of their atoms.
Energy can be converted from one form to another.
For example, as a boy climbs stairs to a diving platform, he is releasing chemical
energy stored in his cells from the food he ate for lunch.
Biology Chapter Notes
The kinetic energy of his muscle movement is converted into potential energy as he
climbs higher.
As he dives, the potential energy is converted back to kinetic energy.
Kinetic energy is transferred to the water as he enters it.
Some energy is converted to heat due to friction.
The energy transformations of life are subject to two laws of thermodynamics.
Thermodynamics is the study of energy transformations.
In this field, the term system refers to the matter under study and the surroundings include
everything outside the system.
A closed system, approximated by liquid in a thermos, is isolated from its surroundings.
In an open system, energy and matter can be transferred between the system and its
surroundings.
Organisms are open systems.
They absorb energy—light or chemical energy in the form of organic molecules—and
release heat and metabolic waste products such as urea or CO2 to their surroundings.
The first law of thermodynamics states that energy can be transferred and transformed,
but it cannot be created or destroyed.
The first law is also known as the principle of conservation of energy.
Plants do not produce energy; they transform light energy to chemical energy.
During every transfer or transformation of energy, some energy is converted to heat,
which is the energy associated with the random movement of atoms and molecules.
A system can use heat to do work only when there is a temperature difference that results
in heat flowing from a warmer location to a cooler one.
If temperature is uniform, as in a living cell, heat can only be used to warm the
organism.
Energy transfers and transformations make the universe more disordered due to this loss
of usable energy.
Entropy is a quantity used as a measure of disorder or randomness.
The more random a collection of matter, the greater its entropy.
The second law of thermodynamics states that every energy transfer or transformation
increases the entropy of the universe.
While order can increase locally, there is an unstoppable trend toward randomization
of the universe.
Much of the increased entropy of the universe takes the form of increasing heat, which
is the energy of random molecular motion.
In most energy transformations, ordered forms of energy are converted at least partly to
heat.
Automobiles convert only 25% of the energy in gasoline into motion; the rest is lost as
heat.
Living cells unavoidably convert organized forms of energy to heat.
For a process to occur on its own, without outside help in the form of energy input, it must
increase the entropy of the universe.
The word spontaneous describes a process that can occur without an input of energy.
Spontaneous processes need not occur quickly.
Biology Chapter Notes
Some spontaneous processes are instantaneous, such as an explosion. Some are very
slow, such as the rusting of an old car.
Another way to state the second law of thermodynamics is for a process to occur
spontaneously, it must increase the entropy of the universe.
Living systems create ordered structures from less ordered starting materials.
For example, amino acids are ordered into polypeptide chains.
The structure of a multicellular body is organized and complex.
However, an organism also takes in organized forms of matter and energy from its
surroundings and replaces them with less ordered forms.
For example, an animal consumes organic molecules as food and catabolizes them to
low-energy carbon dioxide and water.
Over evolutionary time, complex organisms have evolved from simpler ones.
This increase in organization does not violate the second law of thermodynamics.
The entropy of a particular system, such as an organism, may decrease as long as the
total entropy of the universe—the system plus its surroundings—increases.
Organisms are islands of low entropy in an increasingly random universe.
The evolution of biological order is perfectly consistent with the laws of
thermodynamics.
Concept 8.2 The free-energy change of a reaction tells us whether the reaction occurs
spontaneously
How can we determine which reactions occur spontaneously and which ones require an
input of energy?
The concept of free energy provides a useful function for measuring spontaneity of a
system.
Free energy is the portion of a system’s energy that is able to perform work when
temperature and pressure is uniform throughout the system, as in a living cell.
The free energy (G) in a system is related to the total enthalpy (in biological systems,
equivalent to energy) (H) and the entropy (S) by this relationship:
G = H − TS, where T is temperature in Kelvin units.
Increases in temperature amplify the entropy term.
Not all the energy in a system is available for work because the entropy component
must be subtracted from the enthalpy component.
What remains is the free energy that is available for work.
Free energy can be thought of as a measure of the stability of a system.
Systems that are high in free energy—compressed springs, separated charges, organic
polymers—are unstable and tend to move toward a more stable state, one with less free
energy.
Systems that tend to change spontaneously are those that have high enthalpy, low
entropy, or both.
In any spontaneous process, the free energy of a system decreases.
We can represent this change in free energy from the start of a process until its finish by:
G = Gfinal state − Gstarting state
Or G = H − TS
Biology Chapter Notes
For a process to be spontaneous, the system must either give up enthalpy (decrease in H),
give up order (increase in S), or both.
G must be negative for a process to be spontaneous.
Every spontaneous process is characterized by a decrease in the free energy of the
system.
Processes that have a positive or zero G are never spontaneous.
The greater the decrease in free energy, the more work a spontaneous process can
perform.
Nature runs “downhill.”
A system at equilibrium is at maximum stability.
In a chemical reaction at equilibrium, the rates of forward and backward reactions are
equal, and there is no change in the concentration of products or reactants.
At equilibrium G = 0, and the system can do no work.
A process is spontaneous and can perform work only when it is moving toward
equilibrium.
Movements away from equilibrium are nonspontaneous and require the addition of
energy from an outside energy source (the surroundings).
Chemical reactions can be classified as either exergonic or endergonic based on free
energy.
An exergonic reaction proceeds with a net release of free energy; G is negative.
The magnitude of G for an exergonic reaction is the maximum amount of work the
reaction can perform.
The greater the decrease in free energy, the greater the amount of work that can be done.
For the overall reaction of cellular respiration: C6H12O6 + 6O2 -> 6CO2 + 6H2O
G = −686 kcal/mol
For each mole (180 g) of glucose broken down by respiration, 686 kcal of energy are
made available to do work in the cell.
The products have 686 kcal less free energy than the reactants.
An endergonic reaction is one that absorbs free energy from its surroundings.
Endergonic reactions store energy in molecules; G is positive.
Endergonic reactions are nonspontaneous, and the magnitude of G is the quantity of
energy required to drive the reaction.
If cellular respiration releases 686 kcal, then photosynthesis, the reverse reaction, must
require an equivalent investment of energy.
For the conversion of carbon dioxide and water to sugar, G = +686 kcal/mol.
Photosynthesis is strongly endergonic, powered by the absorption of light energy.
Reactions in a closed system eventually reach equilibrium and can do no work.
A cell that has reached metabolic equilibrium has a G = 0 and is dead!
Metabolic disequilibrium is one of the defining features of life.
Cells maintain disequilibrium because they are open systems. The constant flow of
materials into and out of the cell keeps metabolic pathways from ever reaching
equilibrium.
A cell continues to do work throughout its life.
Concept 8.3 ATP powers cellular work by coupling exergonic reactions to endergonic
reactions
A cell does three main kinds of work:
10.Mechanical work, such as the beating of cilia, contraction of muscle cells, and
movement of chromosomes during cellular reproduction.
11.Transport work, the pumping of substances across membranes against the direction of
spontaneous movement.
12.Chemical work, driving endergonic reactions such as the synthesis of polymers from
monomers.
Cells manage their energy resources to do this work by energy coupling, the use of an
exergonic process to drive an endergonic one.
In most cases, the immediate source of energy to power cellular work is ATP.
ATP (adenosine triphosphate) is a type of nucleotide consisting of the nitrogenous base
adenine, the sugar ribose, and a chain of three phosphate groups.
The bonds between phosphate groups can be broken by hydrolysis.
Hydrolysis of the end phosphate group forms adenosine diphosphate.
ATP -> ADP + Pi
This reaction releases 7.3 kcal of energy per mole of ATP under standard
conditions (1 M of each reactant and product, 25°C, pH 7).
In the cell, G for hydrolysis of ATP is about −13 kcal/mol.
While the phosphate bonds of ATP are sometimes referred to as high-energy phosphate
bonds, these are actually fairly weak covalent bonds.
However, they are unstable, and their hydrolysis yields energy because the products
are more stable.
The release of energy during the hydrolysis of ATP comes from the chemical change to a
state of lower free energy, not from the phosphate bonds themselves.
Why does the hydrolysis of ATP yield so much energy?
Each of the three phosphate groups has a negative charge.
These three like charges are crowded together, and their mutual repulsion contributes
to the instability of this region of the ATP molecule.
In the cell, the energy from the hydrolysis of ATP is directly coupled to endergonic
processes by the transfer of the phosphate group to another molecule.
This recipient molecule is now phosphorylated.
This molecule is now more reactive (less stable) than the original unphosphorylated
molecules.
To perform their many tasks, living cells require energy from outside sources.
Energy enters most ecosystems as sunlight and leaves as heat.
Photosynthesis generates oxygen and organic molecules that the mitochondria of
eukaryotes use as fuel for cellular respiration.
Cells harvest the chemical energy stored in organic molecules and use it to regenerate
ATP, the molecule that drives most cellular work.
Respiration has three key pathways: glycolysis, the citric acid cycle, and oxidative
phosphorylation.
Concept 9.3 The citric acid cycle completes the energy-yielding oxidation of organic
molecules
More than three-quarters of the original energy in glucose is still present in the two
molecules of pyruvate.
If oxygen is present, pyruvate enters the mitochondrion where enzymes of the citric acid
cycle complete the oxidation of the organic fuel to carbon dioxide.
Concept 9.5 Fermentation enables some cells to produce ATP without the use of oxygen
Without electronegative oxygen to pull electrons down the transport chain, oxidative
phosphorylation ceases.
Concept 9.6 Glycolysis and the citric acid cycle connect to many other metabolic
pathways
Glycolysis can accept a wide range of carbohydrates for catabolism.
Polysaccharides like starch or glycogen can be hydrolyzed to glucose monomers that
enter glycolysis.
Other hexose sugars, such as galactose and fructose, can also be modified to undergo
glycolysis.
The other two major fuels, proteins and fats, can also enter the respiratory pathways used
by carbohydrates.
Proteins must first be digested to individual amino acids.
Amino acids that will be catabolized must have their amino groups removed via
deamination.
The nitrogenous waste is excreted as ammonia, urea, or another waste product.
The carbon skeletons are modified by enzymes and enter as intermediaries into glycolysis
or the citric acid cycle, depending on their structure.
Catabolism can also harvest energy stored in fats.
Fats must be digested to glycerol and fatty acids.
Glycerol can be converted to glyceraldehyde phosphate, an intermediate of glycolysis.
The rich energy of fatty acids is accessed as fatty acids are split into two-carbon
fragments via beta oxidation.
These molecules enter the citric acid cycle as acetyl CoA.
A gram of fat oxides by respiration generates twice as much ATP as a gram of
carbohydrate.
The metabolic pathways of respiration also play a role in anabolic pathways of the cell.
Intermediaries in glycolysis and the citric acid cycle can be diverted to anabolic pathways.
For example, a human cell can synthesize about half the 20 different amino acids by
modifying compounds from the citric acid cycle.
Glucose can be synthesized from pyruvate; fatty acids can be synthesized from acetyl
CoA.
Biology Chapter Notes
Glycolysis and the citric acid cycle function as metabolic interchanges that enable cells to
convert one kind of molecule to another as needed.
For example, excess carbohydrates and proteins can be converted to fats through
intermediaries of glycolysis and the citric acid cycle.
Metabolism is remarkably versatile and adaptable.
Feedback mechanisms control cellular respiration.
Basic principles of supply and demand regulate the metabolic economy.
If a cell has an excess of a certain amino acid, it typically uses feedback inhibition to
prevent the diversion of intermediary molecules from the citric acid cycle to the
synthesis pathway of that amino acid.
The rate of catabolism is also regulated, typically by the level of ATP in the cell.
If ATP levels drop, catabolism speeds up to produce more ATP.
Control of catabolism is based mainly on regulating the activity of enzymes at strategic
points in the catabolic pathway.
One strategic point occurs in the third step of glycolysis, catalyzed by
phosphofructokinase.
Allosteric regulation of phosphofructokinase sets the pace of respiration.
This enzyme catalyzes the earliest step that irreversibly commits the substrate to
glycolysis.
Phosphofructokinase is an allosteric enzyme with receptor sites for specific inhibitors
and activators.
It is inhibited by ATP and stimulated by AMP (derived from ADP).
When ATP levels are high, inhibition of this enzyme slows glycolysis.
As ATP levels drop and ADP and AMP levels rise, the enzyme becomes active
again and glycolysis speeds up.
Citrate, the first product of the citric acid cycle, is also an inhibitor of
phosphofructokinase.
This synchronizes the rate of glycolysis and the citric acid cycle.
If intermediaries from the citric acid cycle are diverted to other uses (e.g., amino acid
synthesis), glycolysis speeds up to replace these molecules.
Metabolic balance is augmented by the control of other enzymes at other key locations in
glycolysis and the citric acid cycle.
Cells are thrifty, expedient, and responsive in their metabolism.
Concept 10.1 Photosynthesis converts light energy to the chemical energy of food
All green parts of a plant have chloroplasts.
However, the leaves are the major site of photosynthesis for most plants.
There are about half a million chloroplasts per square millimeter of leaf surface.
The color of a leaf comes from chlorophyll, the green pigment in the chloroplasts.
Chlorophyll plays an important role in the absorption of light energy during
photosynthesis.
Biology Chapter Notes
Chloroplasts are found mainly in mesophyll cells forming the tissues in the interior of the
leaf.
O2 exits and CO2 enters the leaf through microscopic pores called stomata in the leaf.
Veins deliver water from the roots and carry off sugar from mesophyll cells to
nonphotosynthetic areas of the plant.
A typical mesophyll cell has 30–40 chloroplasts, each about 2–4 microns by 4–7 microns
long.
Each chloroplast has two membranes around a central aqueous space, the stroma.
In the stroma is an elaborate system of interconnected membranous sacs, the thylakoids.
The interior of the thylakoids forms another compartment, the thylakoid space.
Thylakoids may be stacked into columns called grana.
Chlorophyll is located in the thylakoids.
Photosynthetic prokaryotes lack chloroplasts.
Their photosynthetic membranes arise from infolded regions of the plasma
membranes, folded in a manner similar to the thylakoid membranes of chloroplasts.
Evidence that chloroplasts split water molecules enabled researchers to track atoms
through photosynthesis.
Powered by light, the green parts of plants produce organic compounds and O 2 from CO2
and H2O.
The equation describing the process of photosynthesis is:
6CO2 + 12H2O + light energy C6H12O6 + 6O2+ 6H2O
C6H12O6 is glucose.
Water appears on both sides of the equation because 12 molecules of water are consumed,
and 6 molecules are newly formed during photosynthesis.
We can simplify the equation by showing only the net consumption of water:
6CO2 + 6H2O + light energy C6H12O6 + 6O2
The overall chemical change during photosynthesis is the reverse of cellular respiration.
In its simplest possible form: CO2 + H2O + light energy [CH2O] + O2
[CH2O] represents the general formula for a sugar.
One of the first clues to the mechanism of photosynthesis came from the discovery that
the O2 given off by plants comes from H2O, not CO2.
Before the 1930s, the prevailing hypothesis was that photosynthesis split carbon
dioxide and then added water to the carbon:
Step 1: CO2 C + O2
Step 2: C + H2O CH2O
C. B. van Niel challenged this hypothesis.
In the bacteria that he was studying, hydrogen sulfide (H 2S), not water, is used in
photosynthesis.
These bacteria produce yellow globules of sulfur as a waste, rather than oxygen.
Van Niel proposed this chemical equation for photosynthesis in sulfur bacteria:
CO2 + 2H2S [CH2O] + H2O + 2S
He generalized this idea and applied it to plants, proposing this reaction for their
photosynthesis:
Biology Chapter Notes
CO2 + 2H2O [CH2O] + H2O + O2
Thus, van Niel hypothesized that plants split water as a source of electrons from hydrogen
atoms, releasing oxygen as a byproduct.
Other scientists confirmed van Niel’s hypothesis twenty years later.
They used 18O, a heavy isotope, as a tracer.
They could label either C18O2 or H218O.
They found that the 18O label only appeared in the oxygen produced in photosynthesis
when water was the source of the tracer.
Hydrogen extracted from water is incorporated into sugar, and oxygen is released to the
atmosphere (where it can be used in respiration).
Photosynthesis is a redox reaction.
It reverses the direction of electron flow in respiration.
Water is split and electrons transferred with H+ from water to CO2, reducing it to sugar.
Because the electrons increase in potential energy as they move from water to sugar,
the process requires energy.
The energy boost is provided by light.
Here is a preview of the two stages of photosynthesis.
Photosynthesis is two processes, each with multiple stages.
The light reactions (photo) convert solar energy to chemical energy.
The Calvin cycle (synthesis) uses energy from the light reactions to incorporate CO2 from
the atmosphere into sugar.
In the light reactions, light energy absorbed by chlorophyll in +the thylakoids drives the
transfer of electrons and hydrogen from water to NADP (nicotinamide adenine
dinucleotide phosphate), forming NADPH.
NADPH, an electron acceptor, provides reducing power via energized electrons to the
Calvin cycle.
Water is split in the process, and O2 is released as a by-product.
The light reaction also generates ATP using chemiosmosis, in a process called
photophosphorylation.
Thus light energy is initially converted to chemical energy in the form of two compounds:
NADPH and ATP.
The Calvin cycle is named for Melvin Calvin who, with his colleagues, worked out many
of its steps in the 1940s.
The cycle begins with the incorporation of CO2 into organic molecules, a process known
as carbon fixation.
The fixed carbon is reduced with electrons provided by NADPH.
ATP from the light reactions also powers parts of the Calvin cycle.
Thus, it is the Calvin cycle that makes sugar, but only with the help of ATP and NADPH
from the light reactions.
The metabolic steps of the Calvin cycle are sometimes referred to as the light-independent
reactions, because none of the steps requires light directly.
Nevertheless, the Calvin cycle in most plants occurs during daylight, because that is when
the light reactions can provide the NADPH and ATP the Calvin cycle requires.
Biology Chapter Notes
While the light reactions occur at the thylakoids, the Calvin cycle occurs in the stroma.
Concept 10.2 The light reactions convert solar energy to the chemical energy of ATP
and NADPH
The thylakoids convert light energy into the chemical energy of ATP and NADPH.
Light is a form of electromagnetic radiation.
Like other forms of electromagnetic energy, light travels in rhythmic waves.
The distance between crests of electromagnetic waves is called the wavelength.
Wavelengths of electromagnetic radiation range from less than a nanometer (gamma
rays) to more than a kilometer (radio waves).
The entire range of electromagnetic radiation is the electromagnetic spectrum.
The most important segment for life is a narrow band between 380 to 750 nm, the band of
visible light.
While light travels as a wave, many of its properties are those of a discrete particle, the
photon.
Photons are not tangible objects, but they do have fixed quantities of energy.
The amount of energy packaged in a photon is inversely related to its wavelength.
Photons with shorter wavelengths pack more energy.
While the sun radiates a full electromagnetic spectrum, the atmosphere selectively screens
out most wavelengths, permitting only visible light to pass in significant quantities.
Visible light is the radiation that drives photosynthesis.
When light meets matter, it may be reflected, transmitted, or absorbed.
Different pigments absorb photons of different wavelengths, and the wavelengths that
are absorbed disappear.
A leaf looks green because chlorophyll, the dominant pigment, absorbs red and blue
light, while transmitting and reflecting green light.
A spectrophotometer measures the ability of a pigment to absorb various wavelengths of
light.
It beams narrow wavelengths of light through a solution containing the pigment and
measures the fraction of light transmitted at each wavelength.
An absorption spectrum plots a pigment’s light absorption versus wavelength.
The light reaction can perform work with those wavelengths of light that are absorbed.
There are several pigments in the thylakoid that differ in their absorption spectra.
Chlorophyll a, the dominant pigment, absorbs best in the red and violet-blue
wavelengths and least in the green.
Other pigments with different structures have different absorption spectra.
Collectively, these photosynthetic pigments determine an overall action spectrum for
photosynthesis.
An action spectrum measures changes in some measure of photosynthetic activity (for
example, O2 release) as the wavelength is varied.
The action spectrum of photosynthesis was first demonstrated in 1883 in an elegant
experiment performed by Thomas Engelmann.
Concept 10.3 The Calvin cycle uses ATP and NADPH to convert CO2 to sugar
The Calvin cycle regenerates its starting material after molecules enter and leave the
cycle.
The Calvin cycle is anabolic, using energy to build sugar from smaller molecules.
Carbon enters the cycle as CO2 and leaves as sugar.
The cycle spends the energy of ATP and the reducing power of electrons carried by
NADPH to make sugar.
Concept 10.4 Alternative mechanisms of carbon fixation have evolved in hot, arid
climates
One of the major problems facing terrestrial plants is dehydration.
At times, solutions to this problem require tradeoffs with other metabolic processes,
especially photosynthesis.
Biology Chapter Notes
The stomata are not only the major route for gas exchange (CO 2 in and O2 out), but also
for the evaporative loss of water.
On hot, dry days, plants close their stomata to conserve water. This causes problems for
photosynthesis.
In most plants (C3 plants), initial fixation of CO2 occurs via rubisco, forming a three-
carbon compound, 3-phosphoglycerate.
C3 plants include rice, wheat, and soybeans.
When their stomata partially close on a hot, dry day, CO 2 levels drop as CO2 is consumed
in the Calvin cycle.
At the same time, O2 levels rise as the light reaction converts light to chemical energy.
While rubisco normally accepts CO2, when the O2:CO2 ratio increases (on a hot, dry day
with closed stomata), rubisco can add O2 to RuBP.
When rubisco adds O2 to RuBP, RuBP splits into a three-carbon piece and a two-carbon
piece in a process called photorespiration.
The two-carbon fragment is exported from the chloroplast and degraded to CO 2 by
mitochondria and peroxisomes.
Unlike normal respiration, this process produces no ATP.
In fact, photorespiration consumes ATP.
Unlike photosynthesis, photorespiration does not produce organic molecules.
In fact, photorespiration decreases photosynthetic output by siphoning organic
material from the Calvin cycle.
A hypothesis for the existence of photorespiration is that it is evolutionary baggage.
When rubisco first evolved, the atmosphere had far less O 2 and more CO2 than it does
today.
The inability of the active site of rubisco to exclude O 2 would have made little
difference.
Today it does make a difference.
Photorespiration can drain away as much as 50% of the carbon fixed by the Calvin
cycle on a hot, dry day.
Certain plant species have evolved alternate modes of carbon fixation to minimize
photorespiration.
C4 plants first fix CO2 in a four-carbon compound.
Several thousand plants, including sugarcane and corn, use this pathway.
A unique leaf anatomy is correlated with the mechanism of C4 photosynthesis.
In C4 plants, there are two distinct types of photosynthetic cells: bundle-sheath cells and
mesophyll cells.
Bundle-sheath cells are arranged into tightly packed sheaths around the veins of the
leaf.
Mesophyll cells are more loosely arranged cells located between the bundle sheath and
the leaf surface.
The Calvin cycle is confined to the chloroplasts of the bundle-sheath cells.
However, the cycle is preceded by the incorporation of CO2 into organic molecules in the
mesophyll.
Concept 11.1 External signals are converted into responses within the cell
What messages are passed from cell to cell? How do cells respond to these messages?
We will first consider communication in microbes, to gain insight into the evolution of
cell signaling.
Cell signaling evolved early in the history of life.
One topic of cell “conversation” is sex.
Saccharomyces cerevisiae, the yeast of bread, wine, and beer, identifies potential mates
by chemical signaling.
There are two sexes, a and , each of which secretes a specific signaling molecule, a
factor and factor, respectively.
These factors each bind to receptor proteins on the other mating type.
Once the mating factors have bound to the receptors, the two cells grow toward each other
and undergo other cellular changes.
The two cells fuse, or mate, to form an a/ cell containing the genes of both cells.
The process by which a signal on a cell’s surface is converted into a specific cellular
response is a series of steps called a signal-transduction pathway.
The molecular details of these pathways are strikingly similar in yeast and animal
cells, even though their last common ancestor lived more than a billion years ago.
Signaling systems of bacteria and plants also share similarities.
These similarities suggest that ancestral signaling molecules evolved long ago in
prokaryotes and have since been adopted for new uses by single-celled eukaryotes and
multicellular descendents.
Communicating cells may be close together or far apart.
Multicellular organisms release signaling molecules that target other cells.
Cells may communicate by direct contact.
Biology Chapter Notes
Both animals and plants have cell junctions that connect to the cytoplasm of adjacent
cells.
Signaling substances dissolved in the cytosol can pass freely between adjacent cells.
Animal cells can communicate by direct contact between membrane-bound cell
surface molecules.
Such cell-cell recognition is important to such processes as embryonic development
and the immune response.
In other cases, messenger molecules are secreted by the signaling cell.
Some transmitting cells release local regulators that influence cells in the local
vicinity.
One class of local regulators in animals, growth factors, includes compounds that
stimulate nearby target cells to grow and multiply.
This is an example of paracrine signaling, which occurs when numerous cells
simultaneously receive and respond to growth factors produced by a single cell in their
vicinity.
In synaptic signaling, a nerve cell produces a neurotransmitter that diffuses across a
synapse to a single cell that is almost touching the sender.
The neurotransmitter stimulates the target cell.
The transmission of a signal through the nervous system can also be considered an
example of long-distance signaling.
Local signaling in plants is not well understood. Because of their cell walls, plants must
have different mechanisms from animals.
Plants and animals use hormones for long-distance signaling.
In animals, specialized endocrine cells release hormones into the circulatory system,
by which they travel to target cells in other parts of the body.
Plant hormones, called growth regulators, may travel in vessels but more often travel
from cell to cell or move through air by diffusion.
Hormones and local regulators range widely in size and type.
The plant hormone ethylene (C2H4), which promotes fruit ripening and regulates
growth, is a hydrocarbon of only six atoms, capable of passing through cell walls.
Insulin, which regulates blood sugar levels in mammals, is a protein with thousands of
atoms.
What happens when a cell encounters a signal?
The signal must be recognized by a specific receptor molecule, and the information it
carries must be changed into another form, or transduced, inside the cell before the
cell can respond.
The three stages of cell signaling are reception, transduction, and response.
E. W. Sutherland and his colleagues pioneered our understanding of cell signaling.
Their work investigated how the animal hormone epinephrine stimulates breakdown of
the storage polysaccharide glycogen in liver and skeletal muscle.
Breakdown of glycogen releases glucose derivatives that can be used for fuel in
glycolysis or released as glucose in the blood for fuel elsewhere.
Thus one effect of epinephrine, which is released from the adrenal gland during times
of physical or mental stress, is mobilization of fuel reserves.
Sutherland’s research team discovered that epinephrine activated a cytosolic enzyme,
glycogen phosphorylase.
Biology Chapter Notes
However, epinephrine did not activate the phosphorylase directly in vitro but could
only act via intact cells.
Therefore, there must be an intermediate step or steps occurring inside the cell.
The plasma membrane must be involved in transmitting the epinephrine signal.
The process involves three stages: reception, transduction, and response.
In reception, a chemical signal binds to a cellular protein, typically at the cell’s
surface or inside the cell.
In transduction, binding leads to a change in the receptor that triggers a series of
changes in a series of different molecules along a signal-transduction pathway. The
molecules in the pathway are called relay molecules.
In response, the transduced signal triggers a specific cellular activity.
The ability of organisms to reproduce their kind is the one characteristic that best
distinguishes living things from nonliving matter.
The continuity of life is based on the reproduction of cells, or cell division.
Cell division functions in reproduction, growth, and repair.
The division of a unicellular organism reproduces an entire organism, increasing the
population.
Cell division on a larger scale can produce progeny for some multicellular organisms.
This includes organisms that can grow by cuttings.
Cell division enables a multicellular organism to develop from a single fertilized egg or
zygote.
In a multicellular organism, cell division functions to repair and renew cells that die from
normal wear and tear or accidents.
Cell division is part of the cell cycle, the life of a cell from its origin in the division of a
parent cell until its own division into two.
Concept 12.2 The mitotic phase alternates with interphase in the cell cycle
The mitotic (M) phase of the cell cycle alternates with the much longer interphase.
The M phase includes mitosis and cytokinesis.
Interphase accounts for 90% of the cell cycle.
During interphase, the cell grows by producing proteins and cytoplasmic organelles,
copies its chromosomes, and prepares for cell division.
Interphase has three subphases: the G1 phase (“first gap”), the S phase (“synthesis”), and
the G2 phase (“second gap”).
During all three subphases, the cell grows by producing proteins and cytoplasmic
organelles such as mitochondria and endoplasmic reticulum.
However, chromosomes are duplicated only during the S phase.
Biology Chapter Notes
The daughter cells may then repeat the cycle.
A typical human cell might divide once every 24 hours.
Of this time, the M phase would last less than an hour, while the S phase might take
10–12 hours, or half the cycle.
The rest of the time would be divided between the G1 and G2 phases.
The G1 phase varies most in length from cell to cell.
Mitosis is a continuum of changes.
For convenience, mitosis is usually broken into five subphases: prophase,
prometaphase, metaphase, anaphase, and telophase.
In late interphase, the chromosomes have been duplicated but are not condensed.
A nuclear membrane bounds the nucleus, which contains one or more nucleoli.
The centrosome has replicated to form two centrosomes.
In animal cells, each centrosome features two centrioles.
In prophase, the chromosomes are tightly coiled, with sister chromatids joined together.
The nucleoli disappear.
The mitotic spindle begins to form.
It is composed of centrosomes and the microtubules that extend from them.
The radial arrays of shorter microtubules that extend from the centrosomes are called
asters.
The centrosomes move away from each other, apparently propelled by lengthening
microtubules.
During prometaphase, the nuclear envelope fragments, and microtubules from the spindle
interact with the condensed chromosomes.
Each of the two chromatids of a chromosome has a kinetochore, a specialized protein
structure located at the centromere.
Kinetochore microtubules from each pole attach to one of two kinetochores.
Nonkinetochore microtubules interact with those from opposite ends of the spindle.
The spindle fibers push the sister chromatids until they are all arranged at the metaphase
plate, an imaginary plane equidistant from the poles, defining metaphase.
At anaphase, the centromeres divide, separating the sister chromatids.
Each is now pulled toward the pole to which it is attached by spindle fibers.
By the end, the two poles have equivalent collections of chromosomes.
At telophase, daughter nuclei begin to form at the two poles.
Nuclear envelopes arise from the fragments of the parent cell’s nuclear envelope and
other portions of the endomembrane system.
The chromosomes become less tightly coiled.
Cytokinesis, division of the cytoplasm, is usually well underway by late telophase.
In animal cells, cytokinesis involves the formation of a cleavage furrow, which pinches
the cell in two.
In plant cells, vesicles derived from the Golgi apparatus produce a cell plate at the
middle of the cell.
The mitotic spindle distributes chromosomes to daughter cells: a closer look.
Living organisms are distinguished by their ability to reproduce their own kind.
Offspring resemble their parents more than they do less closely related individuals of the
same species.
The transmission of traits from one generation to the next is called heredity or inheritance.
However, offspring differ somewhat from parents and siblings, demonstrating variation.
Farmers have bred plants and animals for desired traits for thousands of years, but the
mechanisms of heredity and variation eluded biologists until the development of genetics
in the 20th century.
Genetics is the scientific study of heredity and variation.
Concept 13.3 Meiosis reduces the number of chromosome sets from diploid to haploid
Many steps of meiosis resemble steps in mitosis.
Both are preceded by the replication of chromosomes.
However, in meiosis, there are two consecutive cell divisions, meiosis I and meiosis II,
resulting in four daughter cells.
The first division, meiosis I, separates homologous chromosomes.
The second, meiosis II, separates sister chromatids.
The four daughter cells have only half as many chromosomes as the parent cell.
Meiosis I is preceded by interphase, in which the chromosomes are replicated to form
sister chromatids.
These are genetically identical and joined at the centromere.
The single centrosome is replicated, forming two centrosomes.
Division in meiosis I occurs in four phases: prophase I, metaphase I, anaphase I, and
telophase I.
Prophase I
Prophase I typically occupies more than 90% of the time required for meiosis.
During prophase I, the chromosomes begin to condense.
Homologous chromosomes loosely pair up along their length, precisely aligned gene for
gene.
In crossing over, DNA molecules in nonsister chromatids break at corresponding
places and then rejoin the other chromatid.
In synapsis, a protein structure called the synaptonemal complex forms between
homologues, holding them tightly together along their length.
As the synaptonemal complex disassembles in late prophase, each chromosome pair
becomes visible as a tetrad, or group of four chromatids.
Each tetrad has one or more chiasmata, sites where the chromatids of homologous
chromosomes have crossed and segments of the chromatids have been traded.
Biology Chapter Notes
Spindle microtubules form from the centrosomes, which have moved to the poles.
The breakdown of the nuclear envelope and nucleoli take place.
Kinetochores of each homologue attach to microtubules from one of the poles.
Metaphase I
At metaphase I, the tetrads are all arranged at the metaphase plate, with one chromosome
facing each pole.
Microtubules from one pole are attached to the kinetochore of one chromosome of
each tetrad, while those from the other pole are attached to the other.
Anaphase I
In anaphase I, the homologous chromosomes separate. One chromosome moves toward
each pole, guided by the spindle apparatus.
Sister chromatids remain attached at the centromere and move as a single unit toward the
pole.
Telophase I and cytokinesis
In telophase I, movement of homologous chromosomes continues until there is a haploid
set at each pole.
Each chromosome consists of two sister chromatids.
Cytokinesis usually occurs simultaneously, by the same mechanisms as mitosis.
In animal cells, a cleavage furrow forms. In plant cells, a cell plate forms.
No chromosome replication occurs between the end of meiosis I and the beginning of
meiosis II, as the chromosomes are already replicated.
Meiosis II
Meiosis II is very similar to mitosis.
During prophase II, a spindle apparatus forms and attaches to kinetochores of each
sister chromatid.
Spindle fibers from one pole attach to the kinetochore of one sister chromatid, and
those of the other pole attach to kinetochore of the other sister chromatid.
At metaphase II, the sister chromatids are arranged at the metaphase plate.
Because of crossing over in meiosis I, the two sister chromatids of each chromosome
are no longer genetically identical.
The kinetochores of sister chromatids attach to microtubules extending from opposite
poles.
At anaphase II, the centomeres of sister chromatids separate and two newly individual
chromosomes travel toward opposite poles.
In telophase II, the chromosomes arrive at opposite poles.
Nuclei form around the chromosomes, which begin expanding, and cytokinesis
separates the cytoplasm.
At the end of meiosis, there are four haploid daughter cells.
There are key differences between mitosis and meiosis.
Mitosis and meiosis have several key differences.
Concept 13.4 Genetic variation produced in sexual life cycles contributes to evolution
What is the origin of genetic variation?
Mutations are the original source of genetic diversity.
Once different versions of genes arise through mutation, reshuffling during meiosis and
fertilization produce offspring with their own unique set of traits.
Sexual life cycles produce genetic variation among offspring.
The behavior of chromosomes during meiosis and fertilization is responsible for most of
the variation that arises in each generation.
Three mechanisms contribute to genetic variation:
o Independent assortment of chromosomes.
o Crossing over.
o Random fertilization.
Independent assortment of chromosomes contributes to genetic variability due to the
random orientation of homologous pairs of chromosomes at the metaphase plate during
meiosis I.
There is a fifty-fifty chance that a particular daughter cell of meiosis I will get the
maternal chromosome of a certain homologous pair and a fifty-fifty chance that it will
receive the paternal chromosome.
Each homologous pair of chromosomes segregates independently of the other
homologous pairs during metaphase I.
Biology Chapter Notes
Therefore, the first meiotic division results in independent assortment of maternal and
paternal chromosomes into daughter cells.
The numbern of combinations possible when chromosomes assort independently into
gametes is 2 , where n is the haploid number of the organism.
If n = 3, there are 23 = 8 possible combinations.
For humans with n = 23, there are 223, or more than 8 million possible combinations of
chromosomes.
Crossing over produces recombinant chromosomes, which combine genes inherited
from each parent.
Crossing over begins very early in prophase I as homologous chromosomes pair up gene
by gene.
In crossing over, homologous portions of two nonsister chromatids trade places.
For humans, this occurs an average of one to three times per chromosome pair.
Recent research suggests that, in some organisms, crossing over may be essential for
synapsis and the proper assortment of chromosomes in meiosis I.
Crossing over, by combining DNA inherited from two parents into a single chromosome,
is an important source of genetic variation.
At metaphase II, nonidentical sister chromatids sort independently from one another,
increasing by even more the number of genetic types of daughter cells that are formed by
meiosis.
The random nature of fertilization adds to the genetic variation arising from meiosis.
Any sperm can fuse with any egg.
The ovum is one of more than 8 million possible chromosome combinations.
The successful sperm is one of more than 8 million possibilities.
The resulting zygote could contain any one of more than 70 trillion possible
combinations of chromosomes.
Crossing over adds even more variation to this.
Each zygote has a unique genetic identity.
The three sources of genetic variability in a sexually reproducing organism are:
o Independent assortment of homologous chromosomes during meiosis I and of
nonidentical sister chromatids during meiosis II.
o Crossing over between homologous chromosomes during prophase I.
o Random fertilization of an ovum by a sperm.
All three mechanisms reshuffle the various genes carried by individual members of a
population.
Evolutionary adaptation depends on a population’s genetic variation.
Darwin recognized the importance of genetic variation in evolution.
A population evolves through the differential reproductive success of its variant
members.
Those individuals best suited to the local environment leave the most offspring,
transmitting their genes in the process.
This natural selection results in adaptation, the accumulation of favorable genetic
variations.
Every day we observe heritable variations (such as brown, green, or blue eyes) among
individuals in a population.
These traits are transmitted from parents to offspring.
One possible explanation for heredity is a “blending” hypothesis.
This hypothesis proposes that genetic material contributed by each parent mixes in a
manner analogous to the way blue and yellow paints blend to make green.
With blending inheritance, a freely mating population will eventually give rise to a
uniform population of individuals.
Everyday observations and the results of breeding experiments tell us that heritable
traits do not blend to become uniform.
An alternative model, “particulate” inheritance, proposes that parents pass on discrete
heritable units, genes, that retain their separate identities in offspring.
Genes can be sorted and passed on, generation after generation, in undiluted form.
Modern genetics began in an abbey garden, where a monk named Gregor Mendel
documented a particulate mechanism of inheritance.
Concept 14.1 Mendel used the scientific approach to identify two laws of inheritance
Mendel discovered the basic principles of heredity by breeding garden peas in carefully
planned experiments.
Mendel grew up on a small farm in what is today the Czech Republic.
In 1843, Mendel entered an Augustinian monastery.
He studied at the University of Vienna from 1851 to 1853, where he was influenced by a
physicist who encouraged experimentation and the application of mathematics to science
and by a botanist who stimulated Mendel’s interest in the causes of variation in plants.
These influences came together in Mendel’s experiments.
After university, Mendel taught at the Brunn Modern School and lived in the local
monastery.
The monks at this monastery had a long tradition of interest in the breeding of plants,
including peas.
Around 1857, Mendel began breeding garden peas to study inheritance.
Pea plants have several advantages for genetic study.
Pea plants are available in many varieties with distinct heritable features, or
characters, with different variant traits.
Mendel could strictly control which plants mated with which.
Each pea plant has male (stamens) and female (carpal) sexual organs.
In nature, pea plants typically self-fertilize, fertilizing ova with the sperm nuclei from
their own pollen.
Biology Chapter Notes
However, Mendel could also use pollen from another plant for cross-pollination.
Mendel tracked only those characters that varied in an “either-or” manner, rather than a
“more-or-less” manner.
For example, he worked with flowers that were either purple or white.
He avoided traits, such as seed weight, that varied on a continuum.
Mendel started his experiments with varieties that were true-breeding.
When true-breeding plants self-pollinate, all their offspring have the same traits.
In a typical breeding experiment, Mendel would cross-pollinate (hybridize) two
contrasting, true-breeding pea varieties.
The true-breeding parents are the P generation, and their hybrid offspring are the F1
generation.
Mendel would then allow the F1 hybrids to self-pollinate to produce an F2 generation.
It was mainly Mendel’s quantitative analysis of F 2 plants that revealed two fundamental
principles of heredity: the law of segregation and the law of independent assortment.
By the law of segregation, the two alleles for a character are separated during the
formation of gametes.
If the blending model was correct, the F1 hybrids from a cross between purple-flowered
and white-flowered pea plants would have pale purple flowers.
Instead, F1 hybrids all have purple flowers, just as purple as their purple-flowered parents.
When Mendel allowed the F1 plants to self-fertilize, the F2 generation included both
purple-flowered and white-flowered plants.
The white trait, absent in the F1, reappeared in the F2.
Mendel used very large sample sizes and kept accurate records of his results.
Mendel recorded 705 purple-flowered F2 plants and 224 white-flowered F2 plants.
This cross produced a traits ratio of three purple to one white in the F2 offspring.
Mendel reasoned that the heritable factor for white flowers was present in the F 1 plants,
but did not affect flower color.
Purple flower color is a dominant trait, and white flower color is a recessive trait.
The reappearance of white-flowered plants in the F2 generation indicated that the heritable
factor for the white trait was not diluted or “blended” by coexisting with the purple-flower
factor in F1 hybrids.
Mendel found similar 3-to-1 ratios of two traits among F2 offspring when he conducted
crosses for six other characters, each represented by two different traits.
For example, when Mendel crossed two true-breeding varieties, one producing round
seeds and the other producing wrinkled seeds, all the F1 offspring had round seeds.
In the F2 plants, 75% of the seeds were round and 25% were wrinkled.
Mendel developed a hypothesis to explain these results that consisted of four related
ideas. We will explain each idea with the modern understanding of genes and
chromosomes.
o Alternative versions of genes account for variations in inherited characters.
The gene for flower color in pea plants exists in two versions, one for purple flowers
and one for white flowers.
These alternate versions are called alleles.
Each gene resides at a specific locus on a specific chromosome.
Biology Chapter Notes
The DNA at that locus can vary in its sequence of nucleotides.
The purple-flower and white-flower alleles are two DNA variations at the flower-color
locus.
o For each character, an organism inherits two alleles, one from each parent.
A diploid organism inherits one set of chromosomes from each parent.
Each diploid organism has a pair of homologous chromosomes and, therefore, two
copies of each gene.
These homologous loci may be identical, as in the true-breeding plants of the P
generation.
Alternatively, the two alleles may differ.
o If the two alleles at a locus differ, then one, the dominant allele, determines the
organism’s appearance. The other, the recessive allele, has no noticeable effect on the
organism’s appearance.
In the flower-color example, the F1 plants inherited a purple-flower allele from one
parent and a white-flower allele from the other.
They had purple flowers because the allele for that trait is dominant.
o 4. Mendel’s law of segregation states that the two alleles for a heritable character
separate and segregate during gamete production and end up in different gametes.
This segregation of alleles corresponds to the distribution of homologous
chromosomes to different gametes in meiosis.
If an organism has two identical alleles for a particular character, then that allele is
present as a single copy in all gametes.
If different alleles are present, then 50% of the gametes will receive one allele and
50% will receive the other.
Mendel’s law of segregation accounts for the 3:1 ratio that he observed in the F 2
generation.
The F1 hybrids produce two classes of gametes, half with the purple-flower allele and half
with the white-flower allele.
During self-pollination, the gametes of these two classes unite randomly.
This produces four equally likely combinations of sperm and ovum.
A Punnett square predicts the results of a genetic cross between individuals of known
genotype.
Let us describe a Punnett square analysis of the flower-color example.
We will use a capital letter to symbolize the dominant allele and a lowercase letter to
symbolize the recessive allele.
P is the purple-flower allele, and p is the white-flower allele.
What will be the physical appearance of the F2 offspring?
One in four F2 offspring will inherit two white-flower alleles and produce white
flowers.
Half of the F2 offspring will inherit one white-flower allele and one purple-flower
allele and produce purple flowers.
One in four F2 offspring will inherit two purple-flower alleles and produce purple
flowers.
Mendel’s model accounts for the 3:1 ratio in the F2 generation.
An organism with two identical alleles for a character is homozygous for that character.
Biology Chapter Notes
Organisms with two different alleles for a character is heterozygous for that character.
An organism’s traits are called its phenotype.
Its genetic makeup is called its genotype.
Two organisms can have the same phenotype but have different genotypes if one is
homozygous dominant and the other is heterozygous.
For flower color in peas, the only individuals with white flowers are those that are
homozygous recessive (pp) for the flower-color gene.
However, PP and Pp plants have the same phenotype (purple flowers) but different
genotypes (homozygous dominant and heterozygous).
How can we tell the genotype of an individual with the dominant phenotype?
The organism must have one dominant allele, but could be homozygous dominant or
heterozygous.
The answer is to carry out a testcross.
The mystery individual is bred with a homozygous recessive individual.
If any of the offspring display the recessive phenotype, the mystery parent must be
heterozygous.
By the law of independent assortment, each pair of alleles segregates independently into
gametes.
Mendel’s first experiments followed only a single character, such as flower color.
All F1 progeny produced in these crosses were monohybrids, heterozygous for one
character.
A cross between two heterozygotes is a monohybrid cross.
Mendel identified the second law of inheritance by following two characters at the same
time.
In one such dihybrid cross, Mendel studied the inheritance of seed color and seed shape.
The allele for yellow seeds (Y) is dominant to the allele for green seeds (y).
The allele for round seeds (R) is dominant to the allele for wrinkled seeds (r).
Mendel crossed true-breeding plants that had yellow, round seeds (YYRR) with true-
breeding plants that has green, wrinkled seeds (yyrr).
One possibility is that the two characters are transmitted from parents to offspring as a
package.
The Y and R alleles and y and r alleles stay together.
If this were the case, the F1 offspring would produce yellow, round seeds.
The F2 offspring would produce two phenotypes (yellow + round; green + wrinkled) in a
3:1 ratio, just like a monohybrid cross.
This was not consistent with Mendel’s results.
An alternative hypothesis is that the two pairs of alleles segregate independently of each
other.
The presence of a specific allele for one trait in a gamete has no impact on the presence
of a specific allele for the second trait.
In our example, the F1 offspring would still produce yellow, round seeds.
However, when the F1s produced gametes, genes would be packaged into gametes with all
possible allelic combinations.
Biology Chapter Notes
Four classes of gametes (YR, Yr, yR, and yr) would be produced in equal amounts.
When sperm with four classes of alleles and ova with four classes of alleles combined,
there would be 16 equally probable ways in which the alleles can combine in the F 2
generation.
These combinations produce four distinct phenotypes in a 9:3:3:1 ratio.
This was consistent with Mendel’s results.
Mendel repeated the dihybrid cross experiment for other pairs of characters and always
observed a 9:3:3:1 phenotypic ratio in the F2 generation.
Each character appeared to be inherited independently.
If you follow just one character in these crosses, you will observe a 3:1 F 2 ratio, just as if
this were a monohybrid cross.
The independent assortment of each pair of alleles during gamete formation is now called
Mendel’s law of independent assortment.
Mendel’s law of independent assortment states that each pair of alleles segregates
independently during gamete formation.
Strictly speaking, this law applies only to genes located on different, nonhomologous
chromosomes.
Genes located near each other on the same chromosome tend to be inherited together and
have more complex inheritance patterns than those predicted for the law of independent
assortment.
Concept 14.3 Inheritance patterns are often more complex than predicted by simple
Mendelian genetics
In the 20th century, geneticists have extended Mendelian principles not only to diverse
organisms, but also to patterns of inheritance more complex than Mendel described.
In fact, Mendel had the good fortune to choose a system that was relatively simple
genetically.
Each character that Mendel studied is controlled by a single gene.
Each gene has only two alleles, one of which is completely dominant to the other.
The heterozygous F1 offspring of Mendel’s crosses always looked like one of the parental
varieties because one allele was dominant to the other.
The relationship between genotype and phenotype is rarely so simple.
The inheritance of characters determined by a single gene deviates from simple
Mendelian patterns when alleles are not completely dominant or recessive, when a gene
has more than two alleles, or when a gene produces multiple phenotypes.
We will consider examples of each of these situations.
Alleles show different degrees of dominance and recessiveness in relation to each other.
One extreme is the complete dominance characteristic of Mendel’s crosses.
At the other extreme from complete dominance is codominance, in which two alleles
affect the phenotype in separate, distinguishable ways.
For example, the M, N, and MN blood groups of humans are due to the presence of
two specific molecules on the surface of red blood cells.
People of group M (genotype MM) have one type of molecule on their red blood cells,
people of group N (genotype NN) have the other type, and people of group MN
(genotype MN) have both molecules present.
The MN phenotype is not intermediate between M and N phenotypes but rather
exhibits both the M and the N phenotype.
Some alleles show incomplete dominance, in which heterozygotes show a distinct
intermediate phenotype not seen in homozygotes.
This is not blending inheritance because the traits are separable (particulate), as shown
in further crosses.
Concept 15.1 Mendelian inheritance has its physical basis in the behavior of
chromosomes
Around 1900, cytologists and geneticists began to see parallels between the behavior of
chromosomes and the behavior of Mendel’s factors.
Using improved microscopy techniques, cytologists worked out the process of mitosis
in 1875 and meiosis in the 1890s.
Chromosomes and genes are both present in pairs in diploid cells.
Homologous chromosomes separate and alleles segregate during meiosis.
Fertilization restores the paired condition for both chromosomes and genes.
Around 1902, Walter Sutton, Theodor Boveri, and others noted these parallels and a
chromosome theory of inheritance began to take form:
Genes occupy specific loci on chromosomes.
Chromosomes undergo segregation during meiosis.
Chromosomes undergo independent assortment during meiosis.
The behavior of homologous chromosomes during meiosis can account for the
segregation of the alleles at each genetic locus to different gametes.
The behavior of nonhomologous chromosomes can account for the independent
assortment of alleles for two or more genes located on different chromosomes.
Morgan traced a gene to a specific chromosome.
In the early 20th century, Thomas Hunt Morgan was the first geneticist to associate a
specific gene with a specific chromosome.
Like Mendel, Morgan made an insightful choice in his experimental animal. Morgan
worked with Drosophila melanogaster, a fruit fly that eats fungi on fruit.
Fruit flies are prolific breeders and have a generation time of two weeks.
Fruit flies have three pairs of autosomes and a pair of sex chromosomes (XX in
females, XY in males).
Morgan spent a year looking for variant individuals among the flies he was breeding.
He discovered a single male fly with white eyes instead of the usual red.
The normal character phenotype is the wild type.
Concept 15.2 Linked genes tend to be inherited together because they are located near
each other on the same chromosome
Each chromosome has hundreds or thousands of genes.
Genes located on the same chromosome that tend to be inherited together are called
linked genes.
Results of crosses with linked genes deviate from those expected according to
independent assortment.
Morgan observed this linkage and its deviations when he followed the inheritance of
characters for body color and wing size.
The wild-type body color is gray (b+), and the mutant is black (b).
The wild-type wing size is normal (vg+), and the mutant has vestigial wings (vg).
The mutant alleles are recessive to the wild-type alleles.
Neither gene is on a sex chromosome.
Morgan crossed F1 heterozygous females (b+bvg+vg) with homozygous recessive males
(bbvgvg).
According to independent assortment, this should produce 4 phenotypes in a 1:1:1:1 ratio.
Surprisingly, Morgan observed a large number of wild-type (gray-normal) and double-
mutant (black-vestigial) flies among the offspring.
These phenotypes are those of the parents.
Morgan reasoned that body color and wing shape are usually inherited together because
the genes for these characters are on the same chromosome.
The other two phenotypes (gray-vestigial and black-normal) were fewer than expected
from independent assortment (but totally unexpected from dependent assortment).
What led to this genetic recombination, the production of offspring with new
combinations of traits?
Independent assortment of chromosomes and crossing over produce genetic
recombinants.
Biology Chapter Notes
Genetic recombination can result from independent assortment of genes located on
nonhomologous chromosomes or from crossing over of genes located on homologous
chromosomes.
Mendel’s dihybrid cross experiments produced offspring that had a combination of traits
that did not match either parent in the P generation.
If the P generation consists of a yellow-round seed parent (YYRR) crossed with a
green-wrinkled seed parent (yyrr), all F1 plants have yellow-round seeds (YyRr).
A cross between an F1 plant and a homozygous recessive plant (a testcross) produces
four phenotypes.
Half are the parental types, with phenotypes that match the original P parents, with
either yellow-round seeds or green-wrinkled seeds.
Half are recombinants, new combinations of parental traits, with yellow-wrinkled or
green-round seeds.
A 50% frequency of recombination is observed for any two genes located on different
(nonhomologous) chromosomes.
The physical basis of recombination between unlinked genes is the random orientation of
homologous chromosomes at metaphase I of meiosis, which leads to the independent
assortment of alleles.
The F1 parent (YyRr) produces gametes with four different combinations of alleles: YR,
Yr, yR, and yr.
The orientation of the tetrad containing the seed-color gene has no bearing on the
orientation of the tetrad with the seed-shape gene.
In contrast, linked genes, genes located on the same chromosome, tend to move together
through meiosis and fertilization.
Under normal Mendelian genetic rules, we would not expect linked genes to recombine
into assortments of alleles not found in the parents.
If the seed color and seed coat genes were linked, we would expect the F 1 offspring to
produce only two types of gametes, YR and yr, when the tetrads separate.
One homologous chromosome carries the Y and R alleles on the same chromosome,
and the other homologous chromosome carries the y and r alleles.
The results of Morgan’s testcross for body color and wing shape did not conform to either
independent assortment or complete linkage.
Under independent assortment, the testcross should produce a 1:1:1:1 phenotypic ratio.
If completely linked, we should expect to see a 1:1:0:0 ratio with only parental
phenotypes among offspring.
Most of the offspring had parental phenotypes, suggesting linkage between the genes.
However, 17% of the flies were recombinants, suggesting incomplete linkage.
Morgan proposed that some mechanism must occasionally break the physical connection
between genes on the same chromosome.
This process, called crossing over, accounts for the recombination of linked genes.
Crossing over occurs while replicated homologous chromosomes are paired during
prophase of meiosis I.
One maternal and one paternal chromatid break at corresponding points and then rejoin
with each other.
The occasional production of recombinant gametes during meiosis accounts for the
occurrence of recombinant phenotypes in Morgan’s testcross.
Biology Chapter Notes
The percentage of recombinant offspring, the recombination frequency, is related to the
distance between linked genes.
Geneticists can use recombination data to map a chromosome’s genetic loci.
One of Morgan’s students, Alfred Sturtevant, used crossing over of linked genes to
develop a method for constructing a genetic map, an ordered list of the genetic loci along
a particular chromosome.
Sturtevant hypothesized that the frequency of recombinant offspring reflected the distance
between genes on a chromosome.
He assumed that crossing over is a random event, and that the chance of crossing over is
approximately equal at all points on a chromosome.
Sturtevant predicted that the farther apart two genes are, the higher the probability that a
crossover will occur between them, and therefore, the higher the recombination
frequency.
The greater the distance between two genes, the more points there are between them
where crossing over can occur.
Sturtevant used recombination frequencies from fruit fly crosses to map the relative
position of genes along chromosomes.
A genetic map based on recombination frequencies is called a linkage map.
Sturtevant used the testcross design to map the relative position of three fruit fly genes,
body color (b), wing size (vg), and eye color (cn).
The recombination frequency between cn and b is 9%.
The recombination frequency between cn and vg is 9.5%.
The recombination frequency between b and vg is 17%.
The only possible arrangement of these three genes places the eye color gene between
the other two.
Sturtevant expressed the distance between genes, the recombination frequency, as map
units.
One map unit (called a centimorgan) is equivalent to a 1% recombination frequency.
You may notice that the three recombination frequencies in our mapping example are not
quite additive: 9% (b-cn) + 9.5% (cn-vg) > 17% (b-vg).
This results from multiple crossing over events.
A second crossing over “cancels out” the first and reduces the observed number of
recombinant offspring.
Genes father apart (for example, b-vg) are more likely to experience multiple crossing
over events.
Some genes on a chromosome are so far apart that a crossover between them is virtually
certain.
In this case, the frequency of recombination reaches its maximum value of 50% and the
genes behave as if found on separate chromosomes.
In fact, two genes studied by Mendel—for seed color and flower color—are located on
the same chromosome but still assort independently.
Genes located far apart on a chromosome are mapped by adding the recombination
frequencies between the distant genes and the intervening genes.
Sturtevant and his colleagues were able to map the linear positions of genes in Drosophila
into four groups, one for each chromosome.
Biology Chapter Notes
A linkage map provides an imperfect picture of a chromosome.
Map units indicate relative distance and order, not precise locations of genes.
The frequency of crossing over is not actually uniform over the length of a
chromosome.
A linkage map does portray the order of genes along a chromosome, but does not
accurately portray the precise location of those genes.
Combined with other methods like chromosomal banding, geneticists can develop
cytogenetic maps of chromosomes.
These indicate the positions of genes with respect to chromosomal features.
Recent techniques show the physical distances between gene loci in DNA nucleotides.
Concept 15.5 Some inheritance patterns are exceptions to the standard chromosome
theory
The phenotypic effects of some mammalian genes depend on whether they are inherited
from the mother or the father.
For most genes, it is a reasonable assumption that a specific allele will have the same
effect regardless of whether it is inherited from the mother or father.
However, for a few dozen mammalian traits, phenotype varies depending on which parent
passed along the alleles for those traits.
The genes involved are not necessarily sex linked and may or may not lie on the X
chromosome.
Variation in phenotype depending on whether an allele is inherited from the male or
female parent is called genomic imprinting.
Genomic imprinting occurs during formation of gametes and results in the silencing of
certain genes.
Biology Chapter Notes
Imprinted genes are not expressed.
Because different genes are imprinted in sperm and ova, some genes in a zygote are
maternally imprinted, and others are paternally imprinted.
These maternal and paternal imprints are transmitted to all body cells during
development.
For a maternally imprinted gene, only the paternal allele is expressed.
For a paternally imprinted gene, only the maternal allele is expressed.
Patterns of imprinting are characteristic of a given species.
The gene for insulin-like growth factor 2 (Igf2) is one of the first imprinted genes to be
identified.
Although the growth factor is required for normal prenatal growth, only the paternal allele
is expressed.
Evidence that the Igf2 allele is imprinted initially came from crosses between wild-type
mice and dwarf mice homozygous for a recessive mutation in the Igf2 gene.
The phenotypes of heterozygous offspring differ, depending on whether the mutant
allele comes from the mother or the father.
The Igf2 allele is imprinted in eggs, turning off expression of the imprinted allele.
In sperm, the Igf2 allele is not imprinted and functions normally.
What exactly is a genomic imprint?
In many cases, it consists of methyl (—CH3) groups that are added to the cytosine
nucleotides of one of the alleles.
The hypothesis that methylation directly silences an allele is consistent with the evidence
that heavily methylated genes are usually inactive.
Other mechanisms may lead to silencing of imprinted genes.
Most of the known imprinted genes are critical for embryonic development.
In experiments with mice, embryos engineered to inherit both copies of certain
chromosomes from the same parent die before birth, whether their lone parent is male or
female.
Normal development requires that embryonic cells have one active copy of certain genes.
Aberrant imprinting is associated with abnormal development and certain cancers.
Extranuclear genes exhibit a non-Mendelian pattern of inheritance.
Not all of a eukaryote cell’s genes are located on nuclear chromosomes, or even in the
nucleus.
Extranuclear genes are found in small circles of DNA in mitochondria and chloroplasts.
These organelles reproduce themselves and transmit their genes to daughter organelles.
Their cytoplasmic genes do not display Mendelian inheritance, because they are not
distributed to offspring according to the same rules that direct distribution of nuclear
chromosomes during meiosis.
Karl Correns first observed cytoplasmic genes in plants in 1909 when he studied the
inheritance of patches of yellow or white on the leaves of an otherwise green plant.
He determined that the coloration of the offspring was determined by only the maternal
parent.
In April 1953, James Watson and Francis Crick shook the scientific world with an elegant
double-helical model for the structure of deoxyribonucleic acid, or DNA.
Your genetic endowment is the DNA you inherited from your parents.
Nucleic acids are unique in their ability to direct their own replication.
The resemblance of offspring to their parents depends on the precise replication of DNA
and its transmission from one generation to the next.
It is this DNA program that directs the development of your biochemical, anatomical,
physiological, and (to some extent) behavioral traits.
Concept 16.2 Many proteins work together in DNA replication and repair
The specific pairing of nitrogenous bases in DNA was the flash of inspiration that led
Watson and Crick to the correct double helix.
The possible mechanism for the next step, the accurate replication of DNA, was clear to
Watson and Crick from their double helix model.
During DNA replication, base pairing enables existing DNA strands to serve as templates
for new complementary strands.
In a second paper, Watson and Crick published their hypothesis for how DNA replicates.
Essentially, because each strand is complementary to the other, each can form a
template when separated.
The order of bases on one strand can be used to add complementary bases and
therefore duplicate the pairs of bases exactly.
When a cell copies a DNA molecule, each strand serves as a template for ordering
nucleotides into a new complementary strand.
One at a time, nucleotides line up along the template strand according to the base-
pairing rules.
The nucleotides are linked to form new strands.
Watson and Crick’s model, semiconservative replication, predicts that when a double
helix replicates, each of the daughter molecules will have one old strand and one newly
made strand.
Other competing models, the conservative model and the dispersive model, were also
proposed.
Experiments in the late 1950s by Matthew Meselson and Franklin Stahl supported the
semiconservative model proposed by Watson and Crick over the other two models.
In their experiments, 15they labeled the nucleotides of the old strands with a heavy
isotope of nitrogen ( N), while any new nucleotides were indicated by a lighter
isotope (14N).
Replicated strands could be separated by density in a centrifuge.
Each model—the semiconservative model, the conservative model, and the dispersive
model—made specific predictions about the density of replicated DNA strands.
The first replication in the 14N medium produced a band of hybrid (15N-14N) DNA,
eliminating the conservative model.
A second replication produced both light and hybrid DNA, eliminating the dispersive
model and supporting the semiconservative model.
Biology Chapter Notes
A large team of enzymes and other proteins carries out DNA replication.
It takes E. coli 25 minutes to copy each of the 5 million base pairs in its single
chromosome and divide to form two identical daughter cells.
A human cell can copy its 6 billion base pairs and divide into daughter cells in only a few
hours.
This process is remarkably accurate, with only one error per ten billion nucleotides.
More than a dozen enzymes and other proteins participate in DNA replication.
Much more is known about replication in bacteria than in eukaryotes.
The process appears to be fundamentally similar for prokaryotes and eukaryotes.
The replication of a DNA molecule begins at special sites, origins of replication.
In bacteria, this is a specific sequence of nucleotides that is recognized by the replication
enzymes.
These enzymes separate the strands, forming a replication “bubble.”
Replication proceeds in both directions until the entire molecule is copied.
In eukaryotes, there may be hundreds or thousands of origin sites per chromosome.
At the origin sites, the DNA strands separate, forming a replication “bubble” with
replication forks at each end.
The replication bubbles elongate as the DNA is replicated, and eventually fuse.
DNA polymerases catalyze the elongation of new DNA at a replication fork.
As nucleotides align with complementary bases along the template strand, they are added
to the growing end of the new strand by the polymerase.
The rate of elongation is about 500 nucleotides per second in bacteria and 50 per
second in human cells.
In E. coli, two different DNA polymerases are involved in replication: DNA polymerase
III and DNA polymerase I.
In eukaryotes, at least 11 different DNA polymerases have been identified so far.
Each nucleotide that is added to a growing DNA strand is a nucleoside triphosphate.
Each has a nitrogenous base, deoxyribose, and a triphosphate tail.
ATP is a nucleoside triphosphate with ribose instead of deoxyribose.
Like ATP, the triphosphate monomers used for DNA synthesis are chemically reactive,
partly because their triphosphate tails have an unstable cluster of negative charge.
As each nucleotide is added to the growing end of a DNA strand, the last two phosphate
groups are hydrolyzed to form pyrophosphate.
The exergonic hydrolysis of pyrophosphate to two inorganic phosphate molecules
drives the polymerization of the nucleotide to the new strand.
The strands in the double helix are antiparallel.
The sugar-phosphate backbones run in opposite directions.
Each DNA strand has a 3’ end with a free hydroxyl group attached to deoxyribose and
a 5’ end with a free phosphate group attached to deoxyribose.
The 5’ 3’ direction of one strand runs counter to the 3’ 5’ direction of the other
strand.
DNA polymerases can only add nucleotides to the free 3’ end of a growing DNA strand.
A new DNA strand can only elongate in the 5’ 3’ direction.
Biology Chapter Notes
Along one template strand, DNA polymerase III can synthesize a complementary strand
continuously by elongating the new DNA in the mandatory 5’ 3’ direction.
The DNA strand made by this mechanism is called the leading strand.
The other parental strand (5’ 3’ into the fork), the lagging strand, is copied away from
the fork.
Unlike the leading strand, which elongates continuously, the lagging stand is
synthesized as a series of short segments called Okazaki fragments.
Okazaki fragments are about 1,000–2,000 nucleotides long in E. coli and 100–200
nucleotides long in eukaryotes.
Another enzyme, DNA ligase, eventually joins the sugar-phosphate backbones of the
Okazaki fragments to form a single DNA strand.
DNA polymerases cannot initiate synthesis of a polynucleotide.
They can only add nucleotides to the 3’ end of an existing chain that is base-paired
with the template strand.
The initial nucleotide chain is called a primer.
In the initiation of the replication of cellular DNA, the primer is a short stretch of RNA
with an available 3’ end.
The primer is 5–10 nucleotides long in eukaryotes.
Primase, an RNA polymerase, links ribonucleotides that are complementary to the DNA
template into the primer.
RNA polymerases can start an RNA chain from a single template strand.
After formation of the primer, DNA pol III adds a deoxyribonucleotide to the 3’ end of
the RNA primer and continues adding DNA nucleotides to the growing DNA strand
according to the base-pairing rules.
Returning to the original problem at the replication fork, the leading strand requires the
formation of only a single primer as the replication fork continues to separate.
For synthesis of the lagging strand, each Okazaki fragment must be primed separately.
Another DNA polymerase, DNA polymerase I, replaces the RNA nucleotides of the
primers with DNA versions, adding them one by one onto the 3’ end of the adjacent
Okazaki fragment.
The primers are converted to DNA before DNA ligase joins the fragments together.
In addition to primase, DNA polymerases, and DNA ligases, several other proteins have
prominent roles in DNA synthesis.
Helicase untwists the double helix and separates the template DNA strands at the
replication fork.
This untwisting causes tighter twisting ahead of the replication fork, and
topoisomerase helps relieve this strain.
Single-strand binding proteins keep the unpaired template strands apart during
replication.
To summarize, at the replication fork, the leading strand is copied continuously into the
fork from a single primer.
The lagging strand is copied away from the fork in short segments, each requiring a
new primer.
It is conventional and convenient to think of the DNA polymerase molecules as moving
along a stationary DNA template.
Biology Chapter Notes
In reality, the various proteins involved in DNA replication form a single large complex, a
DNA replication “machine.”
Many protein-protein interactions facilitate the efficiency of this machine.
For example, helicase works much more rapidly when it is in contact with primase.
The DNA replication machine is probably stationary during the replication process.
In eukaryotic cells, multiple copies of the machine may anchor to the nuclear matrix, a
framework of fibers extending through the interior of the nucleus.
The DNA polymerase molecules “reel in” the parental DNA and “extrude” newly made
daughter DNA molecules.
Enzymes proofread DNA during its replication and repair damage in existing DNA.
Mistakes during the initial pairing of template nucleotides and complementary nucleotides
occur at a rate of one error per 100,000 base pairs.
DNA polymerase proofreads each new nucleotide against the template nucleotide as soon
as it is added.
If there is an incorrect pairing, the enzyme removes the wrong nucleotide and then
resumes synthesis.
The final error rate is only one per ten billion nucleotides.
DNA molecules are constantly subject to potentially harmful chemical and physical
agents.
Reactive chemicals, radioactive emissions, X-rays, and ultraviolet light can change
nucleotides in ways that can affect encoded genetic information.
DNA bases may undergo spontaneous chemical changes under normal cellular
conditions.
Mismatched nucleotides that are missed by DNA polymerase or mutations that occur after
DNA synthesis is completed can often be repaired.
Each cell continually monitors and repairs its genetic material, with 100 repair
enzymes known in E. coli and more than 130 repair enzymes identified in humans.
In mismatch repair, special enzymes fix incorrectly paired nucleotides.
A hereditary defect in one of these enzymes is associated with a form of colon cancer.
In nucleotide excision repair, a nuclease cuts out a segment of a damaged strand.
DNA polymerase and ligase fill in the gap.
The importance of the proper functioning of repair enzymes is clear from the inherited
disorder xeroderma pigmentosum.
These individuals are hypersensitive to sunlight.
Ultraviolet light can produce thymine dimers between adjacent thymine nucleotides.
This buckles the DNA double helix and interferes with DNA replication.
In individuals with this disorder, mutations in their skin cells are left uncorrected and
cause skin cancer.
The ends of DNA molecules are replicated by a special mechanism.
Limitations of DNA polymerase create problems for the linear DNA of eukaryotic
chromosomes.
The usual replication machinery provides no way to complete the 5’ ends of daughter
DNA strands.
Biology Chapter Notes
Repeated rounds of replication produce shorter and shorter DNA molecules.
Prokaryotes do not have this problem because they have circular DNA molecules without
ends.
The ends of eukaryotic chromosomal DNA molecules, the telomeres, have special
nucleotide sequences.
Telomeres do not contain genes. Instead, the DNA typically consists of multiple
repetitions of one short nucleotide sequence.
In human telomeres, this sequence is typically TTAGGG, repeated between 100 and
1,000 times.
Telomeres protect genes from being eroded through multiple rounds of DNA replication.
Telomeric DNA tends to be shorter in dividing somatic cells of older individuals and
in cultured cells that have divided many times.
It is possible that the shortening of telomeres is somehow connected with the aging
process of certain tissues and perhaps to aging in general.
Telomeric DNA and specific proteins associated with it also prevents the staggered ends
of the daughter molecule from activating the cell’s system for monitoring DNA damage.
Eukaryotic cells have evolved a mechanism to restore shortened telomeres in germ cells,
which give rise to gametes.
If the chromosomes of germ cells became shorter with every cell cycle, essential genes
would eventually be lost.
An enzyme called telomerase catalyzes the lengthening of telomeres in eukaryotic germ
cells, restoring their original length.
Telomerase uses a short molecule of RNA as a template to extend the 3’ end of the
telomere.
There is now room for primase and DNA polymerase to extend the 5’ end.
It does not repair the 3’-end “overhang,” but it does lengthen the telomere.
Telomerase is not present in most cells of multicellular organisms.
Therefore, the DNA of dividing somatic cells and cultured cells tends to become shorter.
Telomere length may be a limiting factor in the life span of certain tissues and of the
organism.
Normal shortening of telomeres may protect organisms from cancer by limiting the
number of divisions that somatic cells can undergo.
Cells from large tumors often have unusually short telomeres, because they have gone
through many cell divisions.
Active telomerase has been found in some cancerous somatic cells.
This overcomes the progressive shortening that would eventually lead to self-
destruction of the cancer.
Immortal strains of cultured cells are capable of unlimited cell division.
Telomerase may provide a useful target for cancer diagnosis and chemotherapy.
The information content of DNA is in the form of specific sequences of nucleotides along
the DNA strands.
The DNA inherited by an organism leads to specific traits by dictating the synthesis of
proteins.
Gene expression, the process by which DNA directs protein synthesis, includes two stages
called transcription and translation.
Proteins are the links between genotype and phenotype.
For example, Mendel’s dwarf pea plants lack a functioning copy of the gene that
specifies the synthesis of a key protein, gibberellin.
Gibberellins stimulate the normal elongation of stems.
Concept 17.6 Comparing gene expression in prokaryotes and eukaryotes reveals key
differences
Although prokaryotes and eukaryotes carry out transcription and translation in very
similar ways, they do have differences in cellular machinery and in details of the
processes.
Eukaryotic RNA polymerases differ from those of prokaryotes and require
transcription factors.
They differ in how transcription is terminated.
Their ribosomes also are different.
One major difference is that prokaryotes can transcribe and translate the same gene
simultaneously.
The new protein quickly diffuses to its operating site.
In eukaryotes, the nuclear envelope segregates transcription from translation.
In addition, extensive RNA processing is carried out between these processes.
This provides additional steps whose regulation helps coordinate the elaborate
activities of a eukaryotic cell.
Eukaryotic cells also have complicated mechanisms for targeting proteins to the
appropriate organelle.
Concept 17.7 Point mutations can affect protein structure and function
Mutations are changes in the genetic material of a cell (or virus).
These include large-scale mutations in which long segments of DNA are affected (for
example, translocations, duplications, and inversions).
A chemical change in just one base pair of a gene causes a point mutation.
If these occur in gametes or cells producing gametes, they may be transmitted to future
generations.
Viruses and bacteria are the simplest biological systems—microbial models in which
scientists find life’s fundamental molecular mechanisms in their most basic, accessible
forms.
Molecular biology was born in the laboratories of microbiologists studying viruses and
bacteria.
Microbes such as E. coli and its viruses are called model systems because of their use
in studies that reveal broad biological principles.
Microbiologists provided most of the evidence that genes are made of DNA, and they
worked out most of the major steps in DNA replication, transcription, and translation.
Techniques enabling scientists to manipulate genes and transfer them from one
organism to another were developed in microbes.
In addition, viruses and bacteria have unique genetic features with implications for
understanding the diseases that they cause.
Bacteria are prokaryotic organisms, with cells that are much smaller and more simply
organized than those of eukaryotes, such as plants and animals.
Viruses are smaller and simpler still, lacking the structure and metabolic machinery of
cells.
Most viruses are little more than aggregates of nucleic acids and protein—genes in a
protein coat.
Concept 18.1 A virus has a genome but can reproduce only within a host cell
Researchers discovered viruses by studying a plant disease.
The story of how viruses were discovered begins in 1883 with research on the cause of
tobacco mosaic disease by Adolf Mayer.
This disease stunts tobacco plant growth and mottles plant leaves.
Mayer concluded that the disease was infectious when he found that he could transmit
the disease by rubbing sap from diseased leaves onto healthy plants.
He concluded that the disease must be caused by an extremely small bacterium.
Ten years later, Dimitri Ivanovsky demonstrated that the sap was still infectious even
after passing through a filter designed to remove bacteria.
In 1897, Martinus Beijerinck ruled out the possibility that the disease was due to a
filterable toxin produced by a bacterium by demonstrating that the infectious agent could
reproduce.
The sap from one generation of infected plants could be used to infect a second
generation of plants that could infect subsequent generations.
Beijerinck also determined that the pathogen could reproduce only within the host,
could not be cultivated on nutrient media, and was not inactivated by alcohol,
generally lethal to bacteria.
Biology Chapter Notes
In 1935, Wendell Stanley crystallized the pathogen, the tobacco mosaic virus (TMV).
A virus is a genome enclosed in a protective coat.
Stanley’s discovery that some viruses could be crystallized was puzzling because not even
the simplest cells can aggregate into regular crystals.
However, viruses are not cells.
They are infectious particles consisting of nucleic acid encased in a protein coat and, in
some cases, a membranous envelope.
The tiniest viruses are only 20 nm in diameter—smaller than a ribosome.
The genome of viruses may consist of double-stranded DNA, single-stranded DNA,
double-stranded RNA, or single-stranded RNA, depending on the kind of virus.
A virus is called a DNA virus or an RNA virus, according to the kind of nucleic acid
that makes up its genome.
The viral genome is usually organized as a single linear or circular molecule of nucleic
acid.
The smallest viruses have only four genes, while the largest have several hundred.
The capsid is the protein shell enclosing the viral genome.
Capsids are built of a large number of protein subunits called capsomeres.
The number of different kinds of proteins making up the capsid is usually small.
The capsid of the tobacco mosaic virus has more than 1,000 copies of the same protein.
Adenoviruses have 252 identical proteins arranged into a polyhedral capsid—as an
icosahedron.
Some viruses have accessory structures to help them infect their hosts.
A membranous envelope surrounds the capsids of flu viruses.
These viral envelopes are derived from the membrane of the host cell.
They also have some host cell viral proteins and glycoproteins, as well as molecules of
viral origin.
Some viruses carry a few viral enzyme molecules within their capsids.
The most complex capsids are found in viruses that infect bacteria, called bacteriophages
or phages.
The T-even phages (T2, T4, T6) that infect Escherichia coli have elongated icosahedral
capsid heads that enclose their DNA and a protein tailpiece that attaches the phage to the
host and injects the phage DNA inside.
Viruses can reproduce only within a host cell.
Viruses are obligate intracellular parasites.
They can reproduce only within a host cell.
An isolated virus is unable to reproduce—or do anything else, except infect an appropriate
host.
Viruses lack the enzymes for metabolism and the ribosomes for protein synthesis.
An isolated virus is merely a packaged set of genes in transit from one host cell to
another.
Each type of virus can infect and parasitize only a limited range of host cells, called its
host range.
Biology Chapter Notes
This host specificity depends on the evolution of recognition systems by the virus.
Viruses identify host cells by a “lock and key” fit between proteins on the outside of
the virus and specific receptor molecules on the host’s surface (which evolved for
functions that benefit the host).
Some viruses have a broad enough host range to infect several species, while others infect
only a single species.
West Nile virus can infect mosquitoes, birds, horses, and humans.
Measles virus can infect only humans.
Most viruses of eukaryotes attack specific tissues.
Human cold viruses infect only the cells lining the upper respiratory tract.
The AIDS virus binds only to certain white blood cells.
A viral infection begins when the genome of the virus enters the host cell.
Once inside, the viral genome commandeers its host, reprogramming the cell to copy viral
nucleic acid and manufacture proteins from the viral genome.
The host provides nucleotides, ribosomes, tRNAs, amino acids, ATP, and other
components for making the viral components dictated by viral genes.
Most DNA viruses use the DNA polymerases of the host cell to synthesize new genomes
along the templates provided by the viral DNA.
RNA viruses use special virus-encoded polymerases that can use RNA as a template.
The nucleic acid molecules and capsomeres then self-assemble into viral particles and exit
the cell.
Tobacco mosaic virus RNA and capsomeres can be assembled to form complete
viruses if the components are mixed together under the right conditions.
The simplest type of viral reproductive cycle ends with the exit of many viruses from the
infected host cell, a process that usually damages or destroys the host cell.
Phages reproduce using lytic or lysogenic cycles.
While phages are the best understood of all viruses, some of them are also among the
most complex.
Research on phages led to the discovery that some double-stranded DNA viruses can
reproduce by two alternative mechanisms: the lytic cycle and the lysogenic cycle.
In the lytic cycle, the phage reproductive cycle culminates in the death of the host.
In the last stage, the bacterium lyses (breaks open) and releases the phages produced
within the cell to infect others.
Each of these phages can infect a healthy cell.
Virulent phages reproduce only by a lytic cycle.
While phages have the potential to wipe out a bacterial colony in just hours, bacteria have
defenses against phages.
Natural selection favors bacterial mutants with receptor sites that are no longer
recognized by a particular type of phage.
Bacteria produce restriction endonucleases, or restriction enzymes, that recognize and
cut up foreign DNA, including certain phage DNA.
Chemical modifications to the bacteria’s own DNA prevent its destruction by
restriction nucleases.
Natural selection also favors phage mutants that are resistant to restriction enzymes.
Biology Chapter Notes
In the lysogenic cycle, the phage genome replicates without destroying the host cell.
Temperate phages, like phage lambda, use both lytic and lysogenic cycles.
The lambda phage that infects E. coli demonstrates the cycles of a temperate phage.
Infection of an E. coli by phage lambda begins when the phage binds to the surface of the
cell and injects its DNA.
What happens next depends on the reproductive mode: lytic or lysogenic cycle.
During a lytic cycle, the viral genes turn the host cell into a lambda phage-producing
factory, and the cell lyses and releases its viral products.
During a lysogenic cycle, the viral DNA molecule is incorporated by genetic
recombination into a specific site on the host cell’s chromosome.
In this prophage stage, one of the viral genes codes for a protein that represses most other
prophage genes.
As a result, the phage genome is largely silent.
A few other prophage genes may also be expressed during lysogenic cycles.
Expression of these genes may alter the host’s phenotype, which can have medical
significance.
Every time the host divides, it copies the phage DNA and passes the copies to daughter
cells.
The viruses propagate without killing the host cells on which they depend.
The term lysogenic implies that prophages are capable of giving rise to active phages that
lyse their host cells.
That happens when the viral genome exits the bacterial chromosome and initiates a lytic
cycle.
Animal viruses are diverse in their modes of infection and replication.
Many variations on the basic scheme of viral infection and reproduction are represented
among animal viruses.
One key variable is the type of nucleic acid that serves as a virus’s genetic material.
Another variable is the presence or absence of a membranous envelope derived from
the host cell membrane.
Most animal viruses with RNA genomes have an envelope, as do some with DNA
genomes.
Viruses equipped with an outer envelope use the envelope to enter the host cell.
Glycoproteins on the envelope bind to specific receptors on the host’s membrane.
The envelope fuses with the host’s membrane, transporting the capsid and viral
genome inside.
The viral genome duplicates and directs the host’s protein synthesis machinery to
synthesize capsomeres with free ribosomes and glycoproteins with bound ribosomes.
After the capsid and viral genome self-assemble, they bud from the host cell covered
with an envelope derived from the host’s plasma membrane, including viral
glycoproteins.
The viral envelope is thus derived from the host’s plasma membrane, although viral genes
specify some of the molecules in the membrane.
These enveloped viruses do not necessarily kill the host cell.
Some viruses have envelopes that are not derived from plasma membrane.
Biology Chapter Notes
The envelope of the herpesvirus is derived from the nuclear envelope of the host.
These double-stranded DNA viruses reproduce within the cell nucleus using viral and
cellular enzymes to replicate and transcribe their DNA.
In some cases, copies of the herpesvirus DNA remain behind as minichromosomes in
the nuclei of certain nerve cells.
There they remain for life until triggered by physical or emotional stress to leave the
genome and initiate active viral production.
The infection of other cells by these new viruses causes cold or genital sores.
The viruses that use RNA as the genetic material are quite diverse, especially those that
infect animals.
In some with single-stranded RNA (class IV), the genome acts as mRNA and is
translated directly.
In others (class V), the RNA genome serves as a template for complementary RNA
strands, which function both as mRNA and as templates for the synthesis of additional
copies of genome RNA.
All viruses that require RNA RNA synthesis to make mRNA use a viral enzyme
that is packaged with the genome inside the capsid.
Retroviruses (class VI) have the most complicated life cycles.
These carry an enzyme called reverse transcriptase that transcribes DNA from an
RNA template.
This provides RNA DNA information flow.
The newly made DNA is inserted as a provirus into a chromosome in the animal cell.
The host’s RNA polymerase transcribes the viral DNA into more RNA molecules.
These can function both as mRNA for the synthesis of viral proteins and as
genomes for new virus particles released from the cell.
Human immunodeficiency virus (HIV), the virus that causes AIDS (acquired
immunodeficiency syndrome) is a retrovirus.
The reproductive cycle of HIV illustrates the pattern of infection and replication in a
retrovirus.
The viral particle includes an envelope with glycoproteins for binding to specific types of
red blood cells, a capsid containing two identical RNA strands as its genome, and two
copies of reverse transcriptase.
After HIV enters the host cell, reverse transcriptase molecules are released into the
cytoplasm and catalyze synthesis of viral DNA.
The host’s polymerase transcribes the proviral DNA into RNA molecules that can
function both as mRNA for the synthesis of viral proteins and as genomes for new virus
particles released from the cell.
Transcription produces more copies of the viral RNA that are translated into viral
proteins, which self-assemble into a virus particle and leave the host.
Viruses may have evolved from other mobile genetic elements.
Viruses do not fit our definition of living organisms.
An isolated virus is biologically inert, and yet it has a genetic program written in the
universal language of life.
Although viruses are obligate intracellular parasites that cannot reproduce independently,
it is hard to deny their evolutionary connection to the living world.
Biology Chapter Notes
Because viruses depend on cells for their own propagation, it is reasonable to assume that
they evolved after the first cells appeared.
Most molecular biologists favor the hypothesis that viruses originated from fragments of
cellular nucleic acids that could move from one cell to another.
A viral genome usually has more in common with the genome of its host than with
those of viruses infecting other hosts.
However, some viruses have genetic sequences that are quite similar to seemingly
distantly related viruses.
This genetic similarity may reflect the persistence of groups of viral genes that were
evolutionarily successful during the early evolution of viruses and their eukaryotic
host cells.
Perhaps the earliest viruses were naked bits of nucleic acids that passed between cells via
injured cell surfaces.
The evolution of capsid genes may have facilitated the infection of undamaged cells.
Candidates for the original sources of viral genomes include plasmids and transposable
elements.
Plasmids are small, circular DNA molecules that are separate from chromosomes.
Plasmids, found in bacteria and in eukaryote yeast, can replicate independently of the
rest of the cell and are occasionally transferred between cells.
Transposable elements are DNA segments that can move from one location to another
within a cell’s genome.
Both plasmids and transposable elements are mobile genetic elements.
The ongoing evolutionary relationship between viruses and the genomes of their hosts is
an association that makes viruses very useful model systems in molecular biology.
Concept 18.2 Viruses, viroids, and prions are formidable pathogens in animals and
plants
The link between viral infection and the symptoms it produces is often obscure.
Some viruses damage or kill cells by triggering the release of hydrolytic enzymes from
lysosomes.
Some viruses cause the infected cell to produce toxins that lead to disease symptoms.
Others have molecular components, such as envelope proteins, that are toxic.
In some cases, viral damage is easily repaired (respiratory epithelium after a cold), but in
others, infection causes permanent damage (nerve cells after polio).
Many of the temporary symptoms associated with a viral infection result from the body’s
own efforts at defending itself against infection.
The immune system is a complex and critical part of the body’s natural defense
mechanism against viral and other infections.
Modern medicine has developed vaccines, harmless variants or derivatives of pathogenic
microbes that stimulate the immune system to mount defenses against the actual pathogen.
Vaccination has eradicated smallpox.
Effective vaccines are available against polio, measles, rubella, mumps, hepatitis B,
and a number of other viral diseases.
Medical technology can do little to cure viral diseases once they occur.
Biology Chapter Notes
Antibiotics, which can kill bacteria by inhibiting enzymes or processes specific to
bacteria, are powerless against viruses, which have few or no enzymes of their own.
Most antiviral drugs resemble nucleosides and interfere with viral nucleic acid
synthesis.
An example is acyclovir, which impedes herpesvirus reproduction by inhibiting the
viral polymerase that synthesizes viral DNA.
Azidothymidine (AZT) curbs HIV reproduction by interfering with DNA synthesis by
reverse transcriptase.
Currently, multidrug “cocktails” are the most effective treatment for HIV.
New viral diseases are emerging.
In recent years, several emerging viruses have risen to prominence.
HIV, the AIDS virus, seemed to appear suddenly in the early 1980s.
Each year new strains of influenza virus cause millions to miss work or class, and
deaths are not uncommon.
The deadly Ebola virus has caused hemorrhagic fevers in central Africa periodically
since 1976.
West Nile virus appeared for the first time in North America in 1999.
A more recent viral disease is severe acute respiratory syndrome (SARS).
Researchers identified the disease agent causing SARS as a coronavirus, a class IV
virus with a single-stranded RNA genome.
The emergence of these new viral diseases is due to three processes: mutation; spread of
existing viruses from one species to another; and dissemination of a viral disease from a
small, isolated population.
Mutation of existing viruses is a major source of new viral diseases.
RNA viruses tend to have high mutation rates because replication of their nucleic acid
lacks proofreading.
Some mutations create new viral strains with sufficient genetic differences from earlier
strains that they can infect individuals who had acquired immunity to these earlier
strains.
This is the case in flu epidemics.
Another source of new viral diseases is the spread of existing viruses from one host
species to another.
It is estimated that about three-quarters of new human diseases originated in other
animals.
For example, hantavirus, which killed dozens of people in 1993, normally infects
rodents, especially deer mice.
In 1993, unusually wet weather in the southwestern United States increased the mice’s
food, exploding the population.
Humans acquired hantavirus when they inhaled dust-containing traces of urine and
feces from infected mice.
The source of the SARS-causing virus is still undetermined, but candidates include the
exotic animal markets in China.
In early 2004, the first cases of a new bird flu were reported in southeast Asia.
If this disease evolves to spread from person to person, the potential for a major
human outbreak is great.
Concept 19.2 Gene expression can be regulated at any stage, but the key step is
transcription
Like unicellular organisms, the tens of thousands of genes in the cells of multicellular
eukaryotes are continually turned on and off in response to signals from their internal and
external environments.
Gene expression must be controlled on a long-term basis during cellular differentiation,
the divergence in form and function as cells in a multicellular organism specialize.
A typical human cell probably expresses about 20% of its genes at any given time.
Biology Chapter Notes
Highly specialized cells, such as nerves or muscles, express only a tiny fraction of
their genes.
Although all the cells in an organism contain an identical genome, the subset of
genes expressed in the cells of each type is unique.
The differences between cell types are due to differential gene expression, the
expression of different genes by cells with the same genome.
The genomes of eukaryotes may contain tens of thousands of genes.
For quite a few species, only a small amount of the DNA—1.5% in humans—codes
for protein.
Of the remaining DNA, a very small fraction consists of genes for rRNA and tRNA.
Most of the rest of the DNA seems to be largely noncoding, although researchers have
found that a significant amount of it is transcribed into RNAs of unknown function.
Problems with gene expression and control can lead to imbalance and diseases, including
cancers.
Our understanding of the mechanisms controlling gene expression in eukaryotes has been
enhanced by new research methods, including advances in DNA technology.
In all organisms, the expression of specific genes is most commonly regulated at
transcription, often in response to signals coming from outside the cell.
The term gene expression is often equated with transcription.
With their greater complexity, eukaryotes have opportunities for controlling gene
expression at additional stages.
Each stage in the entire process of gene expression provides a potential control point
where gene expression can be turned on or off, sped up or slowed down.
A web of control connects different genes and their products.
These levels of control include chromatin packing, transcription, RNA processing,
translation, and various alterations to the protein product.
Chromatin modifications affect the availability of genes for transcription.
In addition to its role in packing DNA inside the nucleus, chromatin organization
regulates gene expression.
Genes of densely condensed heterochromatin are usually not expressed, presumably
because transcription proteins cannot reach the DNA.
A gene’s location relative to nucleosomes and to attachment sites to the chromosome
scaffold or nuclear lamina can affect transcription.
Chemical modifications of chromatin play a key role in chromatin structure and gene
expression.
Chemical modifications of histones play a direct role in the regulation of gene
transcription.
The N-terminus of each histone molecule in a nucleosome protrudes outward from the
nucleosome.
These histone tails are accessible to various modifying enzymes, which catalyze the
addition or removal of specific chemical groups.
Histone acetylation (addition of an acetyl group —COCH3) and deacetylation appear to
play a direct role in the regulation of gene transcription.
Acetylated histones grip DNA less tightly, providing easier access for transcription
proteins in this region.
Biology Chapter Notes
Some of the enzymes responsible for acetylation or deacetylation are associated with
or are components of transcription factors that bind to promoters.
Thus histone acetylation enzymes may promote the initiation of transcription not only
by modifying chromatin structure, but also by binding to and recruiting components of
the transcription machinery.
DNA methylation is the attachment by specific enzymes of methyl groups (—CH3) to
DNA bases after DNA synthesis.
Inactive DNA is generally highly methylated compared to DNA that is actively
transcribed.
For example, the inactivated mammalian X chromosome in females is heavily
methylated.
Genes are usually more heavily methylated in cells where they are not expressed.
Demethylating certain inactive genes turns them on.
However, there are exceptions to this pattern.
DNA methylation proteins recruit histone deacetylation enzymes, providing a mechanism
by which DNA methylation and histone deacetylation cooperate to repress transcription.
In some species, DNA methylation is responsible for long-term inactivation of genes
during cellular differentiation.
Once methylated, genes usually stay that way through successive cell divisions.
Methylation enzymes recognize sites on one strand that are already methylated and
correctly methylate the daughter strand after each round of DNA replication.
This methylation patterns accounts for genomic imprinting in which methylation turns
off either the maternal or paternal alleles of certain genes at the start of development.
The chromatin modifications just discussed do not alter DNA sequence, and yet they may
be passed along to future generations of cells.
Inheritance of traits by mechanisms not directly involving the nucleotide sequence is
called epigenetic inheritance.
Researchers are amassing more and more evidence for the importance of epigenetic
information in the regulation of gene expression.
Enzymes that modify chromatin structure are integral parts of the cell’s machinery for
regulating transcription.
Transcription initiation is controlled by proteins that interact with DNA and with each
other.
Chromatin-modifying enzymes provide initial control of gene expression by making a
region of DNA either more available or less available for transcription.
A cluster of proteins called a transcription initiation complex assembles on the promoter
sequence at the “upstream” end of the gene.
One component, RNA polymerase II, transcribes the gene, synthesizing a primary
RNA transcript or pre-mRNA.
RNA processing includes enzymatic addition of a 5’ cap and a poly-A tail, as well as
splicing out of introns to yield a mature mRNA.
Multiple control elements are associated with most eukaryotic genes.
Control elements are noncoding DNA segments that regulate transcription by binding
certain proteins.
Concept 19.3 Cancer results from genetic changes that affect cell cycle control
Cancer is a disease in which cells escape the control methods that normally regulate cell
growth and division.
The gene regulation systems that go wrong during cancer are the very same systems
that play important roles in embryonic development, the immune response, and other
biological processes.
The genes that normally regulate cell growth and division during the cell cycle include
genes for growth factors, their receptors, and the intracellular molecules of signaling
pathways.
Mutations altering any of these genes in somatic cells can lead to cancer.
The agent of such changes can be random spontaneous mutations or environmental
influences such as chemical carcinogens, X-rays, or certain viruses.
In 1911, Peyton Rous discovered a virus that causes cancer in chickens.
Since then, scientists have recognized a number of tumor viruses that cause cancer in
various animals, including humans.
All tumor viruses transform cells into cancer cells through the integration of viral
nucleic acid into host cell DNA.
Cancer-causing genes, oncogenes, were initially discovered in retroviruses, but close
counterparts, proto-oncogenes, have been found in other organisms.
The products of proto-oncogenes are proteins that stimulate normal cell growth and
division and play essential functions in normal cells.
A proto-oncogene becomes an oncogene following genetic changes that lead to an
increase in the proto-oncogene’s protein production or the activity of each protein
molecule.
These genetic changes include movements of DNA within the genome, amplification
of the proto-oncogene, and point mutations in the control element of the proto-
oncogene.
Cancer cells frequently have chromosomes that have been broken and rejoined
incorrectly.
This may translocate a fragment to a location near an active promoter or other control
element.
Movement of transposable elements may also place a more active promoter near a
proto-oncogene, increasing its expression.
Amplification increases the number of copies of the proto-oncogene in the cell.
A point mutation in the promoter or enhancer of a proto-oncogene may increase its
expression.
A point mutation in the coding sequence may lead to translation of a protein that is
more active or longer-lived.
Mutations to tumor-suppressor genes, whose normal products inhibit cell division, also
contribute to cancer.
Biology Chapter Notes
Any decrease in the normal activity of a tumor-suppressor protein may contribute to
cancer.
Some tumor-suppressor proteins normally repair damaged DNA, preventing the
accumulation of cancer-causing mutations.
Others control the adhesion of cells to each other or to an extracellular matrix, crucial
for normal tissues and often absent in cancers.
Still others are components of cell-signaling pathways that inhibit the cell cycle.
Oncogene proteins and faulty tumor-suppressor proteins interfere with normal signaling
pathways.
The proteins encoded by many proto-oncogenes and tumor-suppressor genes are
components of cell-signaling pathways.
Mutations in the products of two key genes, the ras proto-oncogene, and the p53 tumor
suppressor gene occur in 30% and 50% of human cancers, respectively.
Both the Ras protein and the p53 protein are components of signal-transduction pathways
that convey external signals to the DNA in the cell’s nucleus.
Ras, the product of the ras gene, is a G protein that relays a growth signal from a growth
factor receptor on the plasma membrane to a cascade of protein kinases.
At the end of the pathway is the synthesis of a protein that stimulates the cell cycle.
Many ras oncogenes have a point mutation that leads to a hyperactive version of the
Ras protein that can issue signals on its own, resulting in excessive cell division.
The p53 gene, named for its 53,000-dalton protein product, is often called the “guardian
angel of the genome.”
Damage to the cell’s DNA acts as a signal that leads to expression of the p53 gene.
The p53 protein is a transcription factor for several genes.
It can activate the p21 gene, which halts the cell cycle.
It can turn on genes involved in DNA repair.
When DNA damage is irreparable, the p53 protein can activate “suicide genes” whose
protein products cause cell death by apoptosis.
A mutation that knocks out the p53 gene can lead to excessive cell growth and cancer.
Multiple mutations underlie the development of cancer.
More than one somatic mutation is generally needed to produce the changes characteristic
of a full-fledged cancer cell.
If cancer results from an accumulation of mutations, and if mutations occur throughout
life, then the longer we live, the more likely we are to develop cancer.
Colorectal cancer, with 135,000 new cases and 60,000 deaths in the United States each
year, illustrates a multistep cancer path.
The first sign is often a polyp, a small benign growth in the colon lining.
The cells of the polyp look normal but divide unusually frequently.
Through gradual accumulation of mutations that activate oncogenes and knock out tumor-
suppressor genes, the polyp can develop into a malignant tumor.
About a half dozen DNA changes must occur for a cell to become fully cancerous.
These usually include the appearance of at least one active oncogene and the mutation or
loss of several tumor-suppressor genes.
Biology Chapter Notes
Since mutant tumor-suppressor alleles are usually recessive, mutations must knock out
both alleles.
Most oncogenes behave as dominant alleles and require only one mutation.
In many malignant tumors, the gene for telomerase is activated, removing a natural limit
on the number of times the cell can divide.
Viruses, especially retroviruses, play a role in about 15% of human cancer cases
worldwide.
These include some types of leukemia, liver cancer, and cancer of the cervix.
Viruses promote cancer development by integrating their DNA into that of infected cells.
By this process, a retrovirus may donate an oncogene to the cell.
Alternatively, insertion of viral DNA may disrupt a tumor-suppressor gene or convert a
proto-oncogene to an oncogene.
Some viruses produce proteins that inactivate p53 and other tumor-suppressor proteins,
making the cell more prone to becoming cancerous.
The fact that multiple genetic changes are required to produce a cancer cell helps explain
the predispositions to cancer that run in some families.
An individual inheriting an oncogene or a mutant allele of a tumor-suppressor gene
will be one step closer to accumulating the necessary mutations for cancer to develop.
Geneticists are devoting much effort to finding inherited cancer alleles so that
predisposition to certain cancers can be detected early in life.
About 15% of colorectal cancers involve inherited mutations, especially to DNA repair
genes or to the tumor-suppressor gene adenomatous polyposis coli, or APC.
Normal functions of the APC gene include regulation of cell migration and
adhesion.
Even in patients with no family history of the disease, APC is mutated in about 60%
of colorectal cancers.
Between 5–10% of breast cancer cases show an inherited predisposition.
This is the second most common type of cancer in the United States, striking more
than 180,000 women annually and leading to 40,000 annual deaths.
Mutations to one of two tumor-suppressor genes, BRCA1 and BRCA2, increase the
risk of breast and ovarian cancer.
A woman who inherits one mutant BRCA1 allele has a 60% probability of developing
breast cancer before age 50 (versus a 2% probability in an individual with two normal
alleles).
BRCA1 and BRCA2 are considered tumor-suppressor genes because their wild-type
alleles protect against breast cancer and because their mutant alleles are recessive.
Recent evidence suggests that the BRCA2 protein is directly involved in repairing
breaks that occur in both strands of DNA.
Concept 19.4 Eukaryotic genomes can have many noncoding DNA sequences in
addition to genes
Several trends are evident when we compare the genomes of prokaryotes to those of
eukaryotes.
There is a general trend from smaller to larger genomes, but with fewer genes in a given
length of DNA.
Biology Chapter Notes
Humans have 500 to 1,500 times as many base pairs in their genome as most
prokaryotes, but only 5 to 15 times as many genes.
Most of the DNA in a prokaryote genome codes for protein, tRNA, or rRNA.
The small amount of noncoding DNA consists mainly of regulatory sequences.
In eukaryotes, most of the DNA (98.5% in humans) does not code for protein or RNA.
Gene-related regulatory sequences and introns account for 24% of the human genome.
Introns account for most of the difference in average length of eukaryotic (27,000
base pairs) and prokaryotic genes (1,000 base pairs).
Most intergenic DNA is repetitive DNA, present in multiple copies in the genome.
Transposable elements and related sequences make up 44% of the entire human
genome.
The first evidence for transposable elements came from geneticist Barbara McClintock’s
breeding experiments with Indian corn (maize) in the 1940s and 1950s.
Eukaryotic transposable elements are of two types: transposons, which move within a
genome by means of a DNA intermediate, and retrotransposons, which move by means of
an RNA intermediate, a transcript of the retrotransposon DNA.
Transposons can move by a “cut and paste” mechanism, which removes the element
from its original site, or by a “copy and paste” mechanism, which leaves a copy
behind.
Retrotransposons always leave a copy at the original site, since they are initially
transcribed into an RNA intermediate.
Most transposons are retrotransposons, in which the transcribed RNA includes the code
for an enzyme that catalyzes the insertion of the retrotransposon and may include a gene
for reverse transcriptase.
Reverse transcriptase uses the RNA molecule originally transcribed from the
retrotransposon as a template to synthesize a double-stranded DNA copy.
Multiple copies of transposable elements and related sequences are scattered throughout
eukaryotic genomes.
A single unit is hundreds or thousands of base pairs long, and the dispersed “copies”
are similar but not identical to one another.
Some of the copies are transposable elements and some are related sequences that have
lost the ability to move.
Transposable elements and related sequences make up 25–50% of most mammalian
genomes, and an even higher percentage in amphibians and angiosperms.
In primates, a large portion of transposable element–related DNA consists of a family of
similar sequences called Alu elements.
These sequences account for approximately 10% of the human genome.
Alu elements are about 300 nucleotides long, shorter than most functional transposable
elements, and they do not code for protein.
Many Alu elements are transcribed into RNA molecules.
However, their cellular function is unknown.
Repetitive DNA that is not related to transposable elements probably arose by mistakes
that occurred during DNA replication or recombination.
Repetitive DNA accounts for about 15% of the human genome.
The different versions of each globin subunit are expressed at different times in
development, allowing hemoglobin to function effectively in the changing environment of
the developing animal.
embryonic, fetal, and/or adult stage of development.
In humans, the embryonic and fetal hemoglobins have higher affinity for oxygen than
do adult forms, ensuring transfer of oxygen from mother to developing fetus.
Also found in the globin gene family clusters are several pseudogenes, DNA
sequences similar to real genes that do not yield functional proteins.
One of the great achievements of modern science has been the sequencing of the human
genome, which was largely completed by 2003.
Progress began with the development of techniques for making recombinant DNA, in
which genes from two different sources—and often different species—are combined in
vitro into the same molecule.
The methods for making recombinant DNA are central to genetic engineering, the direct
manipulation of genes for practical purposes.
Applications include the introduction of a desired gene into the DNA of a host that will
produce the desired protein.
DNA technology has launched a revolution in biotechnology, the manipulation of
organisms or their components to make useful products.
Practices that go back centuries, such as the use of microbes to make wine and cheese
and the selective breeding of livestock, are examples of biotechnology.
These techniques exploit naturally occurring mutations and genetic recombination.
Biotechnology based on the manipulation of DNA in vitro differs from earlier
practices by enabling scientists to modify specific genes and move them between
organisms as distinct as bacteria, plants, and animals.
DNA technology is now applied in areas ranging from agriculture to criminal law, but its
most important achievements are in basic research.
Concept 20.1 DNA cloning permits production of multiple copies of a specific gene or
other DNA segment
To study a particular gene, scientists needed to develop methods to isolate the small, well-
defined portion of a chromosome containing the gene of interest.
Techniques for gene cloning enable scientists to prepare multiple identical copies of
gene-sized pieces of DNA.
One basic cloning technique begins with the insertion of a foreign gene into a bacterial
plasmid.
E. coli and its plasmids are commonly used.
First, a foreign gene is inserted into a bacterial plasmid to produce a recombinant DNA
molecule.
The plasmid is returned to a bacterial cell, producing a recombinant bacterium, which
reproduces to form a clone of identical cells.
Every time the bacterium reproduces, the recombinant plasmid is replicated as well.
Under suitable conditions, the bacterial clone will make the protein encoded by the
foreign gene.
The potential uses of cloned genes fall into two general categories.
First, the goal may be to produce a protein product.
Biology Chapter Notes
For example, bacteria carrying the gene for human growth hormone can produce
large quantities of the hormone.
Alternatively, the goal may be to prepare many copies of the gene itself.
This may enable scientists to determine the gene’s nucleotide sequence or provide
an organism with a new metabolic capability by transferring a gene from another
organism.
Most protein-coding genes exist in only one copy per genome, so the ability to clone
rare DNA fragments is very valuable.
Restriction enzymes are used to make recombinant DNA.
Gene cloning and genetic engineering were made possible by the discovery of restriction
enzymes that cut DNA molecules at specific locations.
In nature, bacteria use restriction enzymes to cut foreign DNA, to protect themselves
against phages or other bacteria.
They work by cutting up the foreign DNA, a process called restriction.
Most restriction enzymes are very specific, recognizing short DNA nucleotide sequences
and cutting at specific points in these sequences.
Bacteria protect their own DNA by methylating the sequences recognized by these
enzymes.
Each restriction enzyme cleaves a specific sequence of bases or restriction site.
These are often a symmetrical series of four to eight bases on both strands running in
opposite directions.
If the restriction site on one strand is 3’-CTTAAG-5’, the complementary strand is
5’-GAATTC-3’.
Because the target sequence usually occurs (by chance) many times on a long DNA
molecule, an enzyme will make many cuts.
Copies of a DNA molecule will always yield the same set of restriction fragments
when exposed to a specific enzyme.
Restriction enzymes cut covalent sugar-phosphate backbones of both strands, often in a
staggered way that creates single-stranded sticky ends.
These extensions can form hydrogen-bonded base pairs with complementary single-
stranded stretches (sticky ends) on other DNA molecules cut with the same restriction
enzyme.
These DNA fusions can be made permanent by DNA ligase, which seals the strand by
catalyzing the formation of covalent bonds to close up the sugar-phosphate backbone.
Restriction enzymes and DNA ligase can be used to make a stable recombinant DNA
molecule, with DNA that has been spliced together from two different organisms.
Eukaryotic genes can be cloned in bacterial plasmids.
Recombinant plasmids are produced by splicing restriction fragments from foreign DNA
into plasmids.
The original plasmid used to produce recombinant DNA is called a cloning vector,
defined as a DNA molecule that can carry foreign DNA into a cell and replicate there.
Bacterial plasmids are widely used as cloning vectors for several reasons.
They can be easily isolated from bacteria, manipulated to form recombinant plasmids
by in vitro insertion of foreign DNA, and then reintroduced into bacterial cells.
Concept 20.2 Restriction fragment analysis detects DNA differences that affect
restriction sites
Once we have prepared homogeneous samples of DNA, each containing a large number
of identical segments, we can begin to ask some interesting questions about specific genes
and their functions.
Does a particular gene differ from person to person?
Are certain alleles associated with a hereditary disorder?
Where in the body and when during development is a gene expressed?
What is the location of a gene in the genome?
Is expression of a particular gene related to expression of other genes?
How has a gene evolved, as revealed by interspecific comparisons?
To answer these questions, we need to know the nucleotide sequence of the gene and its
counterparts in other individuals and species, as well as its expression pattern.
One indirect method of rapidly analyzing and comparing genomes is gel electrophoresis.
Gel electrophoresis separates macromolecules—nucleic acids or proteins—on the basis
of their rate of movement through a gel in an electrical field.
Biology Chapter Notes
Rate of movement depends on size, electrical charge, and other physical properties
of the macromolecules.
In restriction fragment analysis, the DNA fragments produced by restriction enzyme
digestion of a DNA molecule are sorted by gel electrophoresis.
When the mixture of restriction fragments from a particular DNA molecule undergoes
electrophoresis, it yields a band pattern characteristic of the starting molecule and the
restriction enzyme used.
The relatively small DNA molecules of viruses and plasmids can be identified simply
by their restriction fragment patterns.
The separated fragments can be recovered undamaged from gels, providing pure
samples of individual fragments.
We can use restriction fragment analysis to compare two different DNA molecules
representing, for example, different alleles of a gene.
Because the two alleles differ slightly in DNA sequence, they may differ in one or
more restriction sites.
If they do differ in restriction sites, each will produce different-sized fragments when
digested by the same restriction enzyme.
In gel electrophoresis, the restriction fragments from the two alleles will produce
different band patterns, allowing us to distinguish the two alleles.
Restriction fragment analysis is sensitive enough to distinguish between two alleles of a
gene that differ by only one base pair in a restriction site.
A technique called Southern blotting combines gel electrophoresis with nucleic acid
hybridization.
Although electrophoresis will yield too many bands to distinguish individually, we can
use nucleic acid hybridization with a specific probe to label discrete bands that derive
from our gene of interest.
The probe is a radioactive single-stranded DNA molecule that is complementary to the
gene of interest.
Southern blotting reveals not only whether a particular sequence is present in the
sample of DNA, but also the size of the restriction fragments that contain the sequence.
One of its many applications is to identify heterozygous carriers of mutant alleles
associated with genetic disease.
In the example below, we compare genomic DNA samples from three individuals: an
individual who is homozygous for the normal ß-globin allele, a homozygote for sickle-cell
allele, and a heterozygote.
We combine several molecular techniques to compare DNA samples from three
individuals.
o We start by adding the same restriction enzyme to each of the three samples to produce
restriction fragments.
o We then separate the fragments by gel electrophoresis.
o We transfer the DNA fragments from the gel to a sheet of nitrocellulose paper, still
separated by size.
This also denatures the DNA fragments.
o Bathing the sheet in a solution containing a radioactively labeled probe allows the
probe to attach by base-pairing to the DNA sequence of interest.
o We can visualize bands containing the label with autoradiography.
Concept 20.5 The practical applications of DNA technology affect our lives in many
ways
DNA technology is reshaping medicine and the pharmaceutical industry.
Modern biotechnology is making enormous contributions both to the diagnosis of diseases
and in the development of pharmaceutical products.
The identification of genes whose mutations are responsible for genetic diseases may
lead to ways to diagnose, treat, or even prevent these conditions.
Susceptibility to many “nongenetic” diseases, from arthritis to AIDS, is influenced by
a person’s genes.
Diseases of all sorts involve changes in gene expression within the affected genes and
within the patient’s immune system.
DNA technology can identify these changes and lead to the development of targets for
prevention or therapy.
PCR and labeled nucleic acid probes can track down the pathogens responsible for
infectious diseases.
Biology Chapter Notes
For example, PCR can amplify and thus detect HIV DNA in blood and tissue samples,
detecting an otherwise elusive infection.
RNA cannot be directly amplified by PCR.
The RNA genome is first converted to double-stranded cDNA by a technique called
RT-PCR, using a probe specific for one of the HIV genes.
Medical scientists can use DNA technology to identify individuals with genetic diseases
before the onset of symptoms, even before birth.
Genetic disorders are diagnosed by using PCR and primers corresponding to cloned
disease genes, and then sequencing the amplified product to look for the disease-
causing mutation.
Cloned disease genes include those for sickle-cell disease, hemophilia, cystic
fibrosis, Huntington’s disease, and Duchenne muscular dystrophy.
It is even possible to identify symptomless carriers of these diseases.
It is possible to detect abnormal allelic forms of genes, even in cases in which the gene
has not yet been cloned.
The presence of an abnormal allele can be diagnosed with reasonable accuracy if a
closely linked RFLP marker has been found.
The closeness of the marker to the gene makes crossing over between them unlikely,
and the marker and gene will almost always be inherited together.
Techniques for gene manipulation hold great potential for treating disease by gene
therapy, the alteration of an afflicted individual’s genes.
A normal allele is inserted into somatic cells of a tissue affected by a genetic disorder.
For gene therapy of somatic cells to be permanent, the cells that receive the normal
allele must be ones that multiply throughout the patient’s life.
Bone marrow cells, which include the stem cells that give rise to blood and immune
system cells, are prime candidates for gene therapy.
A normal allele can be inserted by a retroviral vector into bone marrow cells removed
from the patient.
If the procedure succeeds, the returned modified cells will multiply throughout the
patient’s life and express the normal gene, providing missing proteins.
This procedure was used in a 2000 trial involving ten young children with SCID (severe
combined immunodeficiency disease), a genetic disease in which bone marrow cells do
not produce a vital enzyme because of a single defective gene.
Nine of the children showed significant improvement after two years.
However, two of the children developed leukemia.
It was discovered that the retroviral vector used to carry the normal allele into bone
marrow cells had inserted near a gene involved in proliferation and development of
blood cells, causing leukemia.
The trial has been suspended until researchers learn how to control the location of
insertion of the retroviral vectors.
Gene therapy poses many technical questions.
These include regulation of the activity of the transferred gene to produce the
appropriate amount of the gene product at the right time and place.
In addition, the insertion of the therapeutic gene must not harm other necessary cell
functions.
Gene therapy raises some difficult ethical and social questions.
The application of genetic analysis and DNA technology to the study of development has
brought about a revolution in our understanding of how a complex multicellular organism
develops from a single cell.
In 1995, Swiss researchers identified a gene that functions as a master switch to trigger
the development of the eye in Drosophila.
A similar gene triggers eye development in mammals.
Developmental biologists are discovering remarkable similarities in the mechanisms
that shape diverse organisms.
While geneticists were advancing from Mendel’s laws to an understanding of the
molecular basis of inheritance, developmental biologists were focusing on embryology.
Embryology is the study of the stages of development leading from fertilized egg to
fully formed organism.
In recent years, the concepts and tools of molecular genetics have reached a point where a
real synthesis of genetics and developmental biology has been possible.
When the primary research goal is to understand broad biological principles, the organism
chosen for study is called a model organism.
Researchers select model organisms that are representative of a larger group, suitable
for the questions under investigation, and easy to grow in the lab.
For study of the connections between genes and development, suitable model organisms
have short generation times and small genomes that are suitable for genetic analysis.
Model organisms used in developmental genetics include the fruit fly Drosophila
melanogaster, the nematode Caenorhabditis elegans, the mouse Mus musculus, the
zebra fish Danio rerio, and the plant Arabidopsis thaliana.
The fruit fly Drosophila melanogaster was first chosen as a model organism by geneticist
T. H. Morgan and intensively studied by generations of geneticists after him.
The fruit fly is small and easily grown in the laboratory.
It has a generation time of only two weeks and produces many offspring.
Embryos develop outside the mother’s body.
There are vast amounts of information on its genes and other aspects of its biology.
However, because first rounds of mitosis occur without cytokinesis, parts of its
development are superficially quite different from that of other organisms.
Sequencing of the Drosophila genome was completed in 2000.
It has 180 × 106 base pairs (180 Mb) and contains about 13,700 genes.
The nematode Caenorhabditis elegans normally lives in the soil but is easily grown in
petri dishes.
Only a millimeter long, it has a simple, transparent body with only a few cell types and
grows from zygote to mature adult in only three and a half days.
Its genome has been sequenced. It is 97 Mb long and contains an estimated 19,000
genes.
Because individuals are hermaphrodites, it is easy to detect recessive mutations.
Biology Chapter Notes
Self-fertilization of heterozygotes produces some homozygous recessive offspring
with mutant phenotypes.
Every adult C. elegans has exactly 959 somatic cells.
These arise from the zygote in virtually the same way for every individual.
By following all cell divisions with a microscope, biologists have constructed the
organism’s complete cell lineage, showing the ancestry of every cell in the adult
body.
The mouse Mus musculus has a long history as a mammalian model of development.
Much is known about its biology.
The mouse genome is about 2,600 Mb long with about 25,000 genes, about the same
as the human genome.
Researchers are adept at manipulating mouse genes to make transgenic mice and mice
in which particular genes are “knocked out” by mutation.
Mice are complex animals with a genome as large as ours.
Their embryos develop in the mother’s uterus, hidden from view.
A second vertebrate model, the zebra fish Danio rerio, has some unique advantages.
These small fish (2–4 cm long) are easy to breed in the laboratory in large numbers.
The transparent embryos develop outside the mother’s body.
Although generation time is two to four months, the early stages of development
proceed quickly.
By 24 hours after fertilization, most tissues and early versions of the organs have
formed.
After two days, the fish hatches out of the egg case.
The zebra fish genome is estimated to be 1,700 Mb, and is still being mapped and
sequenced.
For studying the molecular genetics of plant development, researchers are focusing on a
small weed, Arabidopsis thaliana (a member of the mustard family).
One plant can grow and produce thousands of progeny after eight to ten weeks.
A hermaphrodite, each flower makes eggs and sperm.
For gene manipulation research, scientists can induce cultured cells to take up foreign
DNA (genetic transformation).
Its relatively small genome, about 118 Mb, contains an estimated 25,500 genes.
In the development of most multicellular organisms, a single-celled zygote gives rise to
cells of many different types.
Each type has a different structure and corresponding function.
Cells of similar types are organized into tissues, tissues into organs, organs into organ
systems, and organ systems into the whole organism.
Thus, the process of embryonic development must give rise not only to cells of different
types, but also to higher-level structures arranged in a particular way in three dimensions.
Concept 21.1 Embryonic development involves cell division, cell differentiation, and
morphogenesis
An organism arises from a fertilized egg cell as the result of three interrelated processes:
cell division, cell differentiation, and morphogenesis.
Concept 21.2 Different cell types result from differential gene expression in cells with
the same DNA
During differentiation and morphogenesis, embryonic cells behave and function in ways
different from one another, even though all of them have arisen from the same zygote.
The differences between cells in a multicellular organism come almost entirely from
differences in gene expression, not differences in the cell’s genomes.
These differences arise during development, as regulatory mechanisms turn specific genes
off and on.
Different types of cells in an organism have the same DNA.
Much evidence supports the conclusion that nearly all the cells of an organism have
genomic equivalence—that is, they all have the same genes.
An important question that emerges is whether genes are irreversibly inactivated during
differentiation.
One experimental approach to the question of genomic equivalence is to try to generate a
whole organism from differentiated cells of a single type.
In many plants, whole new organisms can develop from differentiated somatic cells.
Biology Chapter Notes
During the 1950s, F. C. Steward and his students found that differentiated root cells
removed from the root could grow into normal adult plants when placed in a medium
culture.
These cloning experiments produced genetically identical individuals, popularly called
clones.
The fact that a mature plant cell can dedifferentiate (reverse its function) and give rise to
all the different kinds of specialized cells of a new plant shows that differentiation does
not necessarily involve irreversible changes in the DNA.
In plants, at least, cells can remain totipotent.
They retain the zygote’s potential to form all parts of the mature organism.
Plant cloning is now used extensively in agriculture.
Differentiated cells from animals often fail to divide in culture, much less develop into a
new organism.
Animal researchers have approached the genomic equivalence question by replacing the
nucleus of an unfertilized egg or zygote with the nucleus of a differentiated cell.
The pioneering experiments in nuclear transplantation were carried out by Robert
Briggs and Thomas King in the 1950s and extended later by John Gordon in the 1980s.
They destroyed or removed the nucleus of a frog egg and transplanted a nucleus from
an embryonic or tadpole cell from the same species into an enucleated egg.
The ability of the transplanted nucleus to support normal development is inversely related
to the donor’s age.
Transplanted nuclei from relatively undifferentiated cells from an early embryo lead to
the development of most eggs into tadpoles.
Transplanted nuclei from fully differentiated intestinal cells lead to fewer than 2% of
the cells developing into normal tadpoles.
Most of the embryos failed to make it through even the earliest stages of
development.
Developmental biologists agree on several conclusions about these results.
First, nuclei do change in some ways as cells differentiate.
While the DNA sequences do not change, histones may be modified or DNA may
be methylated.
In frogs and most other animals, nuclear “potency” tends to be restricted more and
more as embryonic development and cell differentiation progress.
However, chromatin changes are sometimes reversible, and the nuclei of most
differentiated animal cells probably have all the genes required for making an entire
organism.
The ability to clone mammals using nuclei or cells from early embryos has long been
possible.
In 1997, Scottish researchers announced the birth of Dolly, a lamb cloned from an adult
sheep by nuclear transplantation from a differentiated mammary cell.
The mammary cells were fused with sheep egg cells whose nuclei had been removed.
The resulting cells divided to form early embryos, which were implanted into surrogate
mothers.
One of several hundred implanted embryos completed normal development.
Concept 21.3 Pattern formation in animals and plants results from similar genetic and
cellular mechanisms
Before morphogenesis can shape an animal or plant, the organism’s body plan must be
established.
Cytoplasmic determinants and inductive signals contribute to pattern formation, the
development of spatial organization in which the tissues and organs of an organism are
all in their characteristic places.
Pattern formation continues throughout the life of a plant in the apical meristems.
In animals, pattern formation is mostly limited to embryos and juveniles.
Pattern formation begins in the early embryo, when the major axes of an animal and the
root-shoot axis of the plant are established.
The molecular cues that control pattern formation, positional information, tell a cell
its location relative to the body axes and to neighboring cells.
They also determine how the cells and their progeny will respond to future molecular
signals.
Drosophila development is controlled by a cascade of gene activations.
Pattern formation has been most extensively studied in Drosophila melanogaster, where
genetic approaches have had spectacular success.
These studies have established that genes control development and have identified the
key roles that specific molecules play in defining position and directing differentiation.
Combining anatomical, genetic, and biochemical approaches in the study of
Drosophila development, researchers have discovered developmental principles
common to many other species, including humans.
Fruit flies and other arthropods have a modular construction, an ordered series of
segments.
These segments make up the three major body parts: the head, thorax (with wings and
legs), and abdomen.
Biology Chapter Notes
Like other bilaterally symmetrical animals, Drosophila has an anterior-posterior axis
and a dorsal-ventral axis.
Cytoplasmic determinants in the unfertilized egg provide positional information for
the two developmental axes before fertilization.
After fertilization, positional information establishes a specific number of correctly
oriented segments and finally triggers the formation of each segment’s characteristic
structures.
The Drosophila egg cell develops in the female’s ovary, surrounded by ovarian cells
called nurse cells and follicle cells that supply the egg cell with nutrients, mRNAs, and
other substances needed for development.
Development of the fruit fly from egg cell to adult fly occurs in a series of discrete stages.
o Mitosis follows fertilization and egg laying.
Early mitosis occurs without growth of the cytoplasm and without cytokinesis,
producing one big multinucleate cell.
o At the tenth nuclear division, the nuclei begin to migrate to the periphery of the
embryo.
o At division 13, the cytoplasm partitions the 6,000 or so nuclei into separate cells.
The basic body plan—including body axes and segment boundaries—has already been
determined by this time.
A central yolk nourishes the embryo, and the eggshell continues to protect it.
o Subsequent events in the embryo create clearly visible segments, which at first look
very much alike.
o Some cells move to new positions, organs form, and a wormlike larva hatches from the
shell.
During three larval stages, the larva eats, grows, and molts.
o During the third larval stage, the larva transforms into the pupa enclosed in a case.
o Metamorphosis, the change from larva to adult fly, occurs in the pupal case, and the fly
emerges.
Each segment is anatomically distinct, with characteristic appendages.
The results of detailed anatomical observations of development in several species and
experimental manipulations of embryonic tissues laid the groundwork for understanding
the mechanisms of development.
In the 1940s, Edward B. Lewis demonstrated that the study of mutants could be used to
investigate Drosophila development.
He studied bizarre developmental mutations and located the mutations on the fly’s genetic
map.
This research provided the first concrete evidence that genes somehow direct the
developmental process.
In the late 1970s, Christiane Nüsslein-Volhard and Eric Weischaus pushed the
understanding of early pattern formation to the molecular level.
Their goal was to identify all the genes that affect segmentation in Drosophila, but they
faced three problems.
Because Drosophila has about 13,700 genes, there could be only a few genes affecting
segmentation or so many that the pattern would be impossible to discern.
Mutations that affect segmentation are likely to be embryonic lethals, leading to death
at the embryonic or larval stage.
Concept 21.4 Comparative studies help explain how the evolution of development leads
to morphological diversity
Biologists in the field of evolutionary developmental biology, or “evo-devo,” compare
developmental processes of different multicellular organisms.
Their aim is to understand how developmental processes have evolved and how
changes in the processes can modify existing organismal features or lead to new ones.
Biologists are finding that the genomes of related species with strikingly different
forms may have only minor differences in gene sequence or regulation.
All homeotic genes of Drosophila include a 180-nucleotide sequence called the
homeobox, which specifies a 60-amino-acid homeodomain.
An identical, or very similar, sequence of nucleotides (often called Hox genes) is found
in many other animals, including humans.
The vertebrate genes homologous to the homeotic genes of fruit flies have even kept
their chromosomal arrangement.
Related sequences have been found in the regulatory genes of plants, yeasts, and even
prokaryotes.
The homeobox DNA sequence must have evolved very early in the history of life and is
sufficiently valuable that it has been conserved virtually unchanged in animals and plants
for hundreds of millions of years.
Most, but not all, homeobox-containing genes are homeotic genes that are associated with
development.
For example, in Drosophila, homeoboxes are present not only in the homeotic genes,
but also in the egg-polarity gene bicoid, in several segmentation genes, and in the
master regulatory gene for eye development.
The homeobox-encoded homeodomain is part of a protein that binds to DNA when the
protein functions as a transcriptional regulator.
However, the shape of the homeodomain allows it to bind to any DNA segment.
Other, more variable, domains of the overall protein determine which genes it will
regulate.
Interaction of these latter domains with still other transcription factors helps a
homeodomain-protein recognize specific enhancers in the DNA.
Proteins with homeodomains probably regulate development by coordinating the
transcription of batteries of developmental genes.
Biology Chapter Notes
In Drosophila, different combinations of homeobox genes are active in different parts
of the embryo and at different times, leading to pattern formation.
Many other genes involved in development are highly conserved from species to species.
These include numerous genes encoding components of signaling pathways.
How can the same genes be involved in the development of so many different animals?
In some cases, small changes in regulatory sequences of particular genes can lead to
major changes in body form.
For example, varying expression of the Hox genes along the body axis produce
different numbers of leg-bearing segments in insects and crustaceans.
Plants also have homeobox-containing genes.
However, they do not appear to function as master regulatory switches in plants.
Other genes appear to be responsible for pattern formation in plants.
There are some basic similarities—and many differences—in the development of plants
and animals.
The last common ancestor of plants and animals was a single-celled microbe living
hundreds of millions of years ago, so the processes of development evolved independently
in the two lineages.
Plants have rigid cell walls that prevent cell movement, while morphogenetic
movements are very important in animals.
Morphogenesis in plants is dependent on differing planes of cell division and selective
cell enlargement.
Nevertheless, there are some basic similarities of development.
In both plants and animals, development relies on a cascade of transcriptional
regulators turning on or off genes in a finely tuned series.
The genes that direct these processes are very different in plants and animals.
Quite a few of the master regulatory switches in Drosophila are homeobox-containing
Hox genes.
Those in Arabidopsis belong to the Mads-box family of genes.
Although homeobox-containing genes can be found in plants and Mads-box genes can be
found in animals, they do not play the same major roles in development in plants and
animals.
The unity of life is reflected in the similarity of biological mechanisms used to establish
body pattern, although the exact genes directing develop may differ.
The similarities reflect the common ancestry of life on Earth, while the differences have
created the diversity of living organisms.
On November 24, 1859, Charles Darwin published On the Origin of Species by Means of
Natural Selection.
Darwin’s book drew a cohesive picture of life by connecting what had once seemed a
bewildering array of unrelated facts.
Darwin made two major points in The Origin of Species:
o Today’s organisms descended from ancestral species that were different from modern
species.
o Natural selection provided a mechanism for this evolutionary change.
The basic idea of natural selection is that a population can change over time if
individuals that possess certain heritable traits leave more offspring than other
individuals.
Natural selection results in evolutionary adaptation, an accumulation of inherited
characteristics that increase the ability of an organism to survive and reproduce in its
environment.
Eventually, a population may accumulate enough change that it constitutes a new species.
In modern terms, we can define evolution as a change over time in the genetic
composition of a population.
Evolution also refers to the gradual appearance of all biological diversity.
Evolution is such a fundamental concept that its study is relevant to biology at every level,
from molecules to ecosystems.
Evolutionary perspectives continue to transform medicine, agriculture, biotechnology,
and conservation biology.
Concept 22.1 The Darwinian revolution challenged traditional views of a young Earth
inhabited by unchanging species
Western culture resisted evolutionary views of life.
Darwin’s view of life contrasted with the traditional view of an Earth that was a few
thousand years old, populated by life forms that were created at the beginning and had
remained fundamentally unchanged.
The Origin of Species challenged a worldview that had been long accepted.
The Greek philosopher Aristotle (384–322 B.C.E.) opposed any concept of evolution and
viewed species as fixed and unchanging.
Aristotle believed that all living forms could be arranged on a ladder of increasing
complexity (scala naturae) with perfect, permanent species on every rung.
The Old Testament account of creation held that species were individually designed by
God and, therefore, perfect.
Biology Chapter Notes
In the 1700s, natural theology viewed the adaptations of organisms as evidence that the
Creator had designed each species for a purpose.
Carolus Linnaeus (1707–1778), a Swedish physician and botanist, founded taxonomy, a
system for naming species and classifying species into a hierarchy of increasingly
complex categories.
Linnaeus developed the binomial system of naming organisms according to genus and
species.
In contrast to the linear hierarchy of the scala naturae, Linnaeus adopted a nested
classification system, grouping similar species into increasingly general categories.
For Linnaeus, similarity between species did not imply evolutionary kinship but rather
the pattern of their creation.
Darwin’s views were influenced by fossils, remains or traces of organisms from the past
mineralized in sedimentary rocks.
Sedimentary rocks form when mud and sand settle to the bottom of seas, lakes, and
marshes.
New layers of sediment cover older ones, creating layers of rock called strata.
Erosion may later carve through sedimentary rock to expose older strata at the surface.
Fossils within layers of sedimentary rock show that a succession of organisms have
populated Earth throughout time.
Paleontology, the study of fossils, was largely developed by the French anatomist
Georges Cuvier (1769–1832).
In examining rock strata in the Paris Basin, Cuvier noted that the older the strata, the more
dissimilar the fossils from modern life.
Cuvier recognized that extinction had been a common occurrence in the history of life.
Instead of evolution, Cuvier advocated catastrophism, speculating that boundaries
between strata were due to local floods or droughts that destroyed the species then
present.
He suggested that the denuded areas were later repopulated by species immigrating
from unaffected areas.
Theories of geologic gradualism prepared the path for evolutionary biologists.
In contrast to Cuvier’s catastrophism, Scottish geologist James Hutton (1726–1797)
proposed a theory of gradualism that held that profound geological changes took place
through the cumulative effect of slow but continuous processes identical to those currently
operating.
Thus, valleys were formed by rivers flowing through rocks and sedimentary rocks
were formed from soil particles that eroded from land and were carried by rivers to the
sea.
Later, geologist Charles Lyell (1797–1875) proposed a theory of uniformitarianism,
which held that geological processes had not changed throughout Earth’s history.
Hutton’s and Lyell’s observations and theories had a strong influence on Darwin.
First, if geologic changes result from slow, continuous processes rather than sudden
events, then the Earth must be far older than the 6,000 years estimated by theologians
from biblical inference.
Second, slow and subtle processes persisting for long periods of time can also act on
living organisms, producing substantial change over a long period of time.
Lamarck placed fossils in an evolutionary context.
Biology Chapter Notes
In 1809, French biologist Jean-Baptiste de Lamarck (1744–1829) published a theory of
evolution based on his observations of fossil invertebrates in the collections of the Natural
History Museum of Paris.
By comparing fossils and current species, Lamarck found what appeared to be several
lines of descent.
Each was a chronological series of older to younger fossils, leading to a modern
species.
He explained his observations with two principles: use and disuse of parts and the
inheritance of acquired characteristics.
Use and disuse was the concept that body parts that are used extensively become larger
and stronger, while those that are not used deteriorate.
The inheritance of acquired characteristics stated that modifications acquired during
the life of an organism could be passed to offspring.
A classic example is the long neck of the giraffe. Lamarck reasoned that the long,
muscular neck of the modern giraffe evolved over many generations as the ancestors of
giraffes reached for leaves on higher branches and passed this characteristic to their
offspring.
Lamarck thought that evolutionary change was driven by the innate drive of organisms to
increasing complexity.
Lamarck’s theory was a visionary attempt to explain the fossil record and the current
diversity of life with recognition of gradual evolutionary change.
However, modern genetics has provided no evidence that acquired characteristics can
be inherited.
Acquired traits such as a body builder’s bigger biceps do not change the genes
transmitted through gametes to offspring.
Concept 22.2 In The Origin of Species, Darwin proposed that species change through
natural selection
Charles Darwin (1809–1882) was born in western England.
As a boy, he developed a consuming interest in nature.
When Darwin was 16, his father sent him to the University of Edinburgh to study
medicine.
Darwin left Edinburgh without a degree and enrolled at Cambridge University with the
intent of becoming a clergyman.
At that time, most naturalists and scientists belonged to the clergy and viewed the
world in the context of natural theology.
Darwin received his B.A. in 1831.
After graduation Darwin joined the survey ship HMS Beagle as ship naturalist and
conversation companion to Captain Robert FitzRoy.
FitzRoy chose Darwin because of his education, and because his age and social class
were similar to that of the captain.
Field research helped Darwin frame his view of life.
The primary mission of the five-year voyage of the Beagle was to chart poorly known
stretches of the South American coastline.
As you can see, the processes of meiosis and random fertilization have maintained the
same allele and genotype frequencies that existed in the previous generation.
The Hardy-Weinberg theorem states that the repeated shuffling of a population’s gene
pool over generations does not increase the frequency of one allele over another.
Theoretically, the allele frequencies in our flower population should remain at 0.8 for
CR and 0.2 for CW forever.
To generalize the example, in a population with two alleles with frequencies of p and q,
the combined frequencies must add to 100%.
Therefore p + q = 1.
If p + q = 1, then p = 1 − q and q = 1 − p.
In the wildflower W
example, p is the frequency of red alleles (CR) and q is the frequency of
white alleles (C ).
The probability of generating an CRCR offspring is p2 (an application of the rule of
multiplication).
In our example, p = 0.8 and p2 = 0.64.
The probability of generating a CWCW offspring is q2.
In our example, q = 0.2 and q2 = 0.04.
The probability of generating a CRCW offspring is 2pq.
In our example, 2 × 0.8 × 0.2 = 0.32.
The genotype frequencies must add up to 1.0:
p2 + 2pq + q2 = 1.0
For the wildflowers, 0.64 + 0.32 + 0.04 = 1.0.
This general formula is the Hardy-Weinberg equation.
Using this formula, we can calculate frequencies of alleles in a gene pool if we know the
frequency of genotypes, or the frequency of genotypes if we know the frequencies of
alleles.
Five conditions must be met for a population to remain in Hardy-Weinberg equilibrium.
The Hardy-Weinberg theorem describes a hypothetic population that is not evolving.
However, real populations do evolve, and their allele and genotype frequencies do change
over time.
That is because the five conditions for nonevolving populations are rarely met for long in
nature.
A population must satisfy five conditions if it is to remain in Hardy-Weinberg
equilibrium:
o Extremely large population size. In small populations, chance fluctuations in the gene
pool can cause genotype frequencies to change over time. These random changes are
called genetic drift.
o No gene flow. Gene flow, the transfer of alleles due to the migration of individuals or
gametes between populations, can change the proportions of alleles.
o No mutations. Introduction, loss, or modification of genes will alter the gene pool.
Biology Chapter Notes
o Random mating. If individuals pick mates with certain genotypes, or if inbreeding is
common, the mixing of gametes will not be random.
o No natural selection. Differential survival or reproductive success among genotypes
will alter their frequencies.
Evolution usually results when any of these five conditions are not met.
Although natural populations are rarely, if ever, in true Hardy-Weinberg equilibrium, the
rate of evolutionary change in many populations is so slow that they appear to be close to
equilibrium.
In such cases, we can use the Hardy-Weinberg equation to estimate genotype and
allele frequencies.
We can use the theorem to estimate the percentage of the human population that carries
the allele for the inherited disease phenylketonuria (PKU).
About 1 in 10,000 babies born in the United States is born with PKU, a metabolic
condition that results in mental retardation and other problems if left untreated.
The disease is caused by a recessive allele.
Is the U.S. population in Hardy-Weinberg equilibrium with respect to the PKU gene?
o The U.S. population is very large.
o Populations outside the United States have PKU allele frequencies similar to those
seen in the United States, so gene flow will not alter allele frequencies significantly.
o The mutation rate for the PKU gene is very low.
o People do not choose their partners based on whether or not they carry the PKU allele,
and inbreeding (marriage to close relatives) is rare in the United States.
o Selection against PKU only acts against the rare heterozygous recessive individuals.
From the epidemiological data, we know that frequency of homozygous recessive
individuals (q2 in the Hardy-Weinberg theorem) = 1 in 10,000, or 0.0001.
The frequency of the recessive allele (q) is the square root of 0.0001 = 0.01.
The frequency of the dominant allele (p) is p = 1 − q, or 1 − 0.01 = 0.99.
The frequency of carriers (heterozygous individuals) is 2pq = 2 × 0.99 × 0.01 =
0.0198, or about 2%.
Thus, about 2% of the U.S. population carries the PKU allele.
Concept 23.2 Mutation and sexual recombination produce the variation that makes
evolution possible
New genes and new alleles originate only by mutation.
A mutation is a change in the nucleotide sequence of an organism’s DNA.
Most mutations occur in somatic cells and are lost when the individual dies.
Only mutations in cell lines that form gametes can be passed on to offspring, and only a
small fraction of these spread through populations and become fixed.
A new mutation that is transmitted in a gamete to an offspring can immediately change
the gene pool of a population by introducing a new allele.
A point mutation is a change of a single base in a gene.
Biology Chapter Notes
Point mutations can have a significant impact on phenotype, as in the case of sickle-cell
disease.
However, most point mutations are harmless.
Much of the DNA in eukaryotic genomes does not code for protein products.
However, some noncoding regions of DNA do regulate gene expression.
Changes in these regulatory regions of DNA can have profound effects.
Because the genetic code is redundant, some point mutations in genes that code for
proteins may not alter the protein’s amino acid composition.
On rare occasions, a mutant allele may actually make its bearer better suited to the
environment, increasing reproductive success.
This is more likely when the environment is changing.
Some mutations alter gene number or sequence.
Chromosomal mutations that delete or rearrange many gene loci at once are almost
always harmful.
In rare cases, chromosomal rearrangements may be beneficial.
For example, the translocation of part of one chromosome to a different
chromosome could link genes that act together to positive effect.
Gene duplication is an important source of new genetic variation.
Small pieces of DNA can be introduced into the genome through the activity of
transposons.
Such duplicated segments can persist over generations and provide new loci that may
eventually take on new functions by mutation and subsequent selection.
New genes may also arise when the coding subsections of genes known as exons are
shuffled within the genome, within a single locus or between loci.
Such beneficial increases in gene number appear to have played a major role in evolution.
For example, mammalian ancestors carried a single gene for detecting odors that has been
duplicated though various mutational mechanisms.
Modern humans have close to 1,000 olfactory receptor genes.
60% of these genes have been inactivated in humans, due to mutations.
Mice, who rely more on their sense of smell, have lost only 20% of their olfactory
receptor genes.
Mutation rates vary from organism to organism.
Mutation rates are low in animals and plants, averaging about 1 mutation in every
100,000 genes per generation.
In microorganisms and viruses with short generation spans, mutation rates are much
higher and can rapidly generate genetic variation.
Sexual recombination also produces genetic variation.
On a generation-to-generation timescale, sexual recombination is far more important than
mutation in producing the genetic differences that make adaptation possible.
Sexual reproduction rearranges alleles into novel combinations every generation.
Bacteria and viruses can also undergo recombination, but they do so less regularly than
animals and plants.
Bacterial and viral recombination may cross species barriers.
Darwin visited the Galápagos Islands and found them filled with plants and animals
that lived nowhere else in the world.
He realized that he was observing newly emerged species on these young islands.
Speciation—the origin of new species—is at the focal point of evolutionary theory
because the appearance of new species is the source of biological diversity.
Microevolution is the study of adaptive change in a population.
Macroevolution addresses evolutionary changes above the species level.
It deals with questions such as the appearance of evolutionary novelties (e.g.,
feathers and flight in birds) that can be used to define higher taxa.
Speciation addresses the question of how new species originate and develop through
the subdivision and subsequent divergence of gene pools.
The fossil record chronicles two patterns of speciation: anagenesis and cladogenesis.
Anagenesis, phyletic evolution, is the accumulation of changes associated with the
gradual transformation of one species into another.
Cladogenesis, branching evolution, is the budding of one or more new species from a
parent species.
Only cladogenesis promotes biological diversity by increasing the number of
species.
Concept 24.1 The biological species concept emphasizes reproductive isolation
Species is a Latin word meaning “kind” or “appearance.”
Traditionally, morphological differences have been used to distinguish species.
Today, differences in body function, biochemistry, behavior, and genetic makeup
are also used to differentiate species.
Are organisms truly divided into the discrete units we called species, or is this
classification an arbitrary attempt to impose order on the natural world?
In 1942, Ernst Mayr proposed the biological species concept.
Concept 25.1 Phylogenies are based on common ancestries inferred from fossil,
morphological, and molecular evidence
Sedimentary rocks are the richest source of fossils.
Fossils are the preserved remnants or impressions left by organisms that lived in the past.
In essence, they are the historical documents of biology.
Sedimentary rocks form from layers of sand and silt that are carried by rivers to seas and
swamps, where the minerals settle to the bottom along with the remains of organisms.
As deposits pile up, they compress older sediments below them into layers called
strata.
The fossil record is the ordered array in which fossils appear within sedimentary rock
strata.
These rocks record the passing of geological time.
Fossils can be used to construct phylogenies only if we can determine their ages.
The fossil record is a substantial, but incomplete, chronicle of evolutionary change.
The majority of living things were not captured as fossils upon their death.
Of those that formed fossils, later geological processes destroyed many.
Only a fraction of existing fossils have been discovered.
The fossil record is biased in favor of species that existed for a long time, were
abundant and widespread, and had hard shells or skeletons that fossilized readily.
Morphological and molecular similarities may provide clues to phylogeny.
Biology Chapter Notes
Similarities due to shared ancestry are called homologies.
Organisms that share similar morphologies or DNA sequences are likely to be more
closely related than organisms without such similarities.
Morphological divergence between closely related species can be small or great.
Morphological diversity may be controlled by relatively few genetic differences.
Similarity due to convergent evolution is called analogy.
When two organisms from different evolutionary lineages experience similar
environmental pressures, natural selection may result in convergent evolution.
Similar analogous adaptations may evolve in such organisms.
Analogies are not due to shared ancestry.
Distinguishing homology from analogy is critical in the reconstruction of phylogeny.
For example, both birds and bats have adaptations that allow them to fly.
However, a close examination of a bat’s wing shows a greater similarity to a cat’s
forelimb that to a bird’s wing.
Fossil evidence also documents that bat and bird wings arose independently from
walking forelimbs of different ancestors.
Thus a bat’s wing is homologous to other mammalian forelimbs but is analogous in
function to a bird’s wing.
Analogous structures that have evolved independently are also called homoplasies.
In general, the more points of resemblance that two complex structures have, the less
likely it is that they evolved independently.
For example, the skulls of a human and a chimpanzee are formed by the fusion of
many bones.
The two skulls match almost perfectly, bone for bone.
It is highly unlikely that such complex structures have separate origins.
More likely, the genes involved in the development of both skulls were inherited from
a common ancestor.
The same argument applies to comparing genes, which are sequences of nucleotides.
Systematists compare long stretches of DNA and even entire genomes to assess
relationships between species.
If genes in two organisms have closely similar nucleotide sequences, it is highly likely
that the genes are homologous.
It may be difficult to carry out molecular comparisons of nucleic acids.
The first step is to align nucleic acid sequences from the two species being studied.
In closely related species, sequences may differ at only one or a few sites.
Distantly related species may have many differences or sequences of different length.
Over evolutionary time, insertions and deletions accumulate, altering the lengths of
the gene sequences.
Deletions or insertions may shift the remaining sequences, making it difficult to
recognize closely matching nucleotide sequences.
To deal with this, systematists use computer programs to analyze comparable DNA
sequences of differing lengths and align them appropriately.
The fact that molecules have diverged between species does not tell us how long ago their
common ancestor lived.
Biology Chapter Notes
Molecular divergences between lineages with reasonably complete fossil records can
serve as a molecular yardstick to measure the appropriate time span of various degrees
of divergence.
As with morphological characters, it is necessary to distinguish homology from analogy
to determine the usefulness of molecular similarities for reconstruction of phylogenies.
Closely similar sequences are most likely homologies.
In distantly related organisms, identical bases in otherwise different sequences may
simply be coincidental matches or molecular homoplasies.
Scientists have developed mathematical tools that can distinguish “distant” homologies
from coincidental matches in extremely divergent sequences.
For example, such molecular analysis has provided evidence that humans share a
distant common ancestor with bacteria.
Scientists have sequenced more than 20 billion bases worth of nucleic acid data from
thousands of species.
Life is a continuum extending from the earliest organisms to the great variety of forms
alive today.
Organisms interact with their environments.
Geological events that alter environments change the course of biological history.
When glaciers recede and the land rebounds, marine creatures can be trapped in
what gradually become freshwater lakes.
Populations of organisms trapped in these lakes are isolated from parent
populations, and may evolve into new species.
Life changes the planet it inhabits.
The evolution of photosynthetic organisms released oxygen into the air, with a
dramatic effect on Earth’s atmosphere.
The emergence of Homo sapiens has changed the land, water, and air at an
unprecedented rate.
Historical study of any sort is an inexact discipline that depends on the preservation,
reliability, and interpretation of records.
The fossil record of past life is generally less and less complete the further into the past
we delve.
Fortunately, each organism alive today carries traces of its evolutionary history in its
molecules, metabolism, and anatomy.
Still, the evolutionary episodes of greatest antiquity are generally the most obscure.
Concept 26.1 Conditions on early Earth made the origin of life possible
Most biologists now think that it is credible that chemical and physical processes on Earth
produced simple cells.
According to one hypothetical scenario, there were four main stages in this process:
o The abiotic synthesis of small organic molecules (monomers).
o The joining of monomers into polymers.
o The packaging of these molecules into protobionts, droplets with membranes that
maintained a distinct internal chemistry.
o The origin of self-replicating molecules that eventually made inheritance possible.
The scenario is speculative but does lead to predictions that can be tested in laboratory
experiments.
Earth and the other planets in the solar system formed about 4.6 billion years ago,
condensing from a vast cloud of dust and rocks surrounding the young sun.
It is unlikely that life could have originated or survived in the first few hundred million
years after the Earth’s formation.
Biology Chapter Notes
The planet was bombarded by huge bodies of rock and ice left over from the formation
of the solar system.
These collisions generated enough heat to vaporize all available water and prevent the
formation of the seas.
The oldest rocks on the Earth’s surface, located at a site called Isua in Greenland, are 3.8
billion years old.
It is not clear whether these rocks show traces of life.
The first cells may have originated by chemical evolution on a young Earth.
It is credible that chemical and physical processes on early Earth produced the first cells.
According to one hypothesis, there were four main stages to this process:
o Abiotic processes synthesized small organic molecules, such as amino acids and
nucleotides.
o These monomers were joined into polymers, including proteins and nucleic acids.
o Polymers were packaged into “protobionts,” droplets with membranes that maintained
an internal chemistry distinct from their surroundings.
o Self-replicating molecules arose, making inheritance possible.
Abiotic synthesis of organic monomers is a testable hypothesis.
As the bombardment of early Earth slowed, conditions on the planet were very different
from today.
The first atmosphere may have been a reducing atmosphere thick with water vapor,
along with nitrogen and its oxides, carbon dioxide, methane, ammonia, hydrogen, and
hydrogen sulfide.
Similar compounds are released from volcanic eruptions today.
As Earth cooled, the water vapor condensed into the oceans and much of the hydrogen
was lost into space.
In the 1920s, Russian chemist A. I. Oparin and British scientist J. B. S. Haldane
independently postulated that conditions on early Earth favored the synthesis of organic
compounds from inorganic precursors.
They reasoned that this could not happen today because high levels of oxygen in the
atmosphere attack chemical bonds.
A reducing environment in the early atmosphere would have promoted the joining of
simple molecules to form more complex ones.
The considerable energy required to make organic molecules could be provided by
lightning and the intense UV radiation that penetrated the primitive atmosphere.
Young suns emit more UV radiation. The lack of an ozone layer in the early
atmosphere would have allowed this radiation to reach Earth.
Haldane suggested that the early oceans were a solution of organic molecules, a
“primitive soup” from which life arose.
In 1953, Stanley Miller and Harold Urey tested the Oparin-Haldane hypothesis by
creating, in the laboratory, the conditions that had been postulated for early Earth.
They discharged sparks in an “atmosphere” of gases and water vapor.
The Miller-Urey experiments produced a variety of amino acids and other organic
molecules.
Other attempts to reproduce the Miller-Urey experiment with other gas mixtures have
also produced organic molecules, although in smaller quantities.
Biology Chapter Notes
It is unclear whether the atmosphere contained enough methane and ammonia to be
reducing.
There is growing evidence that the early atmosphere was made up primarily of
nitrogen and carbon dioxide.
Miller-Urey-type experiments with such atmospheres have not produced organic
molecules.
It is likely that small “pockets” of the early atmosphere near volcanic openings
were reducing.
Alternate sites proposed for the synthesis of organic molecules include submerged
volcanoes and deep-sea vents where hot water and minerals gush into the deep ocean.
These regions are rich in inorganic sulfur and iron compounds, which are important in
ATP synthesis by present-day organisms.
Some of the organic compounds from which the first life on Earth arose may have come
from space.
Researchers are looking outside of Earth for clues about the origin of life.
Evidence is growing that Mars was relatively warm for a brief period, with liquid
water and an atmosphere rich in carbon dioxide.
During that period, prebiotic chemistry similar to that on early Earth may have
occurred on Mars.
Did life evolve on Mars and then die out, or did dropping temperatures and a thinning
atmosphere terminate prebiotic chemistry before life evolved?
Liquid water lies beneath the ice-covered surface of Europa, one of Jupiter’s moons,
raising the possibility that Europa’s hidden ocean may harbor life.
Detection of free oxygen in the atmosphere of any planets outside our solar system
would be strongly suggestive of oxygenic photosynthesis.
Laboratory simulations of early-Earth conditions have produced organic polymers.
The abiotic origin hypothesis predicts that monomers should link to form polymers
without enzymes and other cellular equipment.
Researchers have produced polymers, including polypeptides, after dripping solutions of
monomers onto hot sand, clay, or rock.
Similar conditions likely existed on early Earth at deep-sea vents or when dilute
solutions of monomers splashed onto fresh lava.
Protobionts can form by self-assembly.
Life is defined by two properties: accurate replication and metabolism.
Neither property can exist without the other.
DNA molecules carry genetic information, including the information needed for accurate
replication.
The replication of DNA requires elaborate enzymatic machinery, along with a copious
supply of nucleotide building blocks provided by cell metabolism.
Although Miller-Urey experiments have yielded some of the nitrogenous bases of DNA
and RNA, they have not produced anything like nucleotides.
Thus, nucleotides were likely not part of the early organic soup.
Self-replicating molecules and a metabolism-like source of the building blocks must have
appeared together.
Concept 26.3 As prokaryotes evolved, they exploited and changed young Earth
The oldest known fossils are 3.5-billion-year-old stromatolites, rocklike structures
composed of layers of cyanobacteria and sediment.
Biology Chapter Notes
If bacterial communities existed 3.5 billion years ago, it seems reasonable that life
originated much earlier, perhaps 3.9 billion years ago, when Earth first cooled to a
temperature where liquid water could exist.
Prokaryotes dominated evolutionary history from 3.5 to 2.0 billion years ago.
The early protobionts must have used molecules present in the primitive soup for their
growth and replication.
Eventually, organisms that could produce all their needed compounds from molecules in
their environment replaced these protobionts.
A rich variety of autotrophs emerged, some of which could use light energy.
The diversification of autotrophs allowed the emergence of heterotrophs, which could live
on molecules produced by the autotrophs.
Prokaryotes were Earth’s sole inhabitants from 3.5 to 2.0 billion years ago.
These organisms transformed the biosphere of the planet.
Relatively early, prokaryotes diverged into two main evolutionary branches, the bacteria
and the archaea.
Representatives from both groups thrive in various environments today.
Metabolism evolved in prokaryotes.
The chemiosmotic mechanism of ATP synthesis is common to all three domains—
Bacteria, Archaea, and Eukarya.
This is evidence of a relatively early origin of chemiosmosis.
Transmembrane proton pumps may have functioned originally to expel H + that
accumulated when fermentation produced organic acids as waste products.
The cell would have to spend a large portion of its ATP to regulate internal pH by
driving H+ pumps.
The first electron +transport pumps may have coupled the oxidation of organic acids to
the transport of H out of the cell.
Finally, in some prokaryotes, electron transport systems efficient enough to expel more H+
than necessary to regulate pH evolved.
These cells could use the inward gradient of H + to reverse the H+ pump, which now
generated ATP instead of consuming it.
Such anaerobic respiration persists in some present-day prokaryotes.
Photosynthesis probably evolved very early in prokaryotic history.
The metabolism of early versions of photosynthesis did not split water and liberate
oxygen.
Some living prokaryotes display such nonoxygenic photosynthesis.
The only living photosynthetic prokaryotes that generate O2 are cyanobacteria.
Most atmospheric oxygen is of biological origin, from the water-splitting step of
photosynthesis.
When oxygenic photosynthesis first evolved, the free oxygen it produced likely
dissolved in the surrounding water until the seas and lakes became saturated with O2.
Additional O2 then reacted with dissolved iron to form the precipitate iron oxide.
These marine sediments were the source of banded iron formations, red layers of rock
containing iron oxide that are a valuable source of iron ore today.
Concept 26.4 Eukaryotic cells arose from symbioses and genetic exchanges between
prokaryotes
Eukaryotic cells differ in many respects from the smaller cells of bacteria and archaea.
Even the simplest single-celled eukaryote is far more complex in structure than any
prokaryote.
While there is some evidence of earlier eukaryotic fossils, the first clearly identified
eukaryote appeared about 2.1 billion years ago.
Other fossils that resemble simple, single-celled algae are slightly older (2.2 billion
years) but may not be eukaryotic.
Traces of molecules similar to cholesterol are found in rocks dating back 2.7 billion
years.
Such molecules are found only by aerobically respiring eukaryotic cells.
If confirmed, this would place the earliest eukaryotes at the same time as the
oxygen revolution that changed the Earth’s environment so dramatically.
Prokaryotes lack internal structures such as the nuclear envelope, endoplasmic reticulum,
and Golgi apparatus.
They have no cytoskeleton and are unable to change cell shape.
Eukaryotic cells have a cytoskeleton and can change shape, enabling them to surround
and engulf other cells.
The first eukaryotes may have been predators of other cells.
A cytoskeleton enables a eukaryotic cell to move structures within the cell and facilitates
the movement of chromosomes in meiosis and mitosis.
Mitosis made it possible to reproduce the large eukaryotic genome.
Meiosis allowed sexual recombination of genes.
How did the complex organization of the eukaryotic cell evolve from the simpler
prokaryotic condition?
A process called endosymbiosis probably led to mitochondria and plastids (the general
term for chloroplasts and related organelles).
The endosymbiotic theory suggests that mitochondria and plastids were formerly small
prokaryotes living within larger cells.
The term endosymbiont is used for a cell that lives within a host cell.
Biology Chapter Notes
The proposed ancestors of mitochondria were aerobic heterotrophic prokaryotes.
The proposed ancestors of plastids were photosynthetic prokaryotes.
The prokaryotic ancestors of mitochondria and plastids probably gained entry to the host
cell as undigested prey or internal parasites.
The symbiosis became mutually beneficial.
A heterotrophic host could use nutrients released from photosynthesis.
An anaerobic host would have benefited from an aerobic endosymbiont.
As they became increasingly interdependent, the host and endosymbionts became a single
organism.
All eukaryotes have mitochondria or their genetic remnants.
The theory of serial endosymbiosis supposes that mitochondria evolved before
plastids.
Overwhelming evidence supports an endosymbiotic origin of plastids and mitochondria.
The inner membranes of both organelles have enzymes and transport systems that are
homologous to those in the plasma membranes of modern prokaryotes.
Mitochondria and plastids replicate by a splitting process similar to prokaryotic binary
fission.
Like prokaryotes, each organelle has a single, circular DNA molecule that is not
associated with histone.
These organelles contain tRNAs, ribosomes, and other molecules needed to transcribe
and translate their DNA into protein.
Ribosomes of mitochondria and plastids are similar to prokaryotic ribosomes in terms
of size, nucleotide sequence, and sensitivity to antibiotics.
Which prokaryotic lineages gave rise to mitochondria and plastids?
Comparisons of small-subunit ribosomal RNA from mitochondria, plastids, and
various living prokaryotes suggest that a group of bacteria called the alpha
proteobacteria are the closest relatives to mitochondria and that cyanobacteria are the
closest relatives to plastids.
Over time, genes have been transferred from mitochondria and plastids to the nucleus.
This process may have been accomplished by transposable elements.
Some mitochondrial and plastic proteins are encoded by the organelle’s DNA, while
others are encoded by nuclear genes.
Some proteins are combinations of polypeptides encoded by genes in both locations.
The origins of other aspects of eukaryotic cells are unclear.
Some researchers have proposed that the nucleus itself evolved from an endosymbiont.
Nuclear genes with close relatives in both bacteria and archaea have been found.
The genome of eukaryotic cells may be the product of genetic annealing, in which
horizontal gene transfers occurred between many different bacterial and archaeal lineages.
These transfers may have taken place during the early evolution of life, or may have
happened repeatedly until the present day.
The origin of other eukaryotic structures is also the subject of active research.
The Golgi apparatus and the endoplasmic reticulum may have originated from
infoldings of the plasma membrane.
The cytoskeletal proteins actin and tubulin have been found in bacteria, where they are
involved in pinching off bacterial cells during cell division.
Biology Chapter Notes
These bacterial proteins may provide information about the origin of the eukaryotic
cytoskeleton.
Some investigators have suggested that eukaryotic flagella and cilia evolved from
symbiotic bacteria.
However, the 9+2 microtubule apparatus of eukaryotic flagella and cilia has not been
found in any prokaryotes.
Concept 26.6 New information has revised our understanding of the tree of life
In recent decades, molecular data have provided new insights into the evolutionary
relationships of life’s diverse forms.
The first taxonomic schemes divided organisms into plant and animal kingdoms.
In 1969, R. H. Whittaker argued for a five-kingdom system: Monera, Protista, Plantae,
Fungi, and Animalia.
The five-kingdom system recognized that there are two fundamentally different types
of cells: prokaryotic (the kingdom Monera) and eukaryotic (the other four kingdoms).
Three kingdoms of multicellular eukaryotes were distinguished by nutrition, in part.
Plants are autotrophic, making organic food by photosynthesis.
Most fungi are decomposers with extracellular digestion and absorptive nutrition.
Most animals ingest food and digest it within specialized cavities.
In Whittaker’s system, Protista included all eukaryotes that did not fit the definition of
plants, fungi, or animals.
Most protists are unicellular.
However, some multicellular organisms, such as seaweeds, were included in Protista
because of their relationships to specific unicellular protists.
The five-kingdom system prevailed in biology for more than 20 years.
During the past three decades, systematists applied cladistic analysis to taxonomy,
constructing cladograms based on molecular data.
These data led to the three-domain system of Bacteria, Archaea, and Eukarya as
“superkingdoms.”
Bacteria differ from Archaea in many key structural, biochemical, and physiological
characteristics.
Many microbiologists have divided the two prokaryotic domains into multiple kingdoms
based on cladistic analysis of molecular data.
A second challenge to the five-kingdom system comes from systematists who are sorting
out the phylogeny of the former members of the kingdom Protista.
Molecular systematics and cladistics have shown that the Protista is not monophyletic.
Some of these organisms have been split among five or more new kingdoms.
Others have been assigned to the Plantae, Fungi, or Animalia.
Clearly, taxonomy at the highest level is a work in progress.
Biology Chapter Notes
There will be much more research before there is anything close to a new consensus for
how the three domains of life are related and how many kingdoms should be included in
each domain.
New data, including the discovery of new groups, will lead to further taxonomic
remodeling.
Keep in mind that phylogenetic trees and taxonomic groupings are hypotheses that fit
the best available data.
Concept 27.2 A great diversity of nutritional and metabolic adaptations have evolved in
prokaryotes
Organisms can be categorized by their nutrition, based on how they obtain energy and
carbon to build the organic molecules that make up their cells.
Nutritional diversity is greater among prokaryotes than among all eukaryotes.
Every type of nutrition observed in eukaryotes is found in prokaryotes, along with some
nutritional modes unique to prokaryotes.
Organisms that obtain energy from light are phototrophs.
Organisms that obtain energy from chemicals in their environment are chemotrophs.
Organisms that need only CO2 as a carbon source are autotrophs.
Organisms that require at least one organic nutrient—such as glucose—as a carbon source
are heterotrophs.
Concept 27.5 Prokaryotes have both harmful and beneficial impacts on humans
Pathogenic prokaryotes represent only a small fraction of prokaryotic species.
Other prokaryotes serve as essential tools in agriculture and industry.
Prokaryotes cause about half of human diseases.
Between 2 and 3 million people a year die of the lung disease tuberculosis, caused by the
bacillus Mycobacterium tuberculosis.
Another 2 million die from diarrhea caused by other prokaryotes.
Lyme disease, caused by a bacterium carried by ticks that live on deer and field mice, is
the most widespread pest-carried disease in the United States.
Concept 28.8 Red algae and green algae are the closest relatives of land plants
More than a billion years ago, a heterotrophic protist acquired a cyanobacterial
endosymbiont.
The photosynthetic descendents of this ancient protist evolved into the red and green
algae.
At least 475 million years ago, the lineage that produced green algae gave rise to the land
plants.
Unlike other eukaryotic algae, red algae have no flagellated stages in their life cycle.
There are more than 6,000 known species of red algae, which are reddish due to the
accessory pigment phycoerythrin.
Coloration varies among species and depends on the depth that they inhabit.
Some species lack pigmentation and are parasites on other red algae.
Red algae are the most common seaweeds in the warm coastal waters of tropical oceans.
Red algae inhabit deeper waters than other photosynthetic eukaryotes.
Their photosynthetic pigments, especially phycobilins, allow them to absorb blue and
green wavelengths that penetrate down to deep water.
Biology Chapter Notes
One red algal species has been discovered off the Bahamas at a depth of more than 260
m.
Some red algae live in fresh water or on land.
Most red algae are multicellular, with some reaching a size large enough to be called
“seaweeds.”
The thalli of many red algal species are filamentous.
The base of the thallus is usually differentiated into a simple holdfast.
The life cycles of red algae are especially diverse.
In the absence of flagella, fertilization depends entirely on water currents to bring
gametes together.
Alternation of generations is common in red algae.
Green algae are named for their grass-green chloroplasts.
These are similar in ultrastructure and pigment composition to those of plants.
Molecular systematics and cellular morphology provide considerable evidence that green
algae and land plants are closely related.
In fact, some systematists advocate the inclusion of green algae into an expanded
“plant” kingdom, Viridiplantae.
Green algae are divided into two main groups, chlorophytes and charophyceans.
Most of the 7,000 species of chlorophytes have been identified.
Most live in fresh water, but many are marine inhabitants.
Some chlorophytes inhabit damp soil, while others are specialized to live on glaciers
and snowfields.
These snow-dwelling chlorophytes carry out photosynthesis despite subfreezing
temperatures and intense visible and ultraviolet radiation.
They are protected by radiation-blocking compounds in their cytoplasm and by the
snow itself, which acts as a shield.
Some chlorophytes live symbiotically with fungi to form lichens, a mutualistic
collective.
Large size and complexity in chlorophytes has evolved by three different mechanisms:
o Formation of colonies of individual cells (e.g., Volvox).
o The repeated division of nuclei without cytoplasmic division to form multinucleate
filaments (e.g., Caulerpa).
o The formation of true multicellular forms by cell division and cell differentiation (e.g.,
Ulva).
Some multicellular marine chlorophytes are large and complex enough to qualify as
seaweeds.
Most green algae have complex life cycles, with both sexual and asexual reproductive
stages.
Most sexual species have biflagellated gametes with cup-shaped chloroplasts.
Alternation of generations evolved in the life cycles of some green algae.
The other main group of green algae is most closely related to land plants.
For the first 3 billion years of Earth’s history, the land was lifeless.
Thin coatings of cyanobacteria existed on land about 1.2 billion years ago.
About 500 million years ago, plants, fungi, and animals joined them.
More than 290,000 species of plants inhabit Earth today.
Most plants live in terrestrial environments, including deserts, grasslands, and forests.
Some species, such as sea grasses, have returned to aquatic habitats.
The presence of plants has enabled other organisms to survive on land.
Plant roots have created habitats for other organisms by stabilizing landscapes.
Plants are the source of oxygen and the ultimate provider of food for land animals.
Concept 29.3 The life cycles of mosses and other bryophytes are dominated by the
gametophyte stage
Bryophytes are represented by three phyla:
Phylum Hepatophyta—liverworts
Phylum Anthocerophyta—hornworts
Phylum Bryophyta—mosses
Note that the name Bryophyta refers only to one phylum, but the informal term bryophyte
refers to all nonvascular plants.
It has not been established whether the diverse bryophytes form a clade.
Systematists continue to debate the sequence in which the three phyla of bryophytes
evolved.
Bryophytes acquired many unique adaptations after their evolutionary split from the
ancestors of modern vascular plants.
They also possess some ancestral traits characteristic of the earliest plants.
In bryophytes, gametophytes are the largest and most conspicuous phase of the life cycle.
Sporophytes are smaller and are present only part of the time.
Bryophyte spores germinate in favorable habitats and grow into gametophytes by mitosis.
Biology Chapter Notes
The gametophyte is a mass of green, branched, filaments that are one cell thick, called a
protonema.
A protonema has a large surface area that enhances absorption of water and minerals.
In favorable conditions, protonema generate gamete-producing structures, the
gametophores.
Bryophytes are anchored by tubular cells or filaments of cells, called rhizoids.
Unlike roots, rhizoids are not composed of tissues, lack specialized conducting cells,
and do not play a primary role in water and mineral absorption.
Bryophyte gametophytes are generally only one or a few cells thick, placing all cells close
to water and dissolved minerals.
Most bryophytes lack conducting tissues to distribute water and organic compounds
within the gametophyte.
Some mosses have conducting tissues in their stems, and a few can grow as tall as 2 m.
It is not clear if conducting tissues in mosses are analogous or homologous to the
xylem and phloem of vascular plants.
Lacking support tissues, most bryophytes are only a few centimeters tall.
The mature gametophores of bryophytes produce gametes in gametangia.
Each vase-shaped archegonium produces a single egg.
Elongated antheridia produce many flagellated sperm.
When plants are coated with a thin film of water, sperm swim toward the archegonia,
drawn by chemical attractants.
They swim into the archegonia and fertilize the eggs.
The zygotes and young sporophytes are retained and nourished by the parent
gametophyte.
Layers of placental nutritive cells transport materials from parent to embryos.
Bryophyte sporophytes disperse enormous numbers of spores.
While the bryophyte sporophyte does have photosynthetic plastids when young, it cannot
live apart from the maternal gametophyte.
A bryophyte sporophyte remains attached to its maternal gametophyte throughout the
sporophyte’s lifetime.
It depends on the gametophyte for sugars, amino acids, minerals, and water.
Bryophytes have the smallest and simplest sporophytes of all modern plant groups,
consistent with the hypothesis that larger and more complex sporophytes evolved only
later in vascular plants.
Moss sporophytes consist of a foot, an elongated stalk (the seta), and a sporangium
(the capsule).
The foot gathers nutrients and water from the parent gametophyte via transfer cells.
The stalk conducts these materials to the capsule.
In most mosses, the seta becomes elongated, elevating the capsule and enhancing spore
dispersal.
The moss capsule (sporangium) is the site of meiosis and spore production.
One capsule can generate more than 50 million spores.
When immature, the capsule is covered by a protective cap of gametophyte tissue, the
calyptra.
Biology Chapter Notes
This is lost when the capsule is ready to release spores.
The upper part of the capsule, the peristome, is often specialized for gradual spore
release.
Liverworts have the simplest sporophytes among the bryophytes.
They consist of a short stalk bearing round sporangia that contain the developing
spores, and a nutritive foot embedded in gametophyte tissues.
Hornwort and moss sporophytes are larger and more complex.
Hornwort sporophytes resemble grass blades and have a cuticle.
The sporophytes of mosses start out green and photosynthetic, but turn tan or brownish
red when ready to release their spores.
The sporophytes of hornworts and mosses have epidermal stomata, like those of
vascular plants.
These pores support photosynthesis by allowing the exchange of CO2 and O2
between the outside air and the interior of the sporophyte.
The fact that stomata are present in mosses and hornworts but absent in liverworts has
led to three hypotheses for their evolution.
o If liverworts are the deepest-branching lineage of land plants, then stomata evolved
once in the ancestor of hornworts, mosses and vascular plants.
o If hornworts are the deepest-branching lineage of land plants, then stomata evolved
once and were lost in the liverwort lineage.
o Perhaps hornworts acquired stomata independently of mosses and vascular plants.
Bryophytes provide many ecological and economic benefits.
Wind dispersal of lightweight spores has distributed bryophytes around the world.
They are common and diverse in moist forests and wetlands.
Some even inhabit extreme environments such as mountaintops, tundra, and deserts.
Phenolic compounds in moss cell walls absorb damaging levels of radiation present in
deserts and at high altitudes and latitudes.
Many mosses can exist in very cold or dry habitats because they are able to lose most of
their body water and then rehydrate and reactivate their cells when moisture again
becomes available.
Few vascular plants can survive the same degree of desiccation.
Sphagnum, a wetland moss, is especially abundant and widespread.
It forms extensive deposits of undecayed organic material, called peat.
Wet regions dominated by Sphagnum or peat moss are known as peat bogs.
Its organic materials do not decay readily because of resistant phenolic compounds and
acidic secretions that inhibit bacterial activity.
Peatlands, extensive high-latitude boreal wetlands occupied by Sphagnum, play an
important role as carbon reservoirs, stabilizing atmospheric carbon dioxide levels.
Sphagnum has been used in the past for diapers and as a natural antiseptic material for
wounds.
Today, it is harvested for use as a soil conditioner and for packing plants’ roots because of
the water storage capacity of its large, dead cells.
Worldwide, an estimated 400 billion tons of organic carbon are stored as peat.
Concept 30.1 The reduced gametophytes of seed plants are protected in ovules and
pollen grains
A number of terrestrial adaptations contributed to the success of seed plants.
These adaptations include the seed, the reduction of the gametophyte generation,
heterospory, ovules, and pollen.
Bryophyte life cycles are dominated by the gametophyte generation, while seedless
vascular plants have sporophyte-dominated life cycles.
The trend to gametophyte reduction continued in the lineage of vascular plants that led to
seed plants.
Seedless vascular plants have tiny gametophytes that are visible to the naked eye.
The gametophytes of seed plants are microscopically small and develop from spores
retained within the moist sporangia of the parental sporophyte.
In seed plants, the delicate female gametophyte and the young sporophyte embryo are
protected from many environmental stresses, including drought and UV radiation.
The gametophytes of seed plants obtain nutrients from their parents, while the free-
living gametophytes of seedless vascular plants must fend for themselves.
Heterospory is the rule among seed plants.
Nearly all seedless plants are homosporous, producing a single kind of spore that forms a
hermaphroditic gametophyte.
Seed plants likely had homosporous ancestors.
All seed plants are heterosporous, producing two different types of sporangia that produce
two types of spores.
Megasporangia produce megaspores, which give rise to female (egg-containing)
gametophytes.
Microsporangia produce microspores, which give rise to male (sperm-containing)
gametophytes.
Biology Chapter Notes
Seed plants produce ovules.
In contrast to the few species of heterosporous seedless vascular plants, seed plants are
unique in retaining their megaspores within the parent sporophyte.
Layers of sporophyte tissue, integuments, envelop and protect the megasporangium.
Gymnosperm megaspores are surrounded by one integument.
Angiosperm megaspores are surrounded by two integuments.
An ovule consists of the megasporangium, megaspores, and integuments.
A female gametophyte develops from a megaspore and produces one or more egg cells.
Pollen eliminated the liquid-water requirement for fertilization.
The microspores develop into pollen grains that are released from the microsporangium.
Pollen grains are covered with a tough coat containing sporopollenin.
They are carried by wind or animals.
The transfer of pollen to the vicinity of the ovule is called pollination.
The pollen grain germinates and grows as a pollen tube into the ovule, where it delivers
one or two sperm into the female gametophyte.
Bryophytes and seedless vascular plants have flagellated sperm cells that swim a few
centimeters through a film of water to reach the egg cells within the archegonium.
In seed plants, the female gametophyte is retained within the sporophyte ovule.
Male gametophytes travel long distances as pollen grains.
The sperm of seed plants lack flagella and do not require a film of water, as they rely
on the pollen tube to reach the egg cell of the female gametophyte within the ovule.
The sperm of some gymnosperm species retain the ancestral flagellated condition,
providing evidence of this evolutionary transition.
The evolution of pollen contributed to the success and diversity of seed plants.
Seeds became an important means of dispersing offspring.
What is a seed?
When a sperm fertilizes an egg of a seed plant, the zygote forms and develops into a
sporophyte embryo.
The ovule develops into a seed, consisting of the embryo and its food supply within a
protective coat derived from the integuments.
The evolution of the seed enabled plants to resist harsh environments and disperse
offspring more widely.
For bryophytes and seedless vascular plants, single-celled spores are the only protective
stage in the life cycle.
Moss spores can survive even if the local environment is too cold, too hot, or too dry
for the moss plants themselves to survive.
Because of their tiny size, the spores themselves can be dispersed in a dormant state to
a new area.
Spores were the main way that plants spread over Earth for the first 100 million years
of life on land.
The seed represents a different solution to resisting harsh environments and dispersing
offspring.
Concept 30.3 The reproductive adaptations of angiosperms include flowers and fruits
Angiosperms, commonly known as flowering plants, are vascular seed plants that produce
flowers and fruits.
They are the most diverse and geographically widespread of all plants, including more
than 90% of plant species.
There are about 250,000 known species of angiosperms.
All angiosperms are placed in a single phylum, the phylum Anthophyta.
The flower is the defining reproductive adaptation of angiosperms.
The flower is an angiosperm structure specialized for sexual reproduction.
In many species of angiosperms, insects and other animals transfer pollen from one
flower to female sex organs of another.
Some species that occur in dense populations, like grasses, are wind pollinated.
A flower is a specialized shoot with up to four circles of modified leaves: sepals, petals,
stamens, and carpals.
The sepals at the base of the flower are modified leaves that are usually green and enclose
the flower before it opens.
Biology Chapter Notes
The petals lie inside the ring of sepals.
These are often brightly colored in plant species that are pollinated by animals.
They typically lack bright coloration in wind-pollinated plant species.
Sepals and petals are sterile floral parts, not directly involved in reproduction.
Stamens, the male reproductive organs, are sporophylls that produce microspores that
will give rise to pollen grains containing male gametophytes.
A stamen consists of a stalk (the filament) and a terminal sac (the anther) where
pollen is produced.
Carpals are female sporophylls that produce megaspores and their products, female
gametophytes.
At the tip of the carpal is a sticky stigma that receives pollen.
A style leads to the ovary at the base of the carpal.
Ovules are protected within the ovary.
Fruits help disperse the seeds of angiosperms.
A fruit usually consists of a mature ovary.
As seeds develop from ovules after fertilization, the wall of the ovary thickens to form
the fruit.
Fruits protect dormant seeds and aid in their dispersal.
The fruit develops after pollination triggers hormonal changes that cause ovarian growth.
The wall of the ovary becomes the pericarp, the thickened wall of the fruit.
The other parts of the flower wither away in many plants.
If a flower has not been pollinated, the fruit usually does not develop, and the entire
flower withers and falls away.
Mature fruits can be fleshy or dry.
Oranges and grapes are fleshy fruits, in which one or more pericarp layers soften
during ripening.
Dry fruits include beans and grains.
The dry, wind-dispersed fruits of grasses are major food staples for humans.
The cereal grains of wheat, rice, and maize are fruits with a dry pericarp that adheres to
the seed coat of the seed.
Fruits are classified according to whether they develop from a single ovary, from multiple
ovaries, or from more than one flower.
By selectively breeding plants, humans have capitalized on the production of edible fruits.
Fruits are adapted to disperse seeds.
Winged seeds may function as kites or propellers to assist wind dispersal.
Coconuts are specialized for water dispersal.
Some fruits are modified as burrs that cling to animal fur.
Many fruits are edible, nutritious, sweet tasting, and colorful.
These fruits rely on animals to eat the fruit and deposit the seeds, along with a
supply of fertilizer, some distance from the parent plant.
The life cycle of an angiosperm is a highly refined version of the alternation of
generations common to all plants.
The honey mushroom Armillaria ostoyae in Malheur National Park in eastern Oregon is
enormous.
Its subterranean mycelium covers 890 hectares, weighs hundreds of tons, and has been
growing for 2,600 years.
Ten thousand species of fungi have been described, but it is estimated that there are
actually up to 1.5 million species of fungi.
Fungal spores have been found 160 km above the ground.
Fungi play an important role in ecosystems, decomposing dead organisms, fallen leaves,
feces, and other organic materials.
This decomposition recycles vital chemical elements back to the environment in forms
other organisms can assimilate.
Most plants depend on mutualistic fungi to help their roots absorb minerals and water
from the soil.
Humans have cultivated fungi for centuries for food, to produce antibiotics and other
drugs, to make bread rise, and to ferment beer and wine.
Concept 31.2 Fungi produce spores through sexual or asexual life cycles
Fungi reproduce by producing vast numbers of spores, either sexually or asexually.
The output of spores from one reproductive structure can be enormous.
Puffballs may release trillions of spores.
Dispersed widely by wind or water, spores germinate to produce mycelia if they land in a
moist place where there is food.
Many fungi have a heterokaryotic stage.
The nuclei of fungal hyphae and spores of most species are haploid, except for transient
diploid stages that form during sexual life cycles.
Sexual reproduction in fungi begins when hyphae from two genetically distinct mycelia
release sexual signaling molecules called pheromones.
Pheromones from each partner bind to receptors on the surface of the other.
The union of the cytoplasm of the two parent mycelia is known as plasmogamy.
In many fungi, the haploid nuclei do not fuse right away.
In some species, heterokaryotic mycelia become mosaics, with different nuclei
remaining in separate parts of the same mycelium or mingling and even exchanging
chromosomes and genes.
In some fungi, the haploid nuclei pair off two to a cell, one from each parent.
Such a mycelium is called dikaryotic, meaning “two nuclei.”
Biology Chapter Notes
In many fungi with sexual life cycles, karyogamy, fusion of haploid nuclei contributed by
two parents, occurs well after plasmogamy, cytoplasmic fusion of cells from the two
parents.
The delay may be hours, days, or even centuries.
During karyogamy, the haploid nuclei contributed by the two parents fuse, producing
diploid cells.
In most fungi, the zygotes of transient structures formed by karyogamy are the only
diploid stage in the life cycle.
These undergo meiosis to produce haploid cells that develop as spores in specialized
reproductive structures.
These spores disperse to form new haploid mycelia.
The sexual processes of karyogamy and meiosis generate genetic variation.
The heterokaryotic condition also offers some of the advantages of diploidy, in that one
haploid genome may be able to compensate for harmful mutations in the other.
Many fungi reproduce asexually.
The processes of asexual reproduction in fungi vary widely.
Some species reproduce only asexually.
Some fungi that can reproduce asexually grow as mold.
Molds grow rapidly as mycelia and produce spores.
Yeasts live in liquid or moist habitats.
Instead of producing spores, yeasts reproduce asexually by simple cell division or by
budding of small cells.
Most molds and yeasts have no known sexual stage.
Such fungi are called deuteromycetes, or imperfect fungi.
Whenever a sexual stage of a deuteromycete is discovered, the species is classified in a
particular phylum depending on its sexual structures.
Fungi can be identified from their sexual stages and by new genetic techniques.
Concept 31.5 Fungi have a powerful impact on ecosystems and human welfare
Ecosystems depend on fungi as decomposers and symbionts.
Biology Chapter Notes
Fungi are important decomposers of organic material, including cellulose and lignin of
plant cell walls.
Fungi and bacteria are essential for providing ecosystems with the inorganic nutrients
responsible for plant growth.
Without decomposers, carbon, nitrogen, and other elements would become tied up in
organic matter.
Fungi form symbiotic relationships with plants, algae, and animals.
Mycorrhizae are extremely important in natural ecosystems and agriculture.
Almost all vascular plants have mycorrhizae and rely on their fungal partners for
essential nutrients.
Some fungi break down plant material in the guts of cows and other grazers.
Many species of ants and termites raise fungi in “farms” and feed them leaves.
The fungi break the leaves down into a substance that the insects can digest.
Some mutualistic associations between “farmer” insects and “farmed” fungi have been
established for more than 50 million years.
In many cases, the fungi can no longer survive without the insects.
Lichens are a symbiotic association of millions of photosynthetic microorganisms held in
a mesh of fungal hyphae.
The fungal component is commonly an ascomycete, but several basidiomycete lichens are
known.
The photosynthetic partners are usually unicellular or filamentous green algae or
cyanobacteria.
The fungal hyphae provide most of the lichen’s mass and give it an overall shape and
structure.
The algae or cyanobacteria usually occupy an inner layer below the lichen surface.
The merger of fungus and algae is so complete that they are actually given genus and
species names, as though they were single organisms.
More than 13,500 species of lichen have been described—a fifth of all known fungi.
In most lichens, each partner provides things the other could not obtain on its own.
For example, the alga provides the fungus with food by “leaking” carbohydrate from
their cells.
The cyanobacteria provide organic nitrogen through nitrogen fixation.
The fungus provides a suitable physical environment for growth, retaining water and
minerals, allowing for gas exchange, shading the algae or cyanobacteria from intense
sunlight with pigments, and deterring consumers with toxic compounds.
The fungus also secretes acids, which aids in the uptake of minerals.
The fungi of many lichens reproduce sexually by forming ascocarps or basidiocarps.
Lichen algae reproduce independently by asexual cell division.
Asexual reproduction of symbiotic units occurs either by fragmentation of the parental
lichen or by the formation of structures called soredia, small clusters of hyphae with
embedded algae.
Phylogenetic studies of lichen DNA have helped illuminate the evolution of this
symbiosis.
Concept 32.1 Animals are multicellular, heterotrophic eukaryotes with tissues that
develop from embryonic layers
There are exceptions to nearly every criterion for distinguishing an animal from other life
forms.
However, five criteria, taken together, comprise a reasonable definition.
o Animals are multicellular, ingestive heterotrophs.
Animals take in preformed organic molecules through ingestion, eating other
organisms or organic material that is decomposing.
o Animal cells lack cell walls that provide structural support for plants and fungi.
The multicellular bodies of animals are held together by extracellular structural
proteins, especially collagen.
Animals have other unique types of intercellular junctions, including tight junctions,
desmosomes, and gap junctions, which hold tissues together.
These junctions are also composed of structural proteins.
o Animals have two unique types of cells: nerve cells for impulse conduction and muscle
cells for movement.
o Most animals reproduce sexually, with the diploid stage usually dominating the life
cycle.
In most species, a small flagellated sperm fertilizes a larger, nonmotile egg.
The zygote undergoes cleavage, a succession of mitotic cell divisions, leading to the
formation of a multicellular, hollow ball of cells called the blastula.
During gastrulation, part of the embryo folds inward, forming layers of embryonic
tissues that will develop into adult body parts.
The resulting development stage is called a gastrula.
Some animals develop directly through transient stages into adults, but others have a
distinct larval stage or stages.
A larva is a sexually immature stage that is morphologically distinct from the adult,
usually eats different foods, and may live in a different habitat from the adult.
Animal larvae eventually undergo metamorphosis, transforming the animal into an
adult.
o Animals share a unique homeobox-containing family of genes known as Hox genes.
All eukaryotes have genes that regulate the expression of other genes.
Many of these regulatory genes contain common modules of DNA sequences called
homeoboxes.
Biology Chapter Notes
All animals share the unique family of Hox genes, suggesting that this gene family
arose in the eukaryotic lineage that gave rise to animals.
Hox genes play important roles in the development of animal embryos, regulating the
expression of dozens or hundreds of other genes.
Hox genes control cell division and differentiation, producing different
morphological features of animals.
Hox genes in sponges regulate the formation of channels, the primary feature of
sponge morphology.
In more complex animals, the Hox gene family underwent further duplication.
In bilaterians, Hox genes regulate patterning of the anterior-posterior axis.
The same conserved genetic network governs the development of a large range of
animals.
Concept 32.2 The history of animals may span more than a billion years
Various studies suggest that animals began to diversify more than a billion years ago.
Some calculations based on molecular clocks estimate that the ancestors of animals
diverged from the ancestors of fungi as much as 1.5 billion years ago.
Similar studies suggest that the common ancestor of living animals lived 1.2 billion to 800
million years ago.
The common ancestor was probably a colonial flagellated protist and may have resembled
modern choanoflagellates.
Neoproterozoic Era (1 billion–542 million years ago)
Although molecular data indicates a much earlier origin of animals, the oldest generally
accepted animal fossils are only 575 million years old.
These fossils are known as the Ediacara fauna, named for the Ediacara Hills of
Australia.
Ediacara fauna consist primarily of cnidarians, but soft-bodied mollusks were also
present, and numerous fossilized burrows and tracks indicate the presence of worms.
Paleozoic Era (542–251 million years ago)
Animals underwent considerable diversification between 542–525 million years ago,
during the Cambrian period of the Paleozoic Era.
During this period, known as the Cambrian explosion, about half of extant animal
phyla arose.
Fossils of Cambrian animals include the first animals with hard, mineralized skeletons.
There are several hypotheses regarding the cause of the Cambrian explosion.
o The new predator-prey relationships that emerged in the Cambrian may have generated
diversity through natural selection.
Predators acquired adaptations that helped them catch prey.
Prey acquired adaptations that helped them resist predation.
o A rise of atmospheric oxygen preceded the Cambrian explosion.
More oxygen may have provided opportunities for animals with higher metabolic rates
and larger body sizes.
o The evolution of the Hox complex provided the developmental flexibility that resulted
in variations in morphology.
Biology Chapter Notes
These hypotheses are not mutually exclusive; all may have played a role.
In the Silurian and Devonian periods, animal diversity continued to increase, punctuated
by episodes of mass extinction.
Vertebrates (fishes) became the top predators of marine food webs.
By 460 million years ago, arthropods began to adapt to terrestrial habitats.
Vertebrates moved to land about 360 million years ago and diversified into many
lineages.
Two of these survive today: amphibians and amniotes.
Mesozoic Era (251–65.5 million years ago)
Few new animal body plans emerged among animals during the Mesozoic era.
Animal phyla began to spread into new ecological niches.
In the oceans, the first coral reefs formed.
On land, birds, pterosaurs, dinosaurs, and tiny nocturnal insect-eating mammals arose.
Cenozoic Era (65.5 million years ago to the present)
Insects and flowering plants both underwent a dramatic diversification during the
Cenozoic era.
This era began with mass extinctions of terrestrial and marine animals.
Among the groups of species that disappeared were large, nonflying dinosaurs and the
marine reptiles.
Large mammalian herbivores and carnivores diversified as mammals exploited vacated
ecological niches.
Some primate species in Africa adapted to open woodlands and savannas as global
climates cooled.
Our ancestors were among these grassland apes.
Concept 32.4 Leading hypotheses agree on major features of the animal phylogenetic
tree
Zoologists currently recognize about 35 animal phyla.
The relationships between these phyla continue to be debated.
Traditionally, zoologists have tested hypotheses about animal phylogeny through
morphological studies.
Currently, zoologists also study the molecular systematics of animals.
New studies of lesser-known phyla and fossil analyses help distinguish between ancestral
and derived traits in various animal groups.
Modern phylogenetic systematics is based on the identification of clades, monophyletic
sets of taxa defined by shared derived features unique to those taxa and their common
ancestor.
This creates a phylogenetic tree that is a hierarchy of clades nested within larger
clades.
Defining the shared derived characteristics is key to a particular hypothesis.
Whether the data are “traditional” morphological characters, “new” molecular
sequences, or some combination of the two, the assumptions and inferences inherent in
the tree are the same.
Two current phylogenetic hypotheses can be compared: one based on systematic analyses
of morphological characters and the other based on recent molecular studies.
The hypotheses agree on the following major features of animal phylogeny.
o All animals share a common ancestor.
Concept 33.1 Sponges are sessile and have a porous body and choanocytes
Sponges (phylum Porifera) are so sedentary that they were mistaken for plants by the
early Greeks.
Living in freshwater and marine environments, sponges are suspension feeders.
The body of a simple sponge resembles a sac perforated with holes.
Water is drawn through the pores into a central cavity, the spongocoel, and flows out
through a larger opening, the osculum.
More complex sponges have folded body walls, and many contain branched water
canals and several oscula.
Sponges range in height from about a few mm to 2 m and most are marine.
About 100 species live in fresh water.
Unlike eumetazoa, sponges lack true issues, groups of similar cells that form a functional
unit.
The germ layers of sponges are loose federations of cells, which are not really tissues
because the cells are relatively unspecialized.
The sponge body does contain different cell types.
Sponges collect food particles from water passing through food-trapping equipment.
Flagellated choanocytes, or collar cells, lining the spongocoel (internal water
chambers) create a flow of water through the sponge with their flagella and trap food
with their collars.
Based on both molecular evidence and the morphology of their choanocytes, sponges
evolved from a colonial choanoflagellate ancestor.
The body of a sponge consists of two cell layers separated by a gelatinous region, the
mesohyl.
Wandering though the mesohyl are amoebocytes.
They take up food from water and from choanocytes, digest it, and carry nutrients to
other cells.
They also secrete tough skeletal fibers within the mesohyl.
In some groups of sponges, these fibers are sharp spicules of calcium carbonate or
silica.
Biology Chapter Notes
Other sponges produce more flexible fibers from a collagen protein called spongin.
We use these pliant, honeycombed skeletons as bath sponges.
Most sponges are sequential hermaphrodites, with each individual producing both sperm
and eggs in sequence.
Gametes arise from choanocytes or amoebocytes.
The eggs are retained, but sperm are carried out the osculum by the water current.
Sperm are drawn into neighboring individuals and fertilize eggs in the mesohyl.
The zygotes develop into flagellated, swimming larvae that disperse from the parent.
When a larva finds a suitable substratum, it develops into a sessile adult.
Sponges produce a variety of antibiotics and other defensive compounds.
Researchers are now isolating these compounds, which may be useful in fighting
human disease.
Concept 33.2 Cnidarians have radial symmetry, a gastrovascular cavity, and cnidocytes
All animals except sponges belong to the Eumetazoa, the animals with true tissues.
The cnidarians (hydras, jellies, sea anemones, and coral animals) have a relatively simple
body construction.
They are a diverse group with more than 10,000 living species, most of which are
marine.
They exhibit a relatively simple, diploblastic body plan that arose 570 million years
ago.
The basic cnidarian body plan is a sac with a central digestive compartment, the
gastrovascular cavity.
A single opening to this cavity functions as both mouth and anus.
This basic body plan has two variations: the sessile polyp and the floating medusa.
The cylindrical polyps, such as hydras and sea anemones, adhere to the substratum by the
aboral end and extend their tentacles, waiting for prey.
Medusas (also called jellies) are flattened, mouth-down versions of polyps that move by
drifting passively and by contracting their bell-shaped bodies.
The tentacles of a jelly dangle from the oral surface.
Some cnidarians exist only as polyps.
Others exist only as medusas.
Still others pass sequentially through both a medusa stage and a polyp stage in their
life cycle.
Cnidarians are carnivores that use tentacles arranged in a ring around the mouth to capture
prey and push the food into the gastrovascular chamber for digestion.
Batteries of cnidocytes on the tentacles defend the animal or capture prey.
Organelles called cnidae evert a thread that can inject poison into the prey, or stick
to or entangle the target.
Cnidae called nematocysts are stinging capsules.
Muscles and nerves exist in their simplest forms in cnidarians.
Cells of the epidermis and gastrodermis have bundles of microfilaments arranged into
contractile fibers.
Biology Chapter Notes
True muscle tissue appears first in triploblastic animals.
When the animal closes its mouth, the gastrovascular cavity acts as a hydrostatic
skeleton against which the contractile cells can work.
Movements are controlled by a noncentralized nerve net associated with simple sensory
receptors that are distributed radially around the body.
The phylum Cnidaria is divided into four major classes: Hydrozoa, Scyphozoa, Cubozoa,
and Anthozoa.
The four cnidarian classes show variations on the same body theme of polyp and medusa.
Most hydrozoans alternate polyp and medusa forms, as in the life cycle of Obelia.
The polyp stage, often a colony of interconnected polyps, is more conspicuous than the
medusa.
Hydras, among the few freshwater cnidarians, are unusual members of the class Hydrozoa
in that they exist only in the polyp form.
When environmental conditions are favorable, a hydra reproduces asexually by
budding, the formation of outgrowths that pinch off from the parent to live
independently.
When environmental conditions deteriorate, hydras form resistant zygotes that remain
dormant until conditions improve.
The medusa generally prevails in the life cycle of class Scyphozoa.
The medusae of most species live among the plankton as jellies.
Most coastal scyphozoans go through small polyp stages during their life cycle.
Jellies that live in the open ocean generally lack the sessile polyp.
Cubozoans have a box-shaped medusa stage.
They can be distinguished from scyphozoans in other significant ways, such as having
complex eyes in the fringe of the medusae.
Cubozoans, which generally live in tropical oceans, are often equipped with highly toxic
cnidocytes.
Sea anemones and corals belong to the class Anthozoa.
They occur only as polyps.
Coral animals live as solitary or colonial forms and secrete a hard external skeleton of
calcium carbonate.
Each polyp generation builds on the skeletal remains of earlier generations to form
skeletons that we call coral.
In tropical seas, coral reefs provide habitat for a great diversity of invertebrates and fishes.
Coral reefs in many parts of the world are currently being destroyed by human activity.
Pollution, overfishing, and global warming are contributing to their demise.
Concept 33.4 Molluscs have a muscular foot, a visceral mass, and a mantle
The phylum Mollusca includes many diverse forms, including snails and slugs, oysters
and clams, and octopuses and squids.
Most molluscs are marine, though some inhabit fresh water, and some snails and slugs
live on land.
Molluscs are soft-bodied animals, but most are protected by a hard shell of calcium
carbonate.
Slugs, squids, and octopuses have reduced or lost their shells completely during their
evolution.
Despite their apparent differences, all molluscs have a similar body plan with a muscular
foot (typically for locomotion), a visceral mass with most of the internal organs, and a
mantle.
The mantle, which secretes the shell, drapes over the visceral mass and creates a water-
filled chamber, the mantle cavity, with gills, anus, and excretory pores.
Biology Chapter Notes
Many molluscs feed by using a straplike rasping organ, a radula, to scrape up food.
Most molluscs have separate sexes, with gonads located in the visceral mass.
However, many snails are hermaphrodites.
The life cycle of many marine molluscs includes a ciliated larva, the trochophore.
This larva is also found in marine annelids (segmented worms) and some other
lophotrochozoans.
The basic molluscan body plan has evolved in various ways in the eight classes of the
phylum.
The four most prominent are the Polyplacophora (chitons), Gastropoda (snails and
slugs), Bivalvia (clams, oysters, and other bivalves), and Cephalopoda (squids,
octopuses, cuttlefish, and chambered nautiluses).
Chitons are marine animals with oval shapes and shells divided into eight dorsal plates.
The chiton body is unsegmented.
Chitons use their muscular foot to grip the rocky substrate tightly and to creep slowly over
the rock surface.
Chitons are grazers that use their radulas to scrape and ingest algae.
Almost three-quarters of all living species of molluscs are gastropods.
Most gastropods are marine, but there are also many freshwater species.
Garden snails and slugs have adapted to land.
During embryonic development, gastropods undergo torsion in which the visceral mass is
rotated up to 180 degrees, so the anus and mantle cavity are above the head in adults.
After torsion, some of the organs that were bilateral are reduced or lost on one side of
the body.
Most gastropods are protected by single, spiral shells into which the animals can retreat if
threatened.
Torsion and formation of the coiled shell are independent developmental processes.
While gastropod shells are typically conical, those of abalones and limpets are somewhat
flattened.
Many gastropods have distinct heads with eyes at the tips of tentacles.
They move by a rippling motion of their foot or by means of cilia.
Most gastropods use their radula to graze on algae or plant material.
Some species are predators.
In these species, the radula is modified to bore holes in the shells of other organisms or
to tear apart tough animal tissues.
In the tropical marine cone snails, teeth on the radula form separate poison darts,
which penetrate and stun their prey, including fishes.
In place of the gills found in most aquatic gastropods, the lining of the mantle cavity of
terrestrial snails functions as a lung.
The class Bivalvia includes clams, oysters, mussels, and scallops.
Bivalves have shells divided into two halves.
The two parts are hinged at the mid-dorsal line, and powerful adductor muscles close
the shell tightly to protect the animal.
Bivalves have no distinct head, and the radula has been lost.
Biology Chapter Notes
Some bivalves have eyes and sensory tentacles along the outer edge of the mantle.
The mantle cavity of a bivalve contains gills that are used for feeding and gas exchange.
Most bivalves are suspension feeders, trapping fine particles in mucus that coats the gills.
Cilia convey the particles to the mouth.
Water flows into the mantle cavity via the incurrent siphon, passes over the gills, and
exits via the excurrent siphon.
Most bivalves live rather sedentary lives, a characteristic suited to suspension feeding.
Sessile mussels secrete strong threads that tether them to rocks, docks, boats, and the
shells of other animals.
Clams can pull themselves into the sand or mud, using the muscular foot as an anchor.
Scallops can swim in short bursts to avoid predators by flapping their shells and jetting
water out their mantle cavity.
Cephalopods are active predators that use rapid movements to dart toward their prey,
which they capture with several long tentacles.
Squids and octopuses use beak-like jaws to bite their prey and then inject poison to
immobilize the victim.
A mantle covers the visceral mass, but the shell is reduced and internal in squids, missing
in many octopuses, and exists externally only in chambered nautiluses.
Fast movements by a squid occur when it contracts its mantle cavity and fires a stream of
water through the excurrent siphon.
By pointing the siphon in different directions, the squid can rapidly move in different
directions.
The foot of a cephalopod has been modified into the muscular siphon and parts of the
tentacles and head.
Cephalopods are the only molluscs with a closed circulatory system.
They also have well-developed sense organs and a complex brain.
The ancestors of octopuses and squid were probably shelled molluscs that took up a
predatory lifestyle.
Shelled cephalopods called ammonites were the dominant invertebrate predators of the
seas for hundreds of millions of years until their disappearance in the mass extinctions at
the end of the Cretaceous period.
Most squid are less than 75 cm long.
In 2003, a squid with a mantle 2.5 meters long was captured near Antarctica.
The specimen was possibly a juvenile, only half the size of an adult.
Large squid are thought to feed on large fish in the deep ocean, where sperm whales
are their only natural predators.
Concept 33.7 Arthropods are segmented coelomates that have an exoskeleton and
jointed appendages
The world arthropod population has been estimated at a billion billion (10 18) individuals.
Nearly a million arthropod species have been described.
Two out of every three known species are arthropods.
This phylum is represented in nearly all habitats in the biosphere.
On the criteria of species diversity, distribution, and sheer numbers, arthropods must be
regarded as the most successful animal phylum.
Biology Chapter Notes
The diversity and success of arthropods are largely due to three features: body
segmentation, a hard exoskeleton, and jointed appendages.
Early arthropods such as the trilobites had pronounced segmentation, but little
variation in their appendages.
Groups of segments and their appendages have become specialized for a variety of
functions, permitting efficient division of labor among regions.
The body of an arthropod is completely covered by the cuticle, an exoskeleton
constructed from layers of protein and chitin.
The exoskeleton protects the animal and provides points of attachment for the muscles
that move appendages.
It is thick and inflexible in some regions, such as crab claws, and thin and flexible in
others, such as joints.
The exoskeleton of arthropods is strong and relatively impermeable to water.
In order to grow, an arthropod must molt its old exoskeleton and secrete a larger one, a
process called ecdysis that leaves the animal temporarily vulnerable to predators and
other dangers.
The exoskeleton’s relative impermeability to water helped prevent desiccation and
provided support on land.
Arthropods moved to land after the colonization of land by plants and fungi.
In 2004, an amateur fossil hunter found a 428-million-year-old fossil of a millipede.
Fossilized arthropod tracks date from 450 million years ago.
Arthropods have well-developed sense organs, including eyes for vision, olfactory
receptors for smell, and antennae for touch and smell.
Most sense organs are located at the anterior end of the animal, which shows extensive
cephalization.
Arthropods have an open circulatory system in which hemolymph fluid is propelled by a
heart through short arteries into sinuses (the hemocoel) surrounding tissues and organs.
Hemolymph returns to the heart through valved pores.
The hemocoel is not a coelom; the true coelom is much reduced in most arthropods.
Open circulatory systems evolved convergently in arthropods and molluscs.
Arthropods have evolved a variety of specialized organs for gas exchange.
Most aquatic species have gills with thin, feathery extensions that have an extensive
surface area in contact with water.
Terrestrial arthropods generally have internal surfaces specialized for gas exchange.
For example, insects have tracheal systems, branched air ducts leading into the
interior from pores in the cuticle.
Molecular systematics is suggesting new hypotheses about arthropod relationships.
Evidence shows that arthropods diverged early in their history into four main
evolutionary lineages: cheliceriformes (sea spiders, horseshoe crabs, scorpions, ticks,
spiders), myriapods (centipedes and millipedes), hexapods (insects and their
wingless, six-legged relatives), and crustaceans (crabs, lobsters, shrimps, barnacles,
and many others).
Cheliceriformes are named for their clawlike feeding appendages, chelicerae, which serve
as pincers or fangs.
Cheliceriformes have an anterior cephalothorax and a posterior abdomen.
They lack sensory antennae, and most have simple eyes (eyes with a single lens).
Biology Chapter Notes
The earliest cheliceriformes were eurypterids, or water scorpions, marine and
freshwater predators that grew up to 3 m long.
Modern marine cheliceriformes include the sea spiders (pycnogonids) and the
horseshoe crabs.
The majority of living cheliceriformes are arachnids, a group that includes scorpions,
spiders, ticks, and mites.
Nearly all ticks are blood-sucking parasites on the body surfaces of reptiles or mammals.
Parasitic mites live on or in a wide variety of vertebrates, invertebrates, and plants.
The arachnid cephalothorax has six pairs of appendages.
There are four pairs of walking legs.
A pair of pedipalps function in sensing or feeding.
The chelicerae usually function in feeding.
Spiders inject poison from glands on the chelicerae to immobilize their prey and while
chewing their prey, spill digestive juices into the tissues and suck up the liquid meal.
In most spiders, gas exchange is carried out by book lungs.
These are stacked plates contained in an internal chamber.
The plates present an extensive surface area, enhancing exchange of gases between the
hemolymph and air.
A unique adaptation of many spiders is the ability to catch flying insects in webs of silk.
The silk protein is produced as a liquid by abdominal glands and spun by spinnerets
into fibers that solidify.
Web designs are characteristic of each species.
Silk fibers have other functions as egg covers, drop lines for a rapid escape, and “gift
wrapping” for nuptial gifts.
Millipedes and centipedes belong to the subphylum Myriapoda, the myriapods.
All living myriapods are terrestrial.
Millipedes (class Diplopoda) have two pairs of walking legs on each of their many
trunk segments, formed by two fused segments.
They eat decaying leaves and plant matter.
They may have been among the earliest land animals.
Centipedes (class Chilopoda) are terrestrial carnivores.
The head has a pair of antennae and three pairs of appendages modified as mouth
parts, including the jawlike mandibles.
Each segment in the trunk region has one pair of walking legs.
Centipedes have poison claws on the anteriormost trunk segment that paralyze prey
and aid in defense.
Insects and their relatives (subphylum Hexapoda) are more species-rich than all other
forms of life combined.
They live in almost every terrestrial habitat and in fresh water, and flying insects fill the
air.
They are rare, but not absent, from the sea, where crustaceans dominate.
The oldest insect fossils date back to the Devonian period, about 416 million years ago.
When insect flight evolved in the Carboniferous and Permian periods, it sparked an
explosion in insect varieties.
Biology Chapter Notes
Diversification of mouthparts for feeding on gymnosperms and other Carboniferous
plants also contributed to the adaptive radiation of insects.
In one widely held hypothesis, the radiation of flowering plants triggered the greatest
diversification of insects in the Cretaceous and early Tertiary periods.
However, new research suggests that insects diversified first and, as pollinators and
herbivores, may have caused the angiosperm radiation.
Flight is one key to the great success of insects.
Flying animals can escape many predators, find food and mates, and disperse to new
habitats faster than organisms that must crawl on the ground.
Many insects have one or two pairs of wings that emerge from the dorsal side of the
thorax.
Wings are extensions of the cuticle and are not true appendages.
Several hypotheses have been proposed for the evolution of wings.
In one hypothesis, wings first evolved as extensions of the cuticle that helped the insect
absorb heat and were later modified for flight.
A second hypothesis argues that wings allowed animals to glide from vegetation to the
ground.
Alternatively, wings may have served as gills in aquatic insects.
Still another hypothesis proposes that insect wings functioned for swimming before
they functioned for flight.
Insect wings are also very diverse.
Dragonflies, among the first insects to fly, have two similar pairs of wings.
The wings of bees and wasps are hooked together and move as a single pair.
Butterfly wings operate similarly because the anterior wings overlap the posterior
wings.
In beetles, the posterior wings function in flight, while the anterior wings act as covers
that protect the flight wings when the beetle is on the ground or burrowing.
The internal anatomy of an insect includes several complex organ systems.
In the complete digestive system, there are regionally specialized organs with discrete
functions.
Metabolic wastes are removed from the hemolymph by Malpighian tubules,
outpockets of the digestive tract.
Respiration is accomplished by a branched, chitin-lined tracheal system that carries
O2 from the spiracles directly to the cells.
The insect nervous system consists of a pair of ventral nerve cords with several segmental
ganglia.
The two cords meet in the head, where the ganglia from several anterior segments are
fused into a cerebral ganglion (brain).
This structure is close to the antennae, eyes, and other sense organs concentrated on
the head.
Metamorphosis is central to insect development.
In incomplete metamorphosis (seen in grasshoppers and some other orders), the
young resemble adults but are smaller and have different body proportions.
Through a series of molts, the young look more and more like adults until they
reach full size.
Vertebrates are named for vertebrae, the series of bones that make up the vertebral
column or backbone.
There are about 52,000 species of vertebrates, far fewer than the 1 million insect species
on Earth.
Plant-eating dinosaurs, at 40,000 kg, were the heaviest animals to walk on land.
The biggest animal that ever existed is the blue whale, at 100,000 kg.
Humans and our closest relatives are vertebrates.
This group includes other mammals, birds, lizards, snakes, turtles, amphibians, and the
various classes of fishes.
Concept 34.1 Chordates have a notochord and a dorsal, hollow nerve cord
The vertebrates belong to one of the two major phyla in the Deuterostomia, the
chordates.
Chordates are bilaterian animals, belonging to the Deuterostomia.
The phylum Chordata includes three subphyla, the vertebrates and two phyla of
invertebrates—the urochordates and the cephalochordates.
Four derived characters define the phylum Chordata.
Although chordates vary widely in appearance, all share the presence of four anatomical
structures at some point in their lifetime.
These chordate characteristics are a notochord; a dorsal, hollow nerve cord; pharyngeal
slits; and a muscular, post-anal tail.
o The notochord, present in all chordate embryos, is a longitudinal, flexible rod located
between the digestive tube and the nerve cord.
It is composed of large, fluid-filled cells encased in fairly stiff, fibrous tissue.
It provides skeletal support throughout most of the length of the animal.
While the notochord persists in the adult stage of some invertebrate chordates and
primitive vertebrates, it remains only as a remnant in vertebrates with a more complex,
jointed skeleton.
For example, it is the gelatinous material of the disks between vertebrae in humans.
o The dorsal, hollow nerve cord of a chordate embryo develops from a plate of ectoderm
that rolls into a tube dorsal to the notochord.
Other animal phyla have solid nerve cords, usually located ventrally.
The nerve cord of the chordate embryo develops into the central nervous system: the
brain and spinal cord.
o The digestive tube of chordates extends from the mouth to the anus.
The region posterior to the mouth is the pharynx.
Concept 34.5 Tetrapods are gnathostomes that have limbs and feet
One of the most significant events in vertebrate history took place 360 million years ago,
when the fins of some lobe-fins evolved into tetrapod limbs and feet.
The most significant character of tetrapods is the four limbs, which allow them to support
their weight on land.
The feet of tetrapods have digits that allow them to transmit muscle-generated forces to
the ground when they walk.
With the move to land, the bones of the pelvic girdle (to which the hind legs are attached)
became fused to the backbone, permitting forces generated by the hind legs against the
ground to be transferred to the rest of the body.
Living tetrapods do not have pharyngeal gill slits.
The ears are adapted to the detection of airborne sounds.
The Devonian coastal wetlands were home to a wide range of lobe-fins. Those that
entered shallow, oxygen-poor water could use their lungs to breathe air.
Some species likely used their stout fins to move across the muddy bottom.
At the water’s edge, leglike appendages were probably better equipment than fins for
paddling and crawling through the dense vegetation in shallow water.
The tetrapod body plan was thus a modification of a preexisting body plan.
In one lineage of lobe-fins, the fins became progressively more limb-like, while the rest of
the body retained adaptations for aquatic life.
For example, fossils of Acanthostega from 365 million years ago had bony gill
supports and rays in its tail to support propulsion in water, but it also had fully formed
legs, ankles, and digits.
Acanthostega is representative of a period of vertebrate evolution when adaptations for
shallow water allowed certain fishes to make a gradual transition to the terrestrial side
of the water’s edge.
A great diversity of tetrapods emerged during the Carboniferous period.
Judging from the morphology and location of the fossils, most of these early tetrapods
remained tied to water.
Class Amphibia: Salamanders, frogs, and caecilians are the three extant amphibian
orders.
Concept 34.6 Amniotes are tetrapods that have a terrestrially adapted egg
The amniote clade consists of the mammals and reptiles (including birds).
The evolution of amniotes from an amphibian ancestor involved many adaptations for
terrestrial living.
The amniotic egg is the major derived character of the clade.
Inside the shell of the amniotic egg are several extraembryonic membranes that function
in gas exchange, waste storage, and the transfer of stored nutrients to the embryo.
The amniotic egg is named for one of these membranes, the amnion, which encloses a
fluid-filled “private pond” that bathes the embryo and acts as a hydraulic shock
absorber.
The amniotic eggs enabled terrestrial vertebrates to complete their life cycles entirely on
land.
In contrast to the shell-less eggs of amphibians, the amniotic eggs of most amniotes
have a shell that retains water and can be laid in a dry place.
The calcareous shells of bird eggs are inflexible, while the leathery eggs of many
reptiles are flexible.
Most mammals have dispensed with the shell.
The embryo implants in the wall of the uterus and obtains its nutrition from the
mother.
Amniotes acquired other adaptations to terrestrial life, including less-permeable skin and
the increasing use of the rib cage to ventilate the lungs.
Amniotes adopt a more elevated stance than earlier tetrapods and living amphibians.
The most recent common ancestor of living amphibians and amniotes lived about 340
million years ago, in the early Carboniferous period.
No fossils of amniotic eggs have been found from that time.
Biology Chapter Notes
Early amniotes lived in drier environments than did earlier tetrapods.
Some were herbivores, with grinding teeth. Others were large and predatory.
The reptile clade includes birds.
The reptile clade includes tuatara, lizards, snakes, turtles, crocodilians, and birds, as well
as extinct groups such as dinosaurs.
Reptiles have several adaptations for terrestrial life not generally found in amphibians.
Scales containing the protein keratin waterproof the skin, preventing dehydration in
dry air.
Crocodiles, which are adapted to water, have evolved more permeable scales called
scutes.
Reptiles obtain all their oxygen with lungs, not through their dry skin.
As an exception, many turtles can use the moist surfaces of their cloaca for gas
exchange.
Most reptiles lay shelled amniotic eggs on land.
Fertilization occurs internally, before the shell is secreted as the egg passes through the
female’s reproductive tract.
Some species of lizards and snakes are viviparous, with their extraembryonic
membranes forming a placenta that enables the embryo to obtain nutrients from its
mother.
Nonbird reptiles are sometimes labeled “cold-blooded” because they do not use their
metabolism extensively to control body temperature.
However, many nonbird reptiles regulate their body temperature behaviorally by
basking in the sun when cool and seeking shade when hot.
Because they absorb external heat rather than generating much of their own, nonbird
reptiles are more appropriately called ectotherms.
One advantage of this strategy is that an ectothermic reptile can survive on less than
10% of the calories required by a mammal of equivalent size.
The reptile clade is not entirely ectothermic.
Birds are endothermic, capable of keeping the body warm through metabolism.
The oldest reptilian fossils date back to the Carboniferous period, about 300 million years
ago.
The first major group of reptiles to emerge was the parareptiles, large, stocky,
quadrupedal herbivores.
Some parareptiles had dermal plates on their skin, which may have provided defense
against predators.
Parareptiles died out 200 million years ago, at the end of the Triassic period.
As parareptiles were dwindling, an equally ancient clade of reptiles, the diapsids, was
diversifying.
The most obvious derived character of diapsids is a pair of holes on each side of the
skull, behind the eye socket.
The diapsids are composed of two main lineages.
One, the lepidosaurs, includes lizards, snakes, and tuataras.
This lineage also produced a number of marine reptiles including plesiosaurs and
ichthyosaurs.
Biology Chapter Notes
The archosaurs include crocodilians, and the extinct pterosaurs and dinosaurs.
Pterosaurs, which originated in the late Triassic, were the first flying tetrapods.
The pterosaur wing is formed from a bristle-covered membrane of skin that stretched
between the hind leg and the tip of an elongated finger.
Well-preserved fossils show the presence of muscles, blood vessels, and nerves in the
wing membrane, suggesting that pterosaurs could dynamically adjust their membranes
to assist their flight.
Dinosaurs were an extremely diverse group varying in body shape, size, and habitat.
There were two main dinosaur lineages: the ornithischians, which were mostly
herbivorous, and the saurischians, which included both long-necked giant herbivores
and bipedal carnivorous theropods.
Theropods included the famous Tyrannosaurus rex as well as the ancestors of birds.
There is increasing evidence that many dinosaurs were agile; fast moving; and, in some
species, social.
Paleontologists have discovered signs of parental care among dinosaurs.
There is continuing debate about whether dinosaurs were endothermic, capable of
keeping their body warm through metabolism.
Some experts are skeptical.
In the warm, consistent Mesozoic climate, behavioral adaptations may have been
sufficient for maintaining a suitable body temperature for terrestrial dinosaurs.
Also, the low surface-to-volume ratios would have reduced the effects of daily
fluctuations in air temperature on the animal’s internal temperature.
Some anatomical evidence supports the hypothesis that at least some dinosaurs were
endotherms.
Paleontologists have found fossils of dinosaurs in both Antarctica and the Arctic,
although the climate in those areas was milder during the Mesozoic than today.
The dinosaur that gave rise to birds was certainly endothermic, as are all birds.
By the end of the Cretaceous, all dinosaurs (except birds) became extinct.
It is uncertain whether dinosaurs were declining before they were finished off by an
asteroid or comet impact.
Lepidosaurs are represented by two living lineages.
One lineage includes the tuatara, two species of lizard-like reptiles found only on 30
islands off the coast of New Zealand.
Tuatara relatives lived at least 220 million years ago, when they thrived on every
continent well into the Cretaceous period.
The other major living lineage of lepidosaurs are the squamates (lizards and snakes).
Lizards are the most numerous and diverse reptiles alive today.
Most are relatively small, but they range in length from 16 mm to 3 m.
Snakes are legless lepidosaurs that evolved from lizards closely related to the Komodo
dragon.
It was once thought that snakes were descendents of lizards that adapted to a burrowing
lifestyle through the loss of limbs.
However, recently discovered fossils of aquatic snakes with complete hind legs
suggest that snakes likely evolved in water and then recolonized land.
Concept 34.7 Mammals are amniotes that have hair and produce milk
Mammals diversified extensively in the wake of the Cretaceous extinctions.
Mammals have a number of derived traits.
All mammalian mothers use mammary glands to nourish their babies with milk, a
balanced diet rich in fats, sugars, proteins, minerals, and vitamins.
All mammals also have hair, made of keratin.
Hair and a layer of fat under the skin retain metabolic heat, contributing to
endothermy in mammals.
Endothermy is supported by an active metabolism, made possible by efficient
respiration and circulation.
Adaptations include a muscular diaphragm and a four-chambered heart.
Mammals generally have larger brains than other vertebrates of equivalent size.
Many species are capable of learning.
The relatively long period of parental care extends the time for offspring to learn
important survival skills by observing their parents.
Feeding adaptations of the jaws and teeth are other important mammalian traits.
Unlike the uniform conical teeth of most reptiles, the teeth of mammals come in a
variety of shapes and sizes adapted for processing many kinds of foods.
During the evolution of mammals from reptiles, two bones formerly in the jaw joint
were incorporated into the mammalian ear and the jaw joint was remodeled.
Mammals belong to a group of amniotes known as synapsids.
Synapsids have a temporal fenestra behind the eye socket on each side of the skull.
Synapsids evolved into large herbivores and carnivores during the Permian period.
Mammal-like synapsids emerged by the end of the Triassic, 200 million years ago.
These animals were not mammals, but they were small and likely hairy, fed on insects
at night, and had a higher metabolism that other synapsids.
They likely laid eggs.
The first true mammals arose in the Jurassic periods.
Early mammals diversified into a number of lineages, all about the size of a shrew.
During the Mesozoic, mammals coexisted with dinosaurs and underwent a great adaptive
radiation in the Cenozoic in the wake of the Cretaceous extinctions.
Modern mammals are split into three groups: monotremes (egg-laying mammals),
marsupials (mammals with pouches), and eutherian (placental) mammals.
The fanwort, an aquatic weed, demonstrates the great developmental plasticity that is
characteristic of plants. The fanwort has feathery underwater leaves and large, flat,
floating surface leaves. Both leaf types have genetically identical cells, but the dissimilar
environments in which they develop cause different genes involved in leaf formation to be
turned on or off.
The form of any plant is controlled by environmental and genetic factors. As a result, no
two plants are identical.
In addition to plastic structural responses of individual plants to specific environments,
plant species have adaptive features that benefit them in their specific environments.
For example, cacti have leaves that are reduced as spines and a stem that serves as the
primary site of photosynthesis. These adaptations reduce water loss in desert
environments.
Angiosperms comprise 90% of plant species and are at the base of the food web of nearly
every terrestrial ecosystem.
Most land animals, including humans, depend on angiosperms directly or indirectly for
sustenance.
Concept 35.1 The plant body has a hierarchy of organs, tissues, and cells
Plants, like multicellular animals, have organs that are composed of different tissues, and
tissues are composed of different cell types.
A tissue is a group of cells with a common structure and function.
An organ consists of several types of tissues that work together to carry out particular
functions.
Vascular plants have three basic organs: roots, stems, and leaves.
The basic morphology of vascular plants reflects their evolutionary history as terrestrial
organisms that inhabit and draw resources from two very different environments.
Plants obtain water and minerals from the soil.
They obtain CO2 and light above ground.
To obtain the resources they need, vascular plants have evolved two systems: a
subterranean root system and an aerial shoot system of stems and leaves.
Each system depends on the other.
Lacking chloroplasts and living in the dark, roots would starve without the sugar and
other organic nutrients imported from the photosynthetic tissues of the shoot system.
Conversely, the shoot system (and its reproductive tissues, flowers) depends on water
and minerals absorbed from the soil by the roots.
A root is an organ that anchors a vascular plant in the soil, absorbs minerals and water,
and stores food.
Concept 35.4 Secondary growth adds girth to stems and roots in woody plants
The stems and roots of most eudicots increase in girth by secondary growth.
The secondary plant body consists of the tissues produced during this secondary
growth in diameter.
Primary and secondary growth occur simultaneously but in different regions.
While elongation of the stem (primary growth) occurs at the apical meristem, increases
in diameter (secondary growth) occur farther down the stem.
The vascular cambium is a cylinder of meristematic cells that forms secondary vascular
tissue.
It forms successive layers of secondary xylem to its interior and secondary phloem to
its exterior.
The accumulation of this tissue over the years accounts for most of the increase in
diameter of a woody plant.
The vascular cambium develops from parenchyma cells that retain the capacity to
divide.
This meristem forms in a layer between the primary xylem and primary phloem of
each vascular bundle and in the ground tissue between the bundles.
The meristematic bands unite to form a continuous cylinder of dividing cells.
This ring of vascular cambium consists of regions of ray initials and fusiform initials.
The tapered, elongated cells of the fusiform initials form secondary xylem to the
inside of the vascular cambium and secondary phloem to the outside.
Ray initials produce vascular rays that transfer water and nutrients laterally within the
woody stem and also store starch and other reserves.
As secondary growth continues over the years, layer upon layer of secondary xylem
accumulates, producing the tissue we call wood.
Wood consists mainly of tracheids, vessel elements (in angiosperms), and fibers.
These cells, dead at functional maturity, have thick, lignified walls that give wood its
hardness and strength.
In temperate regions, secondary growth in perennial plants ceases during the winter.
The first tracheid and vessel cells formed in the spring (early wood) have larger
diameters and thinner walls than cells produced later in the summer (late wood).
The structure of the early wood maximizes delivery of water to new, expanding leaves.
The thick-walled cells of later wood provide more physical support.
This pattern of growth—cambium dormancy, early wood production, and late wood
production—produces annual growth rings.
As a tree or woody shrub ages, the older layers of secondary xylem, known as
heartwood, no longer transport water and minerals.
Biology Chapter Notes
The outer layers, known as sapwood, continue to transport xylem sap.
Only the youngest secondary phloem, closest to the vascular cambium, functions in sugar
transport.
The older secondary phloem dies and is sloughed off as part of the bark.
The cork cambium acts as a meristem for a tough, thick covering for stems and roots that
replaces the epidermis.
Early in secondary growth, the epidermis produced by primary growth splits, dries, and
falls off the stem or root.
It is replaced by two tissues produced by the first cork cambium, which arises in the
outer cortex of stems and in the outer layer of the pericycle of roots.
The first tissue, phelloderm, is a thin layer of parenchyma cells that forms to the
interior of the cork cambium.
Cork cambium also produces cork cells, which accumulate at the cambium’s
exterior.
Waxy material called suberin deposited in the cell walls of cork cells before they
die acts as a barrier against water loss, physical damage, and pathogens.
The cork plus the cork cambium form the periderm, a protective layer that replaces the
epidermis.
In areas called lenticels, spaces develop between the cork cells of the periderm.
These areas within the trunk facilitate gas exchange with the outside air.
Unlike the vascular cambium, cells of the cork cambium do not divide.
The thickening of a stem or root splits the first cork cambium, which loses its
meristematic activity and differentiates into cork cells.
A new cork cambium forms to the inside, resulting in a new layer of periderm.
As this process continues, older layers of periderm are sloughed off.
This produces the cracked, peeling bark of many tree trunks.
Bark refers to all tissues external to the vascular cambium, including secondary phloem,
cork cambium, and cork.
Concept 35.5 Growth, morphogenesis, and differentiation produce the plant body
During plant development, a single cell, the zygote, gives rise to a multicellular plant of
particular form with functionally integrated cells, tissues, and organs.
An increase in mass, or growth, results from cell division and cell expansion.
The development of body form and organization is called morphogenesis.
The specialization of cells with the same set of genetic instructions to produce a
diversity of cell types is called differentiation.
Plants have tremendous developmental plasticity.
Plant form, including height, branching patterns, and reproductive output, is greatly
influenced by environmental factors.
A broad range of morphologies can result from the same genotype as three
developmental processes—growth, morphogenesis, and differentiation—transform a
zygote into an adult plant.
Molecular biology is revolutionizing the study of plants.
Biology Chapter Notes
Modern molecular techniques allow plant biologists to investigate how growth,
morphogenesis, and cellular differentiation give rise to a plant.
Much of this research has focused on Arabidopsis thaliana, a small weed in the
mustard family.
Thousands of these small plants can be cultivated in a few square meters of lab space.
With a generation time of about six weeks, it is an excellent model for genetic studies.
The genome of Arabidopsis is among the tiniest of all known plants.
Arabidopsis was the first plant to have its genome sequenced, in a six-year multinational
project.
Arabidopsis has a total of about 26,000 genes, with fewer than 15,000 different types of
genes.
Now that the DNA sequence of Arabidopsis is known, plant biologists are working to
identify the functions of every one of the plant’s genes by the year 2010.
To aid in this effort, biologists are attempting to create mutants for every gene in the
plant’s genome.
Study of the function of these genes has already expanded our understanding of plant
development.
By identifying each gene’s function, researchers aim to establish a blueprint for how
plants are built.
One day it may be possible to create a computer-generated “virtual plant” that will
enable researchers to visualize which plant genes are activated in different parts of the
plant during the entire course of development.
Growth involves both cell division and cell expansion.
Cell division in meristems increases cell number, increasing the potential for growth.
However, it is cell expansion that accounts for the actual increase in plant mass.
The plane (direction) and symmetry of cell division are important determinants of plant
form.
If the planes of division by a single cell and its descendents are parallel to the plane of
the first cell division, a single file of cells will be produced.
If the planes of cell division of the descendent cells vary at random, an unorganized
clump of cells will result.
While mitosis results in symmetrical redistribution of chromosomes between daughter
cells, cytokinesis may be asymmetrical.
Asymmetrical cell division, in which one cell receives more cytoplasm than the other,
is common in plant cells and usually signals a key developmental event.
For example, guard cells form from an unspecialized epidermal cell through an
asymmetrical cell division and a change in the plane of cell division.
The plane in which a cell will divide is determined during late interphase.
Microtubules in the outer cytoplasm become concentrated into a ring, the preprophase
band.
While this disappears before metaphase, its “imprint” consists of an ordered array of
actin microfilaments that remains after the microtubules disperse and signals the future
plane of cell division.
Cell expansion in animal cells is quite different from cell expansion in plant cells.
Animal cells grow by synthesizing a protein-rich cytoplasm, a metabolically expensive
process.
Biology Chapter Notes
While growing plant cells add some organic material to their cytoplasm, water uptake
by the large central vacuole accounts for 90% of a plant cell’s expansion.
This enables plants to grow economically and rapidly.
Bamboo shoots can elongate more than 2 m per week.
Rapid expansion of shoots and roots increases their exposure to light and soil, an
important evolutionary adaptation to the immobile lifestyle of plants.
The greatest expansion of a plant cell is usually oriented along the plant’s main axis.
The orientations of cellulose microfibrils in the innermost layers of the cell wall cause
this differential growth, as the cell expands mainly perpendicular to the “grain” of the
microfibrils.
Studies of Arabidopsis mutants have confirmed the importance of cortical microtubules in
both cell division and expansion.
For example, fass mutants have unusually squat cells, which follow seemingly random
planes of cell division.
Their roots and stems lack the ordered cell files and layers.
Fass mutants develop into tiny adult plants with all their organs compressed
longitudinally.
The cortical microtubular organization of fass mutants is abnormal.
Although the microtubules involved in chromosome movement and in cell plate
deposition are normal, preprophase bands do not form prior to mitosis.
In interphase cells, the cortical microtubules are randomly positioned.
Therefore, the cellulose microfibrils deposited in the cell wall cannot be arranged to
determine the direction of the cell’s elongation.
Cells with a fass mutation expand in all directions equally and divide in a
haphazard arrangement, leading to stout stature and disorganized tissues.
Morphogenesis depends on pattern formation.
Morphogenesis organizes dividing and expanding cells into multicellular tissues and
organs.
The development of specific structures in specific locations is called pattern
formation.
Pattern formation depends to a large extent on positional information, signals that
continuously indicate each cell’s location within an embryonic structure.
Within a developing organ, each cell responds to positional information by
differentiating into a particular cell type.
Developmental biologists are accumulating evidence that gradients of specific molecules,
generally proteins or mRNAs, provide positional information.
For example, a substance diffusing from a shoot’s apical meristem may “inform” the
cells below of their distance from the shoot tip.
A second chemical signal produced by the outermost cells may enable a cell to gauge
their position relative to the radial axis of the developing organ.
Developmental biologists are testing the hypothesis that diffusible chemical signals
provide plant cells with positional information.
One type of positional information is polarity, the identification of the root end and shoot
end along a well-developed axis.
This polarity results in morphological and physiological differences, and it impacts the
emergence of adventitious roots and shoots from the appropriate ends of plant cuttings.
Biology Chapter Notes
The first division of the zygote is asymmetrical and may initiate the polarization of the
plant body into root and shoot ends.
Once the polarity has been induced, it is very difficult to reverse experimentally.
The establishment of axial polarity is a critical step in plant morphogenesis.
In the gnom mutant of Arabidopsis, the first division is symmetrical, and the resulting
ball-shaped plant lacks roots and leaves.
Other genes that regulate pattern formation and morphogenesis include the homeotic
genes, which mediate many developmental events, such as organ initiation.
For example, the protein product of the KNOTTED-1 homeotic gene is important for
the development of leaf morphology, including production of compound leaves.
Overexpression of this gene causes the compound leaves of a tomato plant to become
“supercompound.”
Cellular differentiation depends on the control of gene expression.
The diverse cell types of a plant, including guard cells, sieve-tube members, and xylem
vessel elements, all descend from a common cell, the zygote, and share the same DNA.
The cloning of whole plants from single somatic cells demonstrates that the genome of a
differentiated cell remains intact and can “dedifferentiate” to give rise to the diverse cell
types of a plant.
Cellular differentiation depends, to a large extent, on control of gene expression.
Cells with the same genomes follow different developmental pathways because they
selectively express certain genes at specific times during differentiation.
For example, two distinct cell types in Arabidopsis, root hair cells and hairless epidermal
cells, develop from immature epidermal cells.
Cells in contact with one underlying cortical cell differentiate into mature, hairless
cells, while those in contact with two underlying cortical cells differentiate into root
hair cells.
The homeotic gene GLABRA-2 is normally expressed only in hairless cells. If it is
rendered dysfunctional, every root epidermal cell develops a root hair.
Clonal analysis of the shoot apex emphasizes the importance of a cell’s location in its
developmental fate.
In the process of shaping a rudimentary organ, patterns of cell division and cell expansion
affect the differentiation of cells by placing them in specific locations relative to other
cells.
Thus, positional information underlies all the processes of development: growth,
morphogenesis, and differentiation.
One approach to studying the relationship among these processes is clonal analysis,
mapping the cell lineages (clones) derived from each cell in an apical meristem as organs
develop.
Researchers induce some change in a cell that tags it in some way such that it (and its
descendents) can be distinguished from its neighbors.
For example, a somatic mutation in an apical cell that prevents chlorophyll production
will produce an “albino” cell.
This cell and all its descendants will appear as a linear file of colorless cells running
down the long axis of the green shoot.
To some extent, the developmental fates of cells in the shoot apex are predictable.
The algal ancestors of plants obtained water, minerals and CO 2 from the water in which
they were completely immersed.
For vascular plants, the evolutionary journey onto land involved the differentiation of the
plant body into roots, which absorb water and minerals from the soil, and shoots, which
absorb light and atmospheric CO2 for photosynthesis.
This morphological solution created a new problem: the need to transport materials
between roots and shoots.
Xylem transports water and minerals from the roots to the shoots.
Phloem transports sugars from the site of production to the regions that need them for
growth and metabolism.
Concept 36.1 Physical forces drive the transport of materials in plants over a range of
distances
Transport in plants occurs on three levels:
o The uptake and loss of water and solutes by individual cells, such as root hairs.
o Short-distance transport of substances from cell to cell at the level of tissues or organs,
such as the loading of sugar from photosynthetic leaf cells into the sieve tubes of
phloem.
o Long-distance transport of sap within xylem and phloem at the level of the whole
plant.
Transport at the cellular level depends on the selective permeability of membranes.
The selective permeability of a plant cell’s plasma membrane controls the movement of
solutes between the cell and the extracellular solution.
Molecules tend to move down their concentration gradient. Diffusion across a
membrane is called passive transport and occurs without the direct expenditure of
metabolic energy by the cell.
Active transport is the pumping of solutes across membranes against their
electrochemical gradients, and requires expenditure of energy by the cell.
The cell must expend metabolic energy, usually in the form of ATP, to transport
solutes “uphill.”
Transport proteins embedded in the membrane can speed movement across the
membrane.
Some transport proteins bind selectively to a solute on one side of the membrane and
release it on the opposite side.
Others act as selective channels, providing a selective passageway across the
membrane.
For example, the membranes of most plant cells have potassium channels that allow
potassium ions (K+) to pass, but not similar ions, such as sodium (Na+).
Concept 36.2 Roots absorb water and minerals from the soil
Water and mineral salts from soil enter the plant through the epidermis of roots, cross the
root cortex, pass into the vascular cylinder, and then flow up xylem vessels to the shoot
system.
o The uptake of soil solution by the hydrophilic epidermal walls of root hairs provides
access to the apoplast, and water and minerals can soak into the cortex along this route.
o Minerals and water that cross the plasma membranes of root hairs enter the symplast.
o Some water and minerals are transported into cells of the epidermis and cortex and
then move inward via the symplast.
o Materials flowing along the apoplastic route are blocked by the waxy Casparian strip at
the endodermis. Some minerals detour around the Casparian strip by crossing the
plasma membrane of an endodermal cell to pass into the vascular cylinder.
o Endodermal and parenchyma cells within the vascular cylinder discharge water and
minerals into their walls (apoplast). The water and minerals enter the dead cells of
xylem vessels and are transported upward into the shoots.
Root hairs, mycorrhizae, and a large surface area of cortical cells enhance water and
mineral absorption.
Much of the absorption of water and minerals occurs near root tips, where the epidermis is
permeable to water and where root hairs are located.
Root hairs, extensions of epidermal cells, account for much of the surface area of roots.
The soil solution flows into the hydrophilic walls of epidermal cells and passes freely
along the apoplast into the root cortex, exposing all the parenchyma cells to soil
solution and increasing membrane surface area.
As the soil solution moves along the apoplast into the roots, cells of the epidermis and
cortex take up water and certain solutes into the symplast.
Biology Chapter Notes
Selective transport proteins of the plasma membrane and tonoplast enable root cells to
extract essential minerals from the dilute soil solution and concentrate them hundreds
of times higher than in the soil solution.
This selective process
+
enables the cell to extract K +, an essential mineral nutrient, and
exclude most Na .
Most plants form partnerships with symbiotic fungi to absorb water and minerals from
soil.
“Infected” roots form mycorrhizae, symbiotic structures consisting of the plant’s roots
united with the fungal hyphae.
Hyphae absorb water and selected minerals, transferring much of these to the host plants.
The mycorrhizae create an enormous surface area for absorption and enable older regions
of the roots to supply water and minerals to the plant.
The endodermis functions as a selective sentry between the root cortex and vascular
tissue.
Water and minerals in the root cortex cannot be transported to the rest of the plant until
they enter the xylem of the vascular cylinder.
The endodermis, the innermost layer of cells in the root cortex, surrounds the vascular
cylinder and functions as a final checkpoint for the selective passage of minerals from
the cortex into the vascular tissue.
Minerals already in the symplast continue through the plasmodesmata of the
endodermal cells and pass into the vascular cylinder.
These minerals were already screened by the selective membrane they crossed to enter
the symplast.
Those minerals that reach the endodermis via the apoplast are blocked by the Casparian
strip in the walls of each endodermal cell.
This strip is a belt of suberin, a waxy material that is impervious to water and dissolved
minerals.
To enter the vascular cylinder, minerals must cross the plasma membrane of the
endodermal cell and enter the vascular cylinder via the symplast.
The endodermis, with its Casparian strip, ensures that no minerals reach the vascular
tissue of the root without crossing a selectively permeable plasma membrane.
The endodermis acts as a sentry on the cortex-vascular cylinder border.
The last segment in the soil-to-xylem pathway is the passage of water and minerals into
the tracheids and vessel elements of the xylem.
Because these cells lack protoplast, the lumen and the cell walls are part of the
apoplast.
Endodermal cells and parenchyma cells within the vascular cylinder discharge
minerals into their walls.
Both diffusion and active transport are involved in the transfer of solutes from the
symplast to apoplast, finally entering the tracheids and xylem vessels.
Concept 36.3 Water and minerals ascend from roots to shoots through the xylem
Xylem sap flows upward to veins that branch throughout each leaf, providing each with
water.
Concept 37.1 Plants require certain chemical elements to complete their life cycle
Early ideas about plant nutrition were not entirely correct and included:
Aristotle’s hypothesis that soil provided the substance for plant growth.
van Helmont’s conclusion from his experiments that plants grow mainly from water.
Hale’s postulate that plants are nourished mostly by air.
In fact, soil, water, and air all contribute to plant growth.
Plants extract mineral nutrients from the soil. Mineral nutrients are essential chemical
elements absorbed from soil in the form of inorganic ions.
For example, many plants acquire nitrogen in the form of nitrate ions (NO 3−).
However, as van Helmont’s data suggested, mineral nutrients from the soil contribute
little to the overall mass of a plant.
About 80–90% of a plant is water. Because water contributes most of the hydrogen ions
and some of the oxygen atoms that are incorporated into organic atoms, one can consider
water a nutrient.
However, only a small fraction of the water entering a plant contributes to organic
molecules.
More than 90% of the water absorbed by a field of corn is lost by transpiration.
Most of the water retained by a plant functions as a solvent, provides most of the mass
for cell elongation, and helps maintain the form of soft tissues by keeping cells turgid.
By weight, the bulk of the organic material of a plant is derived not from water or soil
minerals, but from the CO2 assimilated from the atmosphere.
Concept 37.2 Soil quality is a major determinant of plant distribution and growth
Soil texture and composition are key environmental factors in terrestrial ecosystems.
The texture and chemical composition of soil are major factors determining what kinds of
plants can grow well in a particular location.
Texture is the general structure of soil, including the relative amounts of various sizes
of soil particles.
Composition is the soil’s organic and inorganic components.
Concept 37.3 Nitrogen is often the mineral that has the greatest effect on plant growth
The metabolism of soil bacteria makes nitrogen available to plants.
Of all mineral nutrients, nitrogen has the greatest effect on plant growth and crop yields.
It is ironic that plants sometimes suffer nitrogen deficiencies, for the atmosphere is nearly
80% nitrogen as N2.
Plants cannot use nitrogen in the form of N2.
It must first be converted to ammonium (NH4+) or nitrate (NO3−).
The main source of ammonium and nitrate is the decomposition of humus by
microbes, including ammonifying bacteria.
Nitrogen is lost from this local cycle when soil microbes called denitrifying bacteria
convert NO3− to N2, which diffuses into the atmosphere.
Other bacteria, nitrogen-fixing bacteria, restock nitrogenous minerals in the soil by
converting N2 to NH3 (ammonia) by the metabolic process of nitrogen fixation.
All life on Earth depends on nitrogen fixation, a process performed only by certain
bacterial species.
In soil, these include several species of free-living bacteria and several others that live
in symbiotic relationships with plants.
The reduction of N2 to NH3 is a complicated, multistep process, catalyzed by one
enzyme complex, nitrogenase, and simplified as:
N2 + 8e− + 8H+ + 16ATP -> 2NH3 + H2 + 16ADP + 16Pi
Nitrogen fixation is a very costly process, costing the bacterium 8 ATP for every
ammonia molecule synthesized.
Nitrogen-fixing bacteria are most abundant in soils rich in organic materials, which
provide fuels for cellular respiration to support this expensive metabolic process.
In the soil solution, ammonia picks up another hydrogen ion to form ammonium (NH 4+),
which plants can absorb.
Nitrifying bacteria in the soil oxidize ammonium to nitrate (NO 3−), the required form of
nitrogen for most plants.
After nitrate is absorbed by roots, plant enzymes reduce nitrate back to ammonium,
which other enzymes then incorporate into amino acids and other organic compounds.
Biology Chapter Notes
Most plant species export nitrogen from roots to shoots via the xylem, in the form of
nitrate or organic compounds that have been synthesized in the roots.
Improving the protein yield of crops is a major goal of agricultural research.
The ability of plants to incorporate fixed nitrogen into proteins and other organic
substances has a major impact on human welfare.
Protein deficiency is the most common form of malnutrition.
Either by choice or economic necessity, the majority of the world’s people have a
predominately vegetarian diet.
Unfortunately, plants are a poor source of protein and may be deficient in one or more
of the amino acids that humans need from their diet.
Plant breeding has resulted in new varieties of corn, wheat, and rice that are enriched in
protein.
However, many of these “super” varieties have an extraordinary demand for nitrogen,
which is usually supplied by commercial fertilizer produced by energy-costly
industrial production.
Generally, the countries that most need high-protein crops are the ones least able to
afford to pay for the fossil fuels to power the factories that make fertilizers.
Agricultural scientists are pursuing a variety of strategies to overcome this protein
deficiency.
For example, the use of new nitrogenase-based catalysts to fix nitrogen may make
commercial production of nitrogen fertilizers cheaper.
Alternatively, improvements in the productivity of symbiotic nitrogen fixation may
increase protein yields of crops.
Concept 37.4 Plant nutritional adaptations often involve relationships with other
organisms
The roots of plants belong to subterranean communities that interact with a diversity of
other organisms.
Among these are certain species of bacteria and fungi that have coevolved with
specific plants, forming symbiotic relationships with roots that enhance the nutrition of
both partners.
The two most important examples of mutualistic interactions are nitrogen fixation
(symbiosis of plant roots and bacteria) and the formation of mycorrhizae (symbiosis of
plant roots and fungi).
Symbiotic nitrogen fixation results from intricate interactions between roots and bacteria.
Some plant species form symbiotic relationships with nitrogen-fixing bacteria.
This provides their roots with a built-in source of fixed nitrogen for assimilation into
organic compounds.
Much of the research on this symbiosis has focused on the agriculturally important
members of the legume family, including peas, beans, soybeans, peanuts, alfalfa, and
clover.
A legume’s roots have swellings called nodules, composed of plant cells that contain
nitrogen-fixing bacteria of the genus Rhizobium.
Inside the nodule, Rhizobium bacteria assume a form called bacteriods, which are
contained within vesicles formed by the root cell.
Biology Chapter Notes
Legume-Rhizobium symbioses produce more usable nitrogen for plants than all
industrial fertilizers, at no cost to farmers. Subsequent crops can also benefit from the
usable nitrogen left in the soil by a legume crop.
Nitrogen fixation requires an anaerobic environment.
Lignified external layers of the nodule limit gas exchange.
Nodules produce leghemoglobin, an iron-containing protein that binds reversibly to
oxygen. Leghemoglobin provides oxygen for Rhizobium’s intense respiration, while
protecting nitrogenase from free oxygen.
The development of root nodules begins after bacteria enter the root through an infection
thread.
o Chemical signals from the root attract the Rhizobium bacteria, and chemical signals
from the bacteria lead to the production of an infection thread.
o The bacteria penetrate the root cortex within the infection thread.
o Growth in cortex and pericycle cells which are “infected” with bacteria in vesicles
continues until the two masses of dividing cells fuse, forming the nodule.
o As the nodule continues to grow, vascular tissue connects the nodule to the xylem and
phloem of the stele, providing nutrients to the nodule and carrying nitrogenous
compounds to the rest of the plant.
The symbiotic relationship between a legume and nitrogen-fixing bacteria is mutualistic,
with both partners benefiting.
The bacteria supply the legume with fixed nitrogen.
Most of the ammonium produced by symbiotic nitrogen fixation is used by the
nodules to make amino acids, which are then transported to the shoot and leaves via
the xylem.
The plant provides the bacteria with carbohydrates and other organic compounds and
protects the nitrogenase from free oxygen.
The common agricultural practice of crop rotation exploits symbiotic nitrogen fixation.
One year, a nonlegume crop such as corn is planted. The following year, alfalfa or
another legume is planted to restore the concentration of fixed soil nitrogen.
Often, the legume crop is not harvested but is plowed under to decompose as “green
manure.”
To ensure the formation of nodules, the legume seeds may be soaked in a culture of the
correct Rhizobium bacteria or dusted with bacterial spores before sowing.
Species from many other plant families also benefit from symbiotic nitrogen fixation.
For example, alder trees and certain tropical grasses host nitrogen-fixing bacteria of the
actinomycetes group.
Rice benefits indirectly from symbiotic nitrogen fixation because it is often cultivated
in paddies with the water fern Azolla, which has symbiotic nitrogen-fixing
cyanobacteria.
This increases the fertility of the rice paddy through the activity of the
cyanobacteria.
The growing rice eventually shades and kills the Azolla.
The decomposition of water fern adds more nitrogenous compounds to the paddy.
The molecular biology of root nodule formation is increasingly well understood.
The specific recognition between legume and bacteria and the development of the nodule
is the result of a chemical dialogue between the bacteria and the root.
Biology Chapter Notes
Each partner responds to the chemical signals of the other by expressing certain genes
whose products contribute to nodule formation.
The plant initiates the communication when its roots secrete molecules called
flavonoids, which enter Rhizobium cells living in the vicinity of the roots.
Each particular legume species secretes a type of flavonoid that only a certain
Rhizobium species can detect and absorb.
o A specific flavonoid signal travels from the root to the plant’s Rhizobium partner.
o The flavonoid activates a gene-regulating protein in the bacterium, which switches on
a cluster of bacterial genes called nod (for nodulation genes).
o The nod genes produce enzymes that catalyze production of species-specific molecules
called Nod factors.
o Nod factors signal the root to initiate the infection process, enabling Rhizobium to
enter the root and begin forming the root nodule.
o The plant’s responses 2+ require activation of early nodulin genes by a signal transduction
pathway involving Ca as second messengers.
It may be possible in the future to induce Rhizobium uptake and nodule formation in
crop plants that do not normally form such nitrogen-fixing symbioses.
In the short term, research is focused on improving the efficiency of nitrogen fixation
and protein production.
Mycorrhizae are symbiotic associations of roots and fungi that enhance plant nutrition.
Mycorrhizae (“fungus roots”) are modified roots, consisting of mutualistic associations
of fungi and roots.
The fungus benefits from a hospitable environment and a steady supply of sugar
donated by the host plant.
The fungus provides several potential benefits to the host plant.
First, the fungi increase the surface area for water uptake and selectively absorb
phosphate and other minerals in the soil and supply them to the plant.
The fungi also secrete growth factors that stimulate roots to grow and branch.
The fungi produce antibiotics that may help protect the plant from pathogenic bacteria
and fungi in the soil.
Almost all plant species produce mycorrhizae.
This plant-fungus symbiosis may have been one of the evolutionary adaptations that
made it possible for plants to colonize land in the first place.
Fossilized roots from some of the earliest land plants include mycorrhizae.
Mycorrhizal fungi are more efficient at absorbing minerals than roots, which may have
helped nourish pioneering plants, especially in the nutrient-poor soils present when
terrestrial ecosystems were young.
Today, the first plants to become established on nutrient-poor soils are usually well
endowed with mycorrhizae.
Mycorrhizae take two major forms: ectomycorrhizae and endomycorrhizae.
In ectomycorrhizae, the mycelium forms a dense sheath over the surface of the root.
Some hyphae grow into the cortex in extracellular spaces between root cells. Hyphae
do not penetrate root cells but form a network in the extracellular spaces to facilitate
nutrient exchange.
The mycelium of ectomycorrhizae extends from the mantle surrounding the root into
the soil, greatly increasing the surface area for water and mineral absorption.
Sexual reproduction is not the sole means by which flowering plants reproduce.
Many species can also reproduce asexually, creating offspring that are genetically identical to them.
The propagation of flowering plants by sexual and asexual reproduction forms the basis of agriculture.
For 10,000 years, plant breeders have altered the traits of a few hundred angiosperm species by artificial selection,
transforming them into today’s crops.
Concept 38.2 After fertilization, ovules develop into seeds and ovaries into fruits
Double fertilization gives rise to the zygote and endosperm.
After landing on a receptive stigma, the pollen grain absorbs moisture and germinates, producing a pollen tube that
extends down the style toward the ovary.
The nucleus of the generative cell divides by mitosis to produce two sperm, the male gametes.
The germinated pollen grain contains the mature male gametophyte.
Directed by a chemical attractant, possibly calcium, the tip of the pollen tube enters the ovary, probes through
the micropyle (a gap in the integuments of the ovule), and discharges two sperm within the embryo sac.
Both sperm fuse with nuclei in the embryo sac.
Biology Chapter Notes
One sperm fertilizes the egg to form the zygote.
The other sperm combines with the two polar nuclei to form a triploid nucleus in the central cell.
This large cell will give rise to the endosperm, a food-storing tissue of the seed.
The union of two sperm cells with different nuclei of the embryo sac is termed double fertilization.
Double fertilization ensures that the endosperm will develop only in ovules where the egg has been fertilized.
This prevents angiosperms from squandering nutrients.
Normally nonreproductive tissues surrounding the embryo have prevented researchers from visualizing
fertilization in plants, but recently, scientists have been able to isolate sperm cells and eggs and observe
fertilization in vitro.
The first cellular event after gamete fusion is an increase in cytoplasmic Ca2+ levels, which also occurs during
animal gamete fusion.
In another similarity to animals, plants establish a block to polyspermy, the fertilization of an egg by more than
one sperm cell.
In plants, this may be through deposition of cell wall material that mechanically impedes sperm.
In maize, this barrier is established within 45 seconds after the initial sperm fusion with the egg.
The ovule develops into a seed containing an embryo and a supply of nutrients.
After double fertilization, the ovule develops into a seed, and the ovary develops into a fruit enclosing the seed(s).
As the embryo develops, the seed stockpiles proteins, oils, and starch.
Initially, these nutrients are stored in the endosperm.
Later in seed development in many species, the storage function is taken over by the swelling storage leaves
(cotyledons) of the embryo itself.
Endosperm development usually precedes embryo development.
After double fertilization, the triploid nucleus of the ovule’s central cell divides, forming a multinucleate
“supercell” having a milky consistency.
It becomes multicellular when cytokinesis partitions the cytoplasm between nuclei.
Cell walls form, and the endosperm becomes solid.
Coconut “milk” is an example of liquid endosperm and coconut “meat” is an example of solid endosperm.
The endosperm is rich in nutrients, which it provides to the developing embryo.
In most monocots and some dicots, the endosperm also stores nutrients that can be used by the seedling after
germination.
In many dicots, the food reserves of the endosperm are completely exported to the cotyledons before the seed
completes its development, and consequently the mature seed lacks endosperm.
The first mitotic division of the zygote is transverse, splitting the fertilized egg into a basal cell and a terminal cell.
The terminal cell gives rise to most of the embryo.
The basal cell continues to divide transversely, producing a thread of cells, the suspensor, which anchors the
embryo to its parent.
The suspensor functions in the transfer of nutrients to the embryo from the parent.
The terminal cell divides several times and forms a spherical proembryo attached to the suspensor.
Cotyledons begin to form as bumps on the proembryo.
A eudicot, with its two cotyledons, is heart-shaped at this stage.
Only one cotyledon develops in monocots.
After the cotyledons appear, the embryo elongates.
Cradled between cotyledons is the embryonic shoot apex with the apical meristem of the embryonic shoot.
At the opposite end of the embryo axis is the apex of the embryonic root, also with a meristem.
Biology Chapter Notes
After the seed germinates, the apical meristems at the tips of the shoot and root sustain primary growth as long as
the plant lives.
During the last stages of maturation, a seed dehydrates until its water content is only about 5–15% of its weight.
The embryo stops growing and becomes dormant until the seed germinates.
The embryo and its food supply are enclosed by a protective seed coat formed by the integuments of the
ovule.
In the seed of a common bean, the embryo consists of an elongate structure, the embryonic axis, attached to fleshy
cotyledons.
Below the point at which the fleshy cotyledons are attached, the embryonic axis is called the hypocotyl; above
it is the epicotyl.
At the tip of the epicotyl is the plumule, consisting of the shoot tip with a pair of miniature leaves.
The hypocotyl terminates in the radicle, or embryonic root.
While the cotyledons of the common bean supply food to the developing embryo, the seeds of some dicots, such as
castor beans, retain their food supply in the endosperm and have cotyledons that are very thin.
The cotyledons will absorb nutrients from the endosperm and transfer them to the embryo when the seed
germinates.
The embryo of a monocot has a single cotyledon.
Members of the grass family, including maize and wheat, have a specialized cotyledon called a scutellum.
The scutellum is very thin, with a large surface area pressed against the endosperm, from which the scutellum
absorbs nutrients during germination.
The embryo of a grass seed is enclosed by two sheathes, a coleorhiza, which covers the young root, and a
coleoptile, which covers the young shoot.
The ovary develops into a fruit adapted for seed dispersal.
As the seeds are developing from ovules, the ovary of the flower is developing into a fruit, which protects the
enclosed seeds and aids in their dispersal by wind or animals.
Fertilization triggers hormonal changes that cause the ovary to begin its transformation into a fruit.
If a flower has not been pollinated, fruit usually does not develop, and the entire flower withers and falls away.
The wall of the ovary becomes the pericarp, the thickened wall of the fruit, while other parts of the flower wither
and are shed.
In some angiosperms, other floral parts contribute to the fruit.
In apples, the fleshy part of the fruit is derived mainly from the swollen receptacle, while the core of the apple
fruit develops from the ovary.
Fruits are classified into several types, depending on their developmental origin.
A typical fruit is derived from a single carpel or several fused carpels and is called a simple fruit.
Some simpler fruits are fleshy, like a peach, while others are dry, like a pea pod.
An aggregate fruit results from a single flower that has more than one carpel, each forming a small fruit.
The fruitlets are clustered together on a single receptacle, like a raspberry.
A multiple fruit develops from an inflorescence, a group of flowers tightly clustered together.
When the walls of the ovaries thicken, they fuse together and form one fruit, as in a pineapple.
The fruit usually ripens about the same time as its seeds are completing their development.
For a dry fruit such as a soybean pod, ripening is a little more than senescence of the fruit tissues, which
allows the fruit to open and release the seeds.
The ripening of fleshy fruits is more elaborate, its steps controlled by the complex interactions of hormones.
Ripening results in an edible fruit that serves as an enticement to the animals that help spread the seeds.
Biology Chapter Notes
The “pulp” of the fruit becomes softer as a result of enzymes digesting components of the cell walls.
Color changes from green to red, orange, or yellow.
The fruit becomes sweeter as organic acids or starch molecules are converted to sugar.
Evolutionary adaptations of seed germination contribute to seedling survival.
As a seed matures, it dehydrates and enters a dormancy phase, a condition of extremely low metabolic rate and a
suspension of growth and development.
Conditions required to break dormancy and resume growth and development vary between species.
Some seeds germinate as soon as they are in a suitable environment.
Others remain dormant until some specific environmental cue causes them to break dormancy.
Seed dormancy increases the chances that germination will occur at a time and place most advantageous to the
seedling.
For example, seeds of many desert plants germinate only after a substantial rainfall, ensuring enough water to
complete development.
Where natural fires are common, many seeds require intense heat to break dormancy, allowing them to take
advantage of new opportunities and open space.
Where winters are harsh, seeds may require extended exposure to cold.
Small seeds such as lettuce require light for germination and break dormancy only if they are buried near the
surface.
Other seeds require a chemical attack or physical abrasion as they pass through an animal’s digestive tract
before they can germinate.
The length of time that a dormant seed remains viable and capable of germinating varies from a few days to
decades or longer.
It depends on the species and on environmental conditions.
Most seeds are durable enough to last for a year or two until conditions are favorable for germination.
Thus, the soil has a pool of nongerminated seeds that may have accumulated for several years.
This is one reason vegetation reappears so rapidly after a fire, drought, flood, or some other environmental
disruption.
Germination of seeds depends on imbibition, the uptake of water due to the low water potential of the dry seed.
This causes the expanding seed to rupture its seed coat and triggers metabolic changes in the embryo that
enable it to resume growth.
Enzymes begin digesting the storage materials of endosperm or cotyledons, and the nutrients are transferred to
the growing regions of the embryo.
The first organ to emerge from the germinating seed is the radicle, the embryonic root.
Next, the shoot tip must break through the soil surface.
In garden beans and many other dicots, a hook forms in the hypocotyl, and growth pushes it aboveground.
Stimulated by light, the hypocotyl straightens, raising the cotyledons and epicotyl.
As it rises into the air, the epicotyl spreads its first foliage leaves (true leaves).
These foliage leaves expand, become green, and begin making food by photosynthesis.
After the cotyledons have transferred all their nutrients to the developing plant, they shrivel and fall off the
seedling.
Corn and other grasses, which are monocots, use a different method for breaking ground when they germinate.
The coleoptile, the sheath enclosing and protecting the embryonic shoot, pushes upward through the soil and
into the air.
The shoot tip then grows straight up through the tunnel provided by the tubular coleoptile.
At every stage in the life of a plant, sensitivity to the environment and coordination of
responses are evident.
One part of a plant can send signals to other parts.
Plants can sense gravity and the direction of light.
A plant’s morphology and physiology are constantly tuned to its variable surroundings
by complex interactions between environmental stimuli and internal signals.
At the organismal level, plants and animals respond to environmental stimuli by very
different means.
Animals, being mobile, respond mainly by behavioral mechanisms, moving toward
positive stimuli and away from negative stimuli.
Rooted in one location for life, a plant generally responds to environmental cues by
adjusting its pattern of growth and development.
As a result, plants of the same species vary in body form much more than do
animals of the same species.
At the cellular level, plants and all other eukaryotes are surprisingly similar in their
signaling mechanisms.
Concept 39.4 Plants respond to a wide variety of stimuli other than light
Because of their immobility, plants must adjust to a wide range of environmental
circumstances through developmental and physiological mechanisms.
While light is one important environmental cue, other environmental stimuli also
influence plant development and physiology.
Plants respond to environmental stimuli through a combination of developmental and
physiological mechanisms.
Both the roots and shoots of plants respond to gravity, or gravitropism, although in
diametrically different ways.
Roots demonstrate positive gravitropism, and shoots exhibit negative gravitropism.
Gravitropism ensures that the root grows in the soil and that the shoot reaches sunlight
regardless of how a seed happens to be oriented when it lands.
Auxin plays a major role in gravitropic responses.
Plants may tell up from down by the settling of statoliths, specialized plastids containing
dense starch grains, to the lower portions of cells.
In one hypothesis, the aggregation of statoliths at low points in cells of the root cap
triggers the redistribution of calcium, which in turn causes lateral transport of auxin
within the root.
The high concentrations of auxin on the lower side of the zone of elongation inhibits
cell elongation, slowing growth on that side and curving the root downward.
Experiments with Arabidopsis and tobacco mutants have demonstrated the importance of
“falling statoliths” in root gravitropism, but these have also indicated that other factors or
organelles may be involved.
Mutants lacking statoliths have a slower response than wild-type plants.
Concept 40.1 Physical laws and the environment constrain animal size and shape
An animal’s size and shape, features often called “body plans” or “designs,” are
fundamental aspects of form and function that significantly affect the way an animal
interacts with its environment.
The terms plan and design do not mean that animal body forms are products of
conscious invention.
The body plan or design of an animal results from a pattern of development
programmed by the genome, itself the product of millions of years of evolution due to
natural selection.
Physical requirements constrain what natural selection can “invent.”
An animal such as the mythical winged dragon cannot exist. No animal as large as a
dragon could generate enough lift to take off and fly.
Similarly, the laws of hydrodynamics constrain the shapes that are possible for aquatic
organisms that swim very fast.
Tunas, sharks, penguins, dolphins, seals, and whales are all fast swimmers.
Biology Chapter Notes
All have the same basic fusiform shape, tapered at both ends.
This shape minimizes drag in water, which is about a thousand times denser than air.
The similar forms of speedy fishes, birds, and marine mammals are a consequence of
convergent evolution in the face of the universal laws of hydrodynamics.
Convergence occurs because natural selection shapes similar adaptations when diverse
organisms face the same environmental challenge, such as the resistance of water to
fast travel.
Body size and shape affect interactions with the environment.
An animal’s size and shape have a direct effect on how the animal exchanges energy and
materials with its surroundings.
As a requirement for maintaining the fluid integrity of the plasma membrane of its cells,
an animal’s body must be arranged so that all of its living cells are bathed in an aqueous
medium.
Exchange with the environment occurs as dissolved substances diffuse and are transported
across the plasma membranes between the cells and their aqueous surroundings.
For example, a single-celled protist living in water has a sufficient surface area of
plasma membrane to service its entire volume of cytoplasm.
Surface-to-volume ratio is one of the physical constraints on the size of single-celled
protists.
Multicellular animals are composed of microscopic cells, each with its own plasma
membrane that acts as a loading and unloading platform for a modest volume of
cytoplasm.
This only works if all the cells of the animal have access to a suitable aqueous
environment.
For example, a hydra, built as a sac, has a body wall only two cell layers thick.
Because its gastrovascular cavity opens to the exterior, both outer and inner layers of
cells are bathed in water.
Another way to maximize exposure to the surrounding medium is to have a flat body.
For instance, a parasitic tapeworm may be several meters long, but because it is very
thin, most of its cells are bathed in the intestinal fluid of the worm’s vertebrate host
from which it obtains nutrients.
While two-layered sacs and flat shapes are designs that put a large surface area in contact
with the environment, these solutions do not permit much complexity in internal
organization.
Most animals are more complex and are made up of compact masses of cells, producing
outer surfaces that are relatively small compared to the animal’s volume.
Most organisms have extensively folded or branched internal surfaces specialized for
exchange with the environment.
The circulatory system shuttles material among all the exchange surfaces within the
animal.
Although exchange with the environment is a problem for animals whose cells are mostly
internal, complex forms have distinct benefits.
A specialized outer covering can protect against predators; large muscles can enable
rapid movement; and internal digestive organs can break down food gradually,
controlling the release of stored energy.
Concept 40.2 Animal form and function are correlated at all levels of organization
Life is characterized by hierarchical levels of organization, each with emergent properties.
Animals are multicellular organisms with their specialized cells grouped into tissues.
In most animals, combinations of various tissues make up functional units called organs,
and groups of organs work together as organ systems.
For example, the human digestive system consists of a stomach, small intestine, large
intestine, and several other organs, each a composite of different tissues.
Tissues are groups of cells with a common structure and function.
Different types of tissues have different structures that are suited to their functions.
A tissue may be held together by a sticky extracellular matrix that coats the cells or
weaves them together in a fabric of fibers.
The term tissue is from a Latin word meaning “weave.”
Tissues are classified into four main categories: epithelial tissue, connective tissue,
nervous tissue, and muscle tissue.
Occurring in sheets of tightly packed cells, epithelial tissue covers the outside of the body
and lines organs and cavities within the body.
The cells of an epithelium are closely joined and in many epithelia, the cells are riveted
together by tight junctions.
The epithelium functions as a barrier protecting against mechanical injury, invasive
microorganisms, and fluid loss.
The cells at the base of an epithelial layer are attached to a basement membrane, a dense
mat of extracellular matrix.
The free surface of the epithelium is exposed to air or fluid.
Some epithelia, called glandular epithelia, absorb or secrete chemical solutions.
The glandular epithelia that line the lumen of the digestive and respiratory tracts form
a mucous membrane that secretes a slimy solution called mucus that lubricates the
surface and keeps it moist.
Epithelia are classified by the number of cell layers and the shape of the cells on the free
surface.
A simple epithelium has a single layer of cells, and a stratified epithelium has
multiple tiers of cells.
A “pseudostratified” epithelium is single-layered but appears stratified because the
cells vary in length.
The shapes of cells on the exposed surface may be cuboidal (like dice), columnar (like
bricks on end), or squamous (flat like floor tiles).
Connective tissue functions mainly to bind and support other tissues.
Connective tissues have a sparse population of cells scattered through an extracellular
matrix.
Concept 40.3 Animals use the chemical energy in food to sustain form and function
All organisms require chemical energy for growth, physiological processes, maintenance
and repair, regulation, and reproduction.
Plants use light energy to build energy-rich organic molecules from water and CO2,
and then they use those organic molecules for fuel.
In contrast, animals are heterotrophs and must obtain their chemical energy in food,
which contains organic molecules synthesized by other organisms.
The flow of energy through an animal—its bioenergetics—ultimately limits the animal’s
behavior, growth, and reproduction and determines how much food it needs.
Studying an animal’s bioenergetics tells us a great deal about the animal’s adaptations.
Biology Chapter Notes
Food is digested by enzymatic hydrolysis, and energy-containing food molecules are
absorbed by body cells.
Most fuel molecules are used to generate ATP by the catabolic processes of cellular
respiration and fermentation.
The chemical energy of ATP powers cellular work, enabling cells, organs, and organ
systems to perform the many functions that keep an animal alive.
Since the production and use of ATP generates heat, an animal continuously loses heat
to its surroundings.
After energetic needs of staying alive are met, any remaining food molecules can be used
in biosynthesis.
This includes body growth and repair; synthesis of storage material such as fat; and
production of reproductive structures, including gametes.
Biosynthesis requires both carbon skeletons for new structures and ATP to power their
assembly.
Metabolic rate provides clues to an animal’s bioenergetic “strategy.”
The amount of energy an animal uses in a unit of time is called its metabolic rate—the
sum of all the energy-requiring biochemical reactions occurring over a given time
interval.
Energy is measured in calories (cal) or kilocalories (kcal).
A kilocalorie is 1,000 calories.
The term Calorie, with a capital C, as used by many nutritionists, is actually a
kilocalorie.
Metabolic rate can be determined several ways.
Because nearly all the chemical energy used in cellular respiration eventually appears as
heat, metabolic rate can be measured by monitoring an animal’s heat loss.
A small animal can be placed in a calorimeter, which is a closed, insulated chamber
equipped with a device that records the animal’s heat loss.
A more indirect way to measure metabolic rate is to determine the amount of oxygen
consumed or carbon dioxide produced by an animal’s cellular respiration.
These devices may measure changes in oxygen consumed or carbon dioxide produced
as activity changes.
Over long periods, the rate of food consumption and the energy content of food can be
used to estimate metabolic rate.
A gram of protein or carbohydrate contains about 4.5–5 kcal, and a gram of fat
contains 9 kcal.
This method must account for the energy in food that cannot be used by the animal
(the energy lost in feces and urine).
There are two basic bioenergetic “strategies” used by animals.
Birds and mammals are mainly endothermic, maintaining their body temperature
within a narrow range by heat generated by metabolism.
Endothermy is a high-energy strategy that permits intense, long-duration activity of
a wide range of environmental temperatures.
Most fishes, amphibians, reptiles, and invertebrates are ectothermic, meaning they gain
their heat mostly from external sources.
Concept 40.4 Many animals regulate their internal environment within relatively
narrow limits
More than a century ago, physiologist Claude Bernard made the distinction between
external environments surrounding an animal and the internal environment in which the
cells of the animal actually live.
The internal environment of vertebrates is called the interstitial fluid.
This fluid exchanges nutrients and wastes with blood contained in microscopic vessels
called capillaries.
Bernard also recognized that many animals tend to maintain relatively constant conditions
in their internal environment, even when the external environment changes.
While a pond-dwelling hydra is powerless to affect the temperature of the fluid that
bathes its cells, the human body can maintain its “internal pond” at a more or less
constant temperature of about 37°C.
Similarly, our bodies control the pH of our blood and interstitial fluid to within a tenth
of a pH unit of 7.4.
The amount of sugar in our blood does not fluctuate for long from a concentration of
about 90 mg of glucose per 100 mL of blood.
There are times during the course of the development of an animal when major changes in
the internal environment are programmed to occur.
For example, the balance of hormones in human blood is altered radically during
puberty and pregnancy.
Still, the stability of the internal environment is remarkable.
Biology Chapter Notes
Today, Bernard’s “constant internal milieu” is incorporated into the concept of
homeostasis, which means “steady state,” or internal balance.
Actually the internal environment of an animal always fluctuates slightly.
Homeostasis is a dynamic state, an interplay between outside forces that tend to
change the internal environment and internal control mechanisms that oppose such
changes.
Animals may be regulators or conformers for a particular environmental variable.
Regulating and conforming are two extremes in how animals deal with environmental
fluctuations.
An animal is a regulator for a particular environmental variable if it uses internal control
mechanisms to moderate internal change while external conditions fluctuate.
For example, a freshwater fish maintains a stable internal concentration of solutes in its
blood that is higher than the water in which it lives.
An animal is a conformer for a particular environmental variable if it allows its internal
conditions to vary as external conditions fluctuate.
For example, many marine invertebrates live in environments where solute
concentration (salinity) is relatively stable.
Unlike freshwater fishes, most marine invertebrates do not regulate their internal solute
concentration, but rather conform to the external environment.
No organism is a perfect regulator or conformer.
An animal may maintain homeostasis while regulating some internal conditions and
allowing others to conform to the environment.
For example, most freshwater fishes regulate their internal solute concentration but
allow their internal temperature to conform to external water temperature.
Homeostasis depends on feedback circuits.
Any homeostatic control system has three functional components: a receptor, a control
center, and an effector.
The receptor detects a change in some variable in the animal’s internal environment,
such as a change in temperature.
The control center processes the information it receives from the receptor and directs
an appropriate response by the effector.
One type of control circuit, a negative-feedback system, can control the temperature in a
room.
In this case, the control center, called a thermostat, also contains the receptor, a
thermometer.
When room temperature falls, the thermostat switches on the heater, the effector.
In such a negative-feedback system, a change in the variable being monitored triggers the
control mechanism to counteract further change in the same direction.
Owing to a time lag between receptor and response, the variable drifts slightly above
and below the set point, but fluctuations are moderate.
Negative-feedback mechanisms prevent small changes from becoming too large.
Most homeostatic mechanisms in animals operate on this principle of negative feedback.
Human body temperature is kept close to a set point of 37°C by the cooperation of
several negative-feedback circuits.
All animals eat other organisms—dead or alive, whole or by the piece (including
parasites).
In general, animals fit into one of three dietary categories.
o Herbivores, such as gorillas, cows, hares, and many snails, eat mainly autotrophs
(plants and algae).
o Carnivores, such as sharks, hawks, spiders, and snakes, eat other animals.
o Omnivores, such as cockroaches, bears, raccoons, and humans, consume animal and
plant or algal matter.
Humans evolved as hunters, scavengers, and gatherers.
While the terms herbivore, carnivore, and omnivore represent the kinds of food that an
animal usually eats, most animals are opportunistic, eating foods that are outside their
main dietary category when these foods are available.
For example, cattle and deer, which are herbivores, may occasionally eat small animals
or bird eggs.
Most carnivores obtain some nutrients from plant materials that remain in the digestive
tract of the prey that they eat.
All animals consume bacteria along with other types of food.
For any animal, a nutritionally adequate diet must satisfy three nutritional needs:
o A balanced diet must provide fuel for cellular work.
o It must supply the organic raw materials needed to construct organic molecules.
o Essential nutrients that the animal cannot make from raw materials must be provided in
its food.
Concept 41.2 An animal’s diet must supply carbon skeletons and essential nutrients
In addition to fuel for ATP production, an animal’s diet must supply all the raw materials
for biosynthesis.
This requires organic precursors (carbon skeletons) from its food.
Concept 41.3 The main stages of food processing are ingestion, digestion, absorption,
and elimination
Ingestion, the act of eating, is only the first stage of food processing.
Food is “packaged” in bulk form and contains very complex arrays of molecules,
including large polymers and various substances that may be difficult to process or
even toxic.
Animals cannot use macromolecules like proteins, fats, and carbohydrates in the form of
starch or other polysaccharides.
First, polymers are too large to pass through membranes and enter the cells of the
animal.
Second, the macromolecules that make up an animal are not identical to those of its
food.
In building their macromolecules, however, all organisms use common monomers.
For example, soybeans, fruit flies, and humans all assemble their proteins from the
same 20 amino acids.
Digestion, the second stage of food processing, is the process of breaking food down into
molecules small enough for the body to absorb.
Digestion cleaves macromolecules into their component monomers, which the animal
then uses to make its own molecules or as fuel for ATP production.
Polysaccharides and disaccharides are split into simple sugars.
Fats are digested to glycerol and fatty acids.
Proteins are broken down into amino acids.
Nucleic acids are cleaved into nucleotides.
Digestion reverses the process that a cell uses to link together monomers to form
macromolecules.
Rather than removing a molecule of water for each new covalent bond formed,
digestion breaks bonds with the addition of water via enzymatic hydrolysis.
Biology Chapter Notes
A variety of hydrolytic enzymes catalyze the digestion of each of the classes of
macromolecules found in food.
Chemical digestion is usually preceded by mechanical fragmentation of the food—by
chewing, for instance.
Breaking food into smaller pieces increases the surface area exposed to digestive juices
containing hydrolytic enzymes.
After the food is digested, the animal’s cells take up small molecules such as amino acids
and simple sugars from the digestive compartment, a process called absorption.
During elimination, undigested material passes out of the digestive compartment.
Digestion occurs in specialized compartments.
To avoid digesting their own cells and tissues, most organisms conduct digestion in
specialized compartments.
The simplest digestive compartments are food vacuoles, organelles in which hydrolytic
enzymes break down food without digesting the cell’s own cytoplasm, a process termed
intracellular digestion.
This process begins after a cell has engulfed food by phagocytosis or pinocytosis.
Newly formed food vacuoles fuse with lysosomes, which are organelles containing
hydrolytic enzymes.
Later the vacuole fuses with an anal pore, and its contents are eliminated.
In most animals, at least some hydrolysis occurs by extracellular digestion, the
breakdown of food outside cells.
Extracellular digestion occurs within compartments that are continuous with the
outside of the animal’s body.
This enables organisms to devour much larger prey than can be ingested by
phagocytosis and digested intracellularly.
Many animals with simple body plans, such as cnidarians and flatworms, have digestive
sacs with single openings, called gastrovascular cavities.
These cavities function in both digestion and distribution of nutrients throughout the
body.
For example, the cnidarians called hydras capture their prey with nematocysts and use
tentacles to stuff the prey through the mouth into the gastrovascular cavity.
The prey is then partially digested by enzymes secreted by specialized gland cells
of the gastrodermis.
Nutritive muscular cells in the gastrodermis engulf the food particles.
Most of the actual hydrolysis of macromolecules occurs intracellularly.
Undigested materials are eliminated through the mouth.
In contrast to cnidarians and flatworms, most animals have digestive tubes extending
between a mouth and anus.
These tubes are called complete digestive tracts or alimentary canals.
Because food moves in one direction, the tube can be organized into specialized
regions that carry out digestion and nutrient absorption in a stepwise fashion.
In addition, animals with alimentary canals can eat more food before the earlier meal is
completely digested.
Every organism must exchange materials and energy with its environment, and this
exchange ultimately occurs at the cellular level.
Cells live in aqueous environments.
The resources that they need, such as nutrients and oxygen, move across the plasma
membrane to the cytoplasm.
Metabolic wastes, such as carbon dioxide, move out of the cell.
Most animals have organ systems specialized for exchanging materials with the
environment, and many have an internal transport system that conveys fluid (blood or
interstitial fluid) throughout the body.
For aquatic organisms, structures such as gills present an expansive surface area to the
outside environment.
Oxygen dissolved in the surrounding water diffuses across the thin epithelium covering
the gills and into a network of tiny blood vessels (capillaries).
At the same time, carbon dioxide diffuses out into the water.
Concept 43.3 Humoral and cell-mediated immunity defend against different types of
threats
The immune system can mount two types of responses to antigens: a humoral response
and a cell-mediated response.
Humoral immunity involves B cell activation and clonal selection and results in the
production of antibodies that circulate in the blood plasma and lymph.
Circulating antibodies defend mainly against free bacteria, toxins, and viruses in the
body fluids.
In cell-mediated immunity, activation and clonal selection of cytotoxic T
lymphocytes allows these cells to directly destroy certain target cells, including
“nonself” cancer and transplant cells.
Concept 43.4 The immune system’s ability to distinguish self from nonself limits tissue
transplantation
In addition to attacking pathogens, the immune system will also attack cells from other
individuals.
For example, a skin graft from one person to a nonidentical individual will look
healthy for a day or two, but it will then be destroyed by immune responses.
Interestingly, a pregnant woman does not reject the fetus as a foreign body.
Apparently, the structure of the placenta is the key to this acceptance.
One source of potential problems with blood transfusions is an immune reaction from
individuals with incompatible blood types.
In the ABO blood groups, an individual with type A blood has A antigens on the
surface of red blood cells.
This is not recognized as an antigen by the “owner,” but it can be identified as
foreign if placed in the body of another individual.
B antigens are found on type B red blood cells.
Both A and B antigens are found on type AB red blood cells.
Neither antigen is found on type O red blood cells.
A person with type A blood already has antibodies to the B antigen, even if the person has
never been exposed to type B blood.
These antibodies arise in response to bacteria (normal flora) that have epitopes very
similar to blood group antigens.
Thus, an individual with type A blood does not make antibodies to A-like bacterial
epitopes—these are considered self—but that person does make antibodies to B-like
bacterial epitopes.
If a person with type A blood receives a transfusion of type B blood, the preexisting
anti-B antibodies will induce an immediate and devastating transfusion reaction.
Because blood group antigens are polysaccharides, they induce T-independent responses,
which elicit no memory cells.
Each response is like a primary response, and it generates IgM anti-blood-group
antibodies, not IgG.
Biology Chapter Notes
This is fortunate, because IgM antibodies do not cross the placenta, where they may
harm a developing fetus with a blood type different from its mother’s.
However, another blood group antigen, the Rh factor, can cause mother-fetus problems
because antibodies produced for it are IgG.
This situation arises when a mother that is Rh-negative (lacks the Rh factor) has a fetus
that is Rh-positive, having inherited the factor from the father.
If small amounts of fetal blood cross the placenta late in pregnancy or during delivery,
the mother mounts a humoral response against the Rh factor.
The danger occurs in subsequent Rh-positive pregnancies, when the mother’s Rh-
specific memory B cells produce IgG antibodies that can cross the placenta and destroy
the red blood cells of the fetus.
To prevent this, the mother is injected with anti-Rh antibodies after delivering her first
Rh-positive baby.
She is, in effect, passively immunized (artificially) to eliminate the Rh antigen before
her own immune system responds and generates immunological memory against the
Rh factor, endangering her future Rh-positive babies.
Major histocompatibility complex (MHC) molecules are responsible for stimulating
rejection of tissue grafts and organ transplants.
Because MHC creates a unique protein fingerprint for each individual, foreign MHC
molecules are antigenic, inducing immune responses against the donated tissue or
organ.
To minimize rejection, attempts are made to match MHC of tissue donor and recipient
as closely as possible.
In the absence of identical twins, siblings usually provide the closest tissue-type
match.
In addition to MHC matching, various medicines are used to suppress the immune
response to the transplant.
However, this strategy leaves the recipient more susceptible to infection and cancer
during the course of treatment.
More selective drugs, which suppress helper T cell activation without crippling
nonspecific defense or T-independent humoral responses, have greatly improved the
success of organ transplants.
In bone marrow transplants, it is the graft itself, rather than the recipient, which is the
source of potential immune rejection.
Bone marrow transplants are used to treat leukemia and other cancers as well as
various hematological diseases.
Prior to the transplant, the recipient is typically treated with irradiation to eliminate the
recipient’s immune system, eliminating all abnormal cells and leaving little chance of
graft rejection.
However, the donated marrow, containing lymphocytes, may react against the
recipient, producing graft versus host reaction, unless well matched.
Concept 44.1 Osmoregulation balances the uptake and loss of water and solutes
All animals face the same central problem of osmoregulation.
Over time, the rates of water uptake and loss must balance.
Animal cells—which lack cell walls—swell and burst if there is a continuous net
uptake of water, or shrivel and die if there is a substantial net loss of water.
Water enters and leaves cells by osmosis, the movement of water across a selectively
permeable membrane.
Osmosis occurs whenever two solutions separated by a membrane differ in osmotic
pressure, or osmolarity (moles of solute per liter of solution).
The unit of measurement of osmolarity is milliosmoles per liter (mosm/L).
1 mosm/L is equivalent to a total solute concentration of 10−3 M.
The osmolarity of human blood is about 300 mosm/L, while seawater has an
osmolarity of about 1,000 mosm/L.
If two solutions separated by a selectively permeable membrane have the same
osmolarity, they are said to be isoosmotic.
There is no net movement of water by osmosis between isoosmotic solutions, although
water molecules do cross at equal rates in both directions.
When two solutions differ in osmolarity, the one with the greater concentration of
solutes is referred to as hyperosmotic, and the more dilute solution is hypoosmotic.
Water flows by osmosis from a hypoosmotic solution to a hyperosmotic one.
Osmoregulators expend energy to control their internal osmolarity; osmoconformers are
isoosmotic with their surroundings.
There are two basic solutions to the problem of balancing water gain with water loss.
One—available only to marine animals—is to be isoosmotic to the surroundings as an
osmoconformer.
Although they do not compensate for changes in external osmolarity,
osmoconformers often live in water that has a very stable composition and, hence,
they have a very constant internal osmolarity.
Biology Chapter Notes
In contrast, an osmoregulator is an animal that must control its internal osmolarity
because its body fluids are not isoosmotic with the outside environment.
An osmoregulator must discharge excess water if it lives in a hypoosmotic
environment or take in water to offset osmotic loss if it inhabits a hyperosmotic
environment.
Osmoregulation enables animals to live in environments that are uninhabitable to
osmoconformers, such as freshwater and terrestrial habitats.
It also enables many marine animals to maintain internal osmolarities different from
that of seawater.
Whenever animals maintain an osmolarity difference between the body and the external
environment, osmoregulation has an energy cost.
Because diffusion tends to equalize concentrations in a system, osmoregulators must
expend energy to maintain the osmotic gradients via active transport.
The energy costs depend mainly on how different an animal’s osmolarity is from its
surroundings, how easily water and solutes can move across the animal’s surface, and
how much membrane-transport work is required to pump solutes.
Osmoregulation accounts for nearly 5% of the resting metabolic rate of many marine
and freshwater bony fishes.
Most animals, whether osmoconformers or osmoregulators, cannot tolerate substantial
changes in external osmolarity and are said to be stenohaline.
In contrast, euryhaline animals—which include both some osmoregulators and
osmoconformers—can survive large fluctuations in external osmolarity.
For example, various species of salmon migrate back and forth between freshwater and
marine environments.
The food fish, tilapia, is an extreme example, capable of adjusting to any salt
concentration between freshwater and 2,000 mosm/L, twice that of seawater.
Most marine invertebrates are osmoconformers.
Their osmolarity is the same as seawater.
However, they differ considerably from seawater in their concentrations of most
specific solutes.
Thus, even an animal that conforms to the osmolarity of its surroundings does regulate
its internal composition.
Marine vertebrates and some marine invertebrates are osmoregulators.
For most of these animals, the ocean is a strongly dehydrating environment because it
is much saltier than internal fluids, and water is lost from their bodies by osmosis.
Marine bony fishes, such as cod, are hypoosmotic to seawater and constantly lose
water by osmosis and gain salt by diffusion and from the food they eat.
The fishes balance water loss by drinking seawater and actively transporting chloride
ions out through their skin and gills.
Sodium ions follow passively.
They produce very little urine.
Marine sharks and most other cartilaginous fishes (chondrichthyans) use a different
osmoregulatory “strategy.”
Like bony fishes, salts diffuse into the body from seawater, and these salts are removed
by the kidneys, a special organ called the rectal gland, or in feces.
Concept 44.2 An animal’s nitrogenous wastes reflect its phylogeny and habitat
Because most metabolic wastes must be dissolved in water when they are removed from
the body, the type and quantity of waste products may have a large impact on water
balance.
Nitrogenous breakdown products of proteins and nucleic acids are among the most
important wastes in terms of their effect on osmoregulation.
During their breakdown, enzymes remove nitrogen in the form of ammonia, a small
and very toxic molecule.
Biology Chapter Notes
Some animals excrete ammonia directly, but many species first convert the ammonia
to other compounds that are less toxic but costly to produce.
Animals that excrete nitrogenous wastes as ammonia need access to lots of water.
This is because ammonia is very soluble but can be tolerated only at very low
concentrations.
Therefore, ammonia excretion is most common in aquatic species.
Many invertebrates release ammonia across the whole body surface.
In fishes, most of the ammonia is lost as ammonium ions (NH4+) at the gill epithelium.
Freshwater fishes +are able to exchange NH4+ for Na+ from the environment, which
helps maintain Na concentrations in body fluids.
Ammonia excretion is much less suitable for land animals.
Because ammonia is so toxic, it can be transported and excreted only in large volumes
of very dilute solutions.
Most terrestrial animals and many marine organisms (which tend to lose water to their
environment by osmosis) do not have access to sufficient water.
Instead, mammals, most adult amphibians, sharks, and some marine bony fishes and
turtles excrete mainly urea.
Urea is synthesized in the liver by combining ammonia with carbon dioxide and is
excreted by the kidneys.
The main advantage of urea is its low toxicity, about 100,000 times less than that of
ammonia.
Urea can be transported and stored safely at high concentrations.
This reduces the amount of water needed for nitrogen excretion when releasing a
concentrated solution of urea rather than a dilute solution of ammonia.
The main disadvantage of urea is that animals must expend energy to produce it from
ammonia.
In weighing the relative advantages of urea versus ammonia as the form of nitrogenous
waste, it makes sense that many amphibians excrete mainly ammonia when they are
aquatic tadpoles.
They switch largely to urea when they are land-dwelling adults.
Land snails, insects, birds, and many reptiles excrete uric acid as the main nitrogenous
waste.
Like urea, uric acid is relatively nontoxic.
But unlike either ammonia or urea, uric acid is largely insoluble in water and can be
excreted as a semisolid paste with very little water loss.
While saving even more water than urea, it is even more energetically expensive to
produce.
Uric acid and urea represent different adaptations for excreting nitrogenous wastes with
minimal water loss.
Mode of reproduction appears to have been important in choosing among these
alternatives.
Soluble wastes can diffuse out of a shell-less amphibian egg (ammonia) or be carried
away by the mother’s blood in a mammalian embryo (urea).
However, the shelled eggs of birds and reptiles are not permeable to liquids, which
means that soluble nitrogenous wastes trapped within the egg could accumulate to
dangerous levels.
Biology Chapter Notes
Even urea is toxic at very high concentrations.
Uric acid precipitates out of solution and can be stored within the egg as a harmless
solid left behind when the animal hatches.
The type of nitrogenous waste also depends on habitat.
For example, terrestrial turtles (which often live in dry areas) excrete mainly uric acid,
while aquatic turtles excrete both urea and ammonia.
In some species, individuals can change their nitrogenous wastes when environmental
conditions change.
For example, certain tortoises that usually produce urea shift to uric acid when
temperature increases and water becomes less available.
Excretion of nitrogenous wastes is a good illustration of how response to the environment
occurs on two levels.
Over generations, evolution determines the limits of physiological responses for a
species.
During their lives, individual organisms make adjustments within these evolutionary
constraints.
The amount of nitrogenous waste produced is coupled to the energy budget and depends
on how much and what kind of food an animal eats.
Because they use energy at high rates, endotherms eat more food—and thus produce
more nitrogenous wastes—per unit volume than ectotherms.
Carnivores (which derive much of their energy from dietary proteins) excrete more
nitrogen than animals that obtain most of their energy from lipids or carbohydrates.
Concept 44.4 Nephrons and associated blood vessels are the functional units of the
mammalian kidney
Mammals have a pair of bean-shaped kidneys.
Each kidney is supplied with blood by a renal artery and drained by a renal vein.
In humans, the kidneys account for less than 1% of body weight, but they receive
about 20% of resting cardiac output.
Urine exits each kidney through a duct called the ureter, and both ureters drain through a
common urinary bladder.
During urination, urine is expelled from the urinary bladder through a tube called the
urethra, which empties to the outside near the vagina in females or through the penis
in males.
Sphincter muscles near the junction of the urethra and the bladder control urination.
The mammalian kidney has two distinct regions, an outer renal cortex and an inner renal
medulla.
Both regions are packed with microscopic excretory tubules, nephrons, and their
associated blood vessels.
Each nephron consists of a single long tubule and a ball of capillaries, called the
glomerulus.
The blind end of the tubule forms a cup-shaped swelling, called Bowman’s capsule,
that surrounds the glomerulus.
Each human kidney contains about a million nephrons, with a total tubule length of 80
km.
Filtration occurs as blood pressure forces fluid from the blood in the glomerulus into the
lumen of Bowman’s capsule.
The porous capillaries, along with specialized capsule cells called podocytes, are
permeable to water and small solutes but not to blood cells or large molecules such as
plasma proteins.
The filtrate in Bowman’s capsule contains salt, glucose, amino acids, vitamins,
nitrogenous wastes such as urea, and other small molecules.
From Bowman’s capsule, the filtrate passes through three regions of the nephron: the
proximal tubule; the loop of Henle, a hairpin turn with a descending limb and an
ascending limb; and the distal tubule.
The distal tubule empties into a collecting duct, which receives processed filtrate from
many nephrons.
The many collecting ducts empty into the renal pelvis, which is drained by the ureter.
In the human kidney, about 80% of the nephrons, the cortical nephrons, have reduced
loops of Henle and are almost entirely confined to the renal cortex.
The other 20%, the juxtamedullary nephrons, have well-developed loops that extend
deeply into the renal medulla.
Only mammals and birds have juxtamedullary nephrons; the nephrons of other
vertebrates lack loops of Henle.
Concept 44.5 The mammalian kidney’s ability to conserve water is a key terrestrial
adaptation
The osmolarity of human blood is about 300 mosm/L, but the kidney can excrete urine up
to four times as concentrated—about 1,200 mosm/L.
At an extreme of water conservation, Australian hopping mice, which live in desert
regions, can produce urine concentrated to 9,300 mosm/L—25 times as concentrated
as their body fluid.
In a mammalian kidney, the cooperative action and precise arrangement of the loops of
Henle and the collecting ducts are largely responsible for the osmotic gradients that
concentrate the urine.
Biology Chapter Notes
In addition, the maintenance of osmotic differences and the production of
hyperosmotic urine are only possible because considerable energy is expended by the
active transport of solutes against concentration gradients.
In essence, the nephrons can be thought of as tiny energy-consuming machines whose
function is to produce a region of high osmolarity in the kidney, which can then extract
water from the urine in the collecting duct.
The two primary solutes in this osmolarity gradient are NaCl and urea.
The juxtamedullary nephrons, which maintain an osmotic gradient in the kidney and use
that gradient to excrete a hyperosmotic urine, are the key to understanding the physiology
of the mammalian kidney as a water-conserving organ.
Filtrate passing from Bowman’s capsule to the proximal tubule has an osmolarity of
about 300 mosm/L.
As the filtrate flows through the proximal tubule in the renal cortex, large amounts of
water and salt are reabsorbed.
The volume of the filtrate decreases substantially, but its osmolarity remains about the
same.
The ability of the mammalian kidney to convert interstitial fluid at 300 mosm/L to 1,200
mosm/L as urine depends on a countercurrent multiplier between the ascending and
descending limbs of the loop of Henle.
As the filtrate flows from the cortex to the medulla in the descending limb of the loop of
Henle, water leaves the tubule by osmosis.
The osmolarity of the filtrate increases as solutes, including NaCl, become more
concentrated.
The highest molarity occurs at the elbow of the loop of Henle.
This maximizes the diffusion of salt out of the tubule as the filtrate rounds the curve
and enters the ascending limb, which is permeable to salt but not to water.
The descending limb produces progressively saltier filtrate, and the ascending limb
exploits this concentration of NaCl to help maintain a high osmolarity in the interstitial
fluid of the renal medulla.
The loop of Henle has several qualities of a countercurrent system.
Although the two limbs of the loop are not in direct contact, they are close enough to
exchange substances through the interstitial fluid.
The nephron can concentrate salt in the inner medulla largely because exchange
between opposing flows in the descending and ascending limbs overcomes the
tendency for diffusion to even out salt concentrations throughout the kidney’s
interstitial fluid.
The vasa recta is also a countercurrent system, with descending and ascending vessels
carrying blood in opposite directions through the kidney’s osmolarity gradient.
As the descending vessel conveys blood toward the inner medulla, water is lost from
the blood and NaCl diffuses into it.
These fluxes are reversed as blood flows back toward the cortex in the ascending
vessel.
Thus, the vasa recta can supply the kidney with nutrients and other important
substances without interfering with the osmolarity gradient necessary to excrete a
hyperosmotic urine.
The countercurrent-like characteristics of the loop of Henle and the vasa recta maintain
the steep osmotic gradient between the medulla and the cortex.
Concept 44.6 Diverse adaptations of the vertebrate kidney have evolved in different
environments
Variations in nephron structure and function equip the kidneys of different vertebrates for
osmoregulation in their various habitats.
Mammals that excrete the most hyperosmotic urine, such as hopping mice and other
desert mammals, have exceptionally long loops of Henle.
This maintains steep osmotic gradients, resulting in very concentrated urine.
In contrast, beavers, which rarely face problems of dehydration, have nephrons with
short loops, resulting in a much lower ability to concentrate urine.
Birds, like mammals, have kidneys with juxtamedullary nephrons that specialize in
conserving water.
However, the nephrons of birds have much shorter loops of Henle than do mammalian
nephrons.
Biology Chapter Notes
Bird kidneys cannot concentrate urine to the osmolarities achieved by mammalian
kidneys.
The main water conservation adaptation of birds is the use of uric acid as the nitrogen
excretion molecule.
The kidneys of other reptiles, having only cortical nephrons, produce urine that is, at
most, isoosmotic to body fluids.
However, the epithelium of the cloaca helps conserve fluid by reabsorbing some of the
water present in urine and feces.
Also, like birds, most other terrestrial reptiles excrete nitrogenous wastes as uric acid.
In contrast to mammals and birds, a freshwater fish must excrete excess water because the
animal is hyperosmotic to its surroundings.
Instead of conserving water, the nephrons produce a large volume of very dilute urine.
Freshwater fishes conserve salts by reabsorption of ions from the filtrate in the
nephrons.
Amphibian kidneys function much like those of freshwater fishes.
When in fresh water, the skin of the frog accumulates certain salts from the water by
active transport, and the kidneys excrete dilute urine.
On land, where dehydration is the most pressing problem, frogs conserve body fluid by
reabsorbing water across the epithelium of the urinary bladder.
Marine bony fishes, being hypoosmotic to their surroundings, have the opposite problem
of their freshwater relatives.
In many species, nephrons have small glomeruli or lack glomeruli altogether.
Concentrated urine is produced by secreting ions into excretory tubules.
The kidneys of marine fishes excrete very little urine and function mainly to get rid of
divalent ions such as Ca2+, Mg2+, and SO42−, which the fish takes in by its incessant
drinking of seawater.
Its gills excrete mainly monovalent ions +
such as Na + and Cl− and the bulk of its
nitrogenous wastes in the form of NH4 .
An animal hormone is a chemical signal that is secreted into the circulatory system that
communicates regulatory messages within the body.
A hormone may reach all parts of the body, but only specific target cells respond to
specific hormones.
A given hormone traveling in the bloodstream elicits specific responses from its target
cells, while other cell types ignore that particular hormone.
Concept 45.1 The endocrine system and the nervous system act individually and
together in regulating an animal’s physiology
Animals have two systems of internal communication and regulation, the nervous system
and the endocrine system.
Collectively, all of an animal’s hormone-secreting cells constitute its endocrine system.
Hormones coordinate slow but long-acting responses to stimuli such as stress,
dehydration, and low blood glucose levels.
Hormones also regulate long-term developmental processes such as growth and
development of primary and secondary sexual characteristics.
Hormone-secreting organs called endocrine glands secrete hormones directly into the
extracellular fluid, where they diffuse into the blood.
The nervous and endocrine systems overlap to some extent.
Certain specialized nerve cells known as neurosecretory cells release hormones into
the blood.
The hormones produced by these cells are sometimes called neurohormones.
Chemicals such as epinephrine serve as both hormones of the endocrine system and
neurotransmitters in the nervous system.
The nervous system plays a role in certain sustained responses—controlling day/night
cycles and reproductive cycles in many animals, for example—often by increasing or
decreasing secretions from endocrine glands.
The fundamental concepts of biological control systems are important in regulation by
hormones.
A receptor, or sensor, detects a stimulus and sends information to a control center.
After comparing the incoming information to a set point, the control center sends out a
signal that directs an effector to respond.
In endocrine and neuroendocrine pathways, this outgoing signal, called an efferent
system, is a hormone or neurohormone, which acts on particular effector tissues and
elicits specific physiological or developmental changes.
The three types of simple hormonal pathways (simple endocrine pathway, simple
neurohormone pathway, and simple neuroendocrine pathway) include these basic
functional components.
Concept 45.2 Hormones and other chemical signals bind to target cell receptors,
initiating pathways that culminate in specific cell responses
Hormones convey information via the bloodstream to target cells throughout the body.
Other chemical signals—local regulators—transmit information to target cells near the
secreting cells.
Pheromones carry messages to different individuals of a species.
Three major classes of molecules function as hormones in vertebrates: proteins and
peptides, amines, and steroids.
Most protein/peptides and amine hormones are water-soluble, unlike steroid hormones.
Signaling by all hormones involves three key events: reception, signal transduction, and
response.
Reception of the signal occurs when the signal molecule binds to a specific receptor
protein in or on the target cell.
Binding of a signal molecule to a receptor protein triggers signal transduction within
the target cell that results in a response, a change in the cell’s behavior.
Cells that lack receptors for a particular chemical signal are unresponsive to that
signal.
Water-soluble hormones have cell-surface receptors.
The receptors for water-soluble hormones are embedded in the plasma membrane.
Binding of a hormone to its receptor initiates a signal transduction pathway, a series of
changes in cellular proteins that converts an extracellular chemical signal to a specific
intracellular response.
The response may be the activation of an enzyme, a change in uptake or secretion of
specific molecules, or rearrangement of the cytoskeleton.
Signal transduction from some cell-surface receptors activates proteins in the
cytoplasm that move into the nucleus and directly or indirectly regulate gene
transcription.
Concept 45.5 Invertebrate regulatory systems also involve endocrine and nervous
system interactions
Invertebrates produce a variety of hormones in endocrine and neurosecretory cells.
Some invertebrate hormones have homeostatic functions, such as regulation of water
balance.
Others function in reproduction and development.
Biology Chapter Notes
In hydras, one hormone functions in growth and budding (asexual reproduction) but
prevents sexual reproduction.
In the mollusc Aplysia, specialized nerve cells secrete a neurohormone that stimulates
the laying of thousands of eggs and inhibits feeding and locomotion, activities that
interfere with reproduction.
All groups of arthropods have extensive endocrine systems.
Crustaceans have hormones for growth and reproduction, water balance, movement of
pigments in the integument and eyes, and regulation of metabolism.
Crustaceans and insects grow in spurts, shedding the old exoskeleton and secreting a new
one with each molt.
Insects acquire their adult characteristics in a single terminal molt.
In all arthropods with exoskeletons, molting is triggered by a hormone.
The hormonal control of insect development is well understood.
Brain hormone, produced by neurosecretory cells in the brain, stimulates the release of
ecdysone from the prothoracic glands, a pair of endocrine glands behind the head.
Ecdysone promotes molting and the development of adult characteristics.
Brain hormone and ecdysone are balanced by juvenile hormone, secreted by the corpora
allata, a pair of small endocrine glands that are somewhat analogous to the anterior
pituitary of vertebrates.
As the name suggests, juvenile hormone promotes the retention of larval (juvenile)
characteristics.
In the presence of a high concentration of juvenile hormone, ecdysone still stimulates
molting, but the product is simply a larger larva.
Only when the level of juvenile hormone declines can ecdysone-induced molting produce
a pupa.
Within the pupa, metamorphosis produces the adult form.
Synthetic juvenile hormone is used as insecticide to prevent insects from maturing to
reproductive adults.
Concept 46.1 Both asexual and sexual reproduction occur in the animal kingdom
Asexual reproduction involves the formation of individuals whose genes come from a
single parent.
There is no fusion of sperm and egg.
Sexual reproduction is the formation of offspring by the fusion of haploid gametes to
form a diploid zygote.
The female gamete, the unfertilized egg, or ovum, is usually large and nonmotile.
The male gamete is the sperm, which is usually small and motile.
Sexual reproduction increases genetic variation among offspring by generating unique
combinations of genes inherited from two parents.
Diverse mechanisms of asexual reproduction enable animals to produce identical
offspring rapidly.
Many invertebrates can reproduce asexually by fission, in which a parent separates into
two or more approximately equal-sized individuals.
Budding is also common among invertebrates. This is a form of asexual reproduction
in which new individuals split off from existing ones.
In fragmentation, the body breaks into several pieces, some or all of which develop
into complete adults.
Reproducing in this way requires regeneration of lost body parts.
Many animals can also replace new appendages by regeneration.
Asexual reproduction has a number of advantages.
It allows isolated animals to reproduce without needing to find a mate.
It can create numerous offspring in a short period of time.
In stable environments, it allows for the perpetuation of successful genotypes.
Concept 46.2 Fertilization depends on mechanisms that help sperm meet eggs of the
same species
The mechanisms of fertilization, the union of sperm and egg, play an important part in
sexual reproduction.
In external fertilization, eggs are released by the female into a wet environment,
where they are fertilized by the male.
In species with internal fertilization, sperm are deposited in or near the female
reproductive tract, and fertilization occurs within the tract.
A moist habitat is almost always required for external fertilization, both to prevent
gametes from drying out and to allow the sperm to swim to the eggs.
In species with external fertilization, timing is crucial to ensure that mature sperm
encounter ripe eggs.
Environmental cues such as temperature or day length may cause gamete release by
the whole population.
Individuals may engage in courtship behavior that leads to fertilization of the eggs of
one female by one male.
Internal fertilization is an adaptation to terrestrial life that enables sperm to reach an egg
in a dry environment.
Internal fertilization requires sophisticated reproductive systems, including copulatory
organs that deliver sperm and receptacles for their storage and transport to ripe eggs.
The sexual response can be divided into four phases: excitement, plateau, orgasm, and
resolution.
Excitement prepares the vagina and penis for coitus.
Vasocongestion is evident in the erection of the penis and clitoris; the enlargement of
the testes, labia, and breasts; and vaginal lubrication.
Myotonia may result in nipple erection or tension in the arms and legs.
Biology Chapter Notes
In the plateau phase, these responses continue.
Stimulation by the autonomic nervous system increases breathing and heart rate.
In females, plateau includes vasocongestion of the outer third of the vagina, expansion
of the inner two-thirds of the vagina, and elevation of the uterus to form a depression
that receives sperm at the back of the vagina.
Orgasm is the shortest phase of the sexual response cycle.
It is characterized by rhythmic, involuntary contractions of the reproductive structures
in both sexes.
In male orgasm, emission is the contraction of the glands and ducts of the reproductive
tract, which forces semen into the urethra.
Ejaculation occurs with the contraction of the urethra and expulsion of semen.
In female orgasm, the uterus and outer vagina contract.
Resolution completes the cycle and reverses the responses of earlier stages.
Vasocongested organs return to their normal sizes and colors; muscles relax.
Concept 46.5 In humans and other placental mammals, an embryo grows into a
newborn in the mother’s uterus
In placental mammals, pregnancy or gestation is the condition of carrying one or more
embryos.
A human pregnancy averages 266 days.
Many rodents have gestation periods of 21 days. Cows have a gestation of 27 days, and
elephant gestation lasts 600 days.
Fertilization or conception occurs in the oviduct.
Twenty-four hours later, cleavage begins.
Three to four days after fertilization, the embryo reaches the uterus as a ball of cells.
By one week past fertilization, the blastocyst forms as a sphere of cells containing a
cavity.
After a few more days, the blastocyst implants in the endometrium.
The embryo secretes hormones to signal its presence and control the mother’s
reproductive system.
Human chorionic gonadotropin (HCG) acts like pituitary LH to maintain secretion
of progesterone and estrogens by the corpus luteum for the first few weeks of
pregnancy.
Some HCG is excreted in the urine, where it is detected by pregnancy tests.
Human gestation is divided into three trimesters of three months each.
For the first 2–4 weeks of development, the embryo obtains nutrients from the
endometrium.
The outer layer of the blastocyst, called the trophoblast invades the endometrium,
eventually helping to form the placenta.
Biology Chapter Notes
The placenta allows diffusion of material between maternal and embryonic
circulations, providing nutrients, exchanging respiratory gases, and disposing of
metabolic wastes for the embryo.
Blood from the embryo travels to the placenta and returns via the umbilical vein.
Organogenesis occurs during the first trimester.
By the end of week four, the heart is beating.
By the end of week eight, all the major structures of the adult are present in
rudimentary form.
The rapidity of development makes this a time when the embryo is especially
sensitive to environmental insults such as radiation or drugs.
High levels of progesterone initiate changes in the maternal reproductive system.
These include increased mucus in the cervix to form a protective plug, growth of
the maternal part of the placenta, enlargement of the uterus, and cessation of
ovarian and menstrual cycling.
The breasts enlarge rapidly and may be very tender.
During the second trimester, the fetus grows rapidly to 30 cm and is very active.
The mother may feel movements during the early part of the second trimester.
Hormonal levels stabilize as HCG declines, the corpus luteum deteriorates, and the
placenta takes over the secretion of progesterone, which maintains the pregnancy.
During the third trimester, the fetus grows rapidly to about 3–4 kg in weight and 50 cm in
length.
Fetal activity may decrease as the fetus fills the space available to it.
Maternal abdominal organs become compressed and displaced, leading to frequent
urination, digestive blockages, and back strain.
A complex interplay of local regulators (prostaglandins) and hormones (estrogen and
oxytocin) induces and regulates labor.
The mechanism that triggers labor is not fully understood.
In one possible model, high levels of estrogen induce the formation of oxytocin
receptors on the uterus.
Oxytocin, produced by the fetus and the mother’s posterior pituitary, stimulates
powerful contractions by the smooth muscles of the uterus.
Oxytocin also stimulates the placenta to secrete prostaglandins, which enhance the
contractions.
The physical and emotional stress associated with the contractions stimulate the release
of more oxytocin and prostaglandins, a positive feedback system that underlies the
process of labor.
Birth, or parturition, is brought about by strong, rhythmic uterine contractions.
The process of labor has three stages.
The first stage is the opening up and thinning of the cervix, ending in complete
dilation.
The second stage is the expulsion of the baby as a result of strong uterine
contractions.
The third stage is the expulsion of the placenta.
Lactation is unique to mammals.
After birth, decreasing levels of progesterone free the anterior pituitary from negative
feedback and allow prolactin secretion.
Biology Chapter Notes
Prolactin stimulates milk production 2–3 days after birth.
The release of milk from the mammary glands is controlled by oxytocin.
Reproductive immunologists are working to understand why mammalian mothers do not
reject the embryo as a foreign body, despite its paternal antigens.
The trophoblast may inhibit a maternal immune response against the embryo by
releasing signal molecules with immunosuppressive effects.
These include HCG, a variety of protein “factors,” a prostaglandin, several
interleukins, and an interferon.
Some combination of these substances may interfere with immune rejection by
acting on the mother’s T lymphocytes.
A different hypothesis suggests that the trophoblast and later the placenta secrete an
enzyme that rapidly breaks down local supplies of tryptophan, an amino acid necessary
for T cell survival and function.
This enzyme is essential for maintaining pregnancy in mice.
Another possibility is the absence of certain histocompatibility antigens on placenta
cells and the secretion of a hormone that induces synthesis of a “death activator”
protein (FasL) on placental cells.
Activated T cells have a complementary “death receptor” (Fas), and the binding of
FasL to Fas causes the T cells to self-destruct by apoptosis.
Contraception can be achieved in several ways.
Some methods prevent the release of mature secondary oocytes and sperm from
gonads, others prevent fertilization by keeping sperm and egg apart, and still others
prevent implantation of an embryo.
Fertilization can be prevented by abstinence from sexual intercourse or by any of
several barriers that keep sperm and egg apart.
Temporary abstinence is called the rhythm method of birth control.
This means of natural family planning depends on refraining from intercourse
when conception is most likely.
Ovulation can be detected by noting changes in cervical mucus and body
temperature during the menstrual cycle.
Natural family planning brings a pregnancy rate of 10–20%.
As a method of preventing fertilization, coitus interruptus, or withdrawal (removal of
the penis from the vagina before ejaculation), is unreliable.
Sperm may be present in secretions that precede ejaculation.
The several barrier methods of contraception that block sperm from meeting the egg
have pregnancy rates of less than 10%.
The condom used by the male is a thin latex or natural membrane sheath that fits over
the penis to collect the semen.
The diaphragm is a dome-shaped rubber cap that fits into the upper portion of the
vagina before intercourse.
Both methods are more effective when used in conjunction with a spermicide.
Birth control pills are chemical contraceptives with a pregnancy rate of less than 1%.
The most commonly used birth control pills are a combination of a synthetic estrogen
and progestin (progesterone-like hormone).
This combination acts by negative feedback to stop the release of GnRH by the
hypothalamus and, thus, of FSH and LH by the pituitary.
The prevention of LH release prevents ovulation.
Biology Chapter Notes
As a backup mechanism, the inhibition of FSH secretion by the low dose of
estrogen in the pills prevents follicles from developing.
A second type of birth control pill, the minipill, contains only progestin.
It does not effectively block ovulation, and it is not as effective a contraceptive as the
combination pill.
The minipill prevents fertilization mainly by causing thickening of a woman’s
cervical mucus so it blocks sperm from entering the uterus.
It also causes changes in the endometrium that interfere with implantation.
Combination pills carry a slightly elevated risk of abnormal blood clotting, high blood
pressure, heart attack, and stroke.
However, they decrease the risk of ovarian and endometrial cancers.
Sterilization is the permanent prevention of gamete release.
Tubal ligation in women involves cauterization or ligation of a section of the
oviducts to prevent the eggs from traveling into the uterus.
Vasectomy in men is the cutting of each vas deferens to prevent sperm from
entering the urethra.
Abortion is the termination of a pregnancy.
Spontaneous abortion or miscarriage occurs in as many of one-third of all
pregnancies.
In addition, 1.5 million American women choose abortions performed by
physicians each year.
A drug called mifepristone, or RU-486, enables a woman to terminate pregnancy
nonsurgically within the first seven weeks.
An analogue of progesterone, RU-486 blocks progesterone reception in the
uterus, preventing progesterone from maintaining pregnancy.
It is taken with a small amount of prostaglandin to induce uterine contractions.
The hypoblast is required for normal development and seems to help direct the formation
of the primitive streak.
Some hypoblast cells later form portions of the yolk sac.
In organogenesis, the organs of the animal body form from the three embryonic germ
layers.
Various regions of the three embryonic germ layers develop into the rudiments of organs
during the process of organogenesis.
While gastrulation involves mass cell movements, organogenesis involves more localized
morphogenetic changes in tissue and cell shape.
The first organs to form in the frog are the neural tube and notochord.
Biology Chapter Notes
The notochord is formed from dorsal mesoderm that condenses above the archenteron.
Signals sent from the notochord to the overlying ectoderm cause that region of notochord
to become neural plate.
This process is often seen in organogenesis: one germ layer signaling another to
determine the fate of the second layer.
The neural plate curves inward, rolling itself into a neural tube that runs along the
anterior-posterior axis of the embryo.
The neural tube becomes the brain and spinal cord.
Unique to vertebrate embryos is a band of cells called the neural crest, which develops
along the border where the neural tube pinches off from the ectoderm.
Neural crest cells migrate throughout the embryo, forming many cell types.
Some have proposed calling neural crest cells the “fourth germ layer.”
Somites form in strips of mesoderm lateral to the notochord.
The somites are arranged serially on both sides along the length of the notochord.
Mesenchyme cells migrate from the somites to new locations.
The notochord is the core around which the vertebrae form.
Parts of the notochord persist into adulthood as the inner portions of vertebral disks.
Somite cells also form the muscles associated with the axial skeleton.
Lateral to the somites, the mesoderm splits into two layers that form the lining of the
coelom.
As organogenesis progresses, morphogenesis and cell differentiation refine the organs that
form from the three germ layers.
Embryonic development leads to an aquatic, herbivorous tadpole larva, which later
metamorphoses into a terrestrial, carnivorous adult frog.
The derivatives of the ectoderm germ layer include epidermis of skin and its derivatives,
epithelial lining of the mouth and rectum, cornea and lens of the eyes, the nervous system,
adrenal medulla, tooth enamel, and the epithelium of the pineal and pituitary glands.
The endoderm germ layer contributes to the epithelial linings of the digestive tract (except
the mouth and rectum), respiratory system, pancreas, thyroid, parathyroids, thymus,
urethra, urinary bladder, and reproductive system.
Derivatives of the mesoderm germ layer are the notochord, the skeletal and muscular
systems, the circulatory and lymphatic systems, the excretory system, the reproductive
system (except germ cells), the dermis of skin, the lining of the body cavity, and the
adrenal cortex.
Amniote embryos develop in a fluid-filled sac within a shell or uterus.
The amniote embryo is the solution to reproduction in a dry environment.
The shelled eggs of birds and other reptiles, as well as monotreme mammals, and the
uterus of placental mammals provide an aqueous environment for development.
Within the shell or uterus, the embryos of these animals are surrounded by fluid within
a sac formed by a membrane called the amnion.
Reptiles (including birds) and mammals are thus amniotes.
Amniote development includes the formation of four extraembryonic membranes: yolk
sac, amnion, chorion, and allantois.
The cells of the yolk sac digest yolk, providing nutrients to the embryo.
Biology Chapter Notes
The amnion encloses the embryo in a fluid-filled amniotic sac that protects the embryo
from drying out.
The chorion cushions the embryo against mechanical shocks and works with the
allantois to exchange gases between the embryo and the surrounding air.
The allantois functions as a disposal sac for uric acid and functions with the chorion as
a respiratory organ.
Mammalian development has some unique features.
The eggs of most mammals are very small, storing little food.
Early cleavage is relatively slow in mammals.
In humans, the first division is complete after 36 hours, the second division after 60
hours, and the third division after 72 hours.
Relatively slow cleavage produces equal-sized blastomeres.
At the eight-cell stage, the blastomeres become tightly adhered to one another, causing
the outer surface to appear smooth.
At completion of cleavage, the embryo has more than 100 cells arranged around a
central cavity.
The blastocyst travels down the oviduct to reach the uterus.
Clustered at one end of the blastocyst is a group of cells called the inner cell mass that
develops into the embryo and contributes to all the extraembryonic membranes.
The trophoblast, the outer epithelium of the blastocyst, secretes enzymes that break down
the endometrium to facilitate implantation of the blastocyst.
The trophoblast thickens, projecting fingerlike projections into the surrounding
maternal tissue, which is rich in vascular tissue.
Invasion by the trophoblast leads to erosion of the capillaries in the surrounding
endometrium, causing the blood to spill out and bathe trophoblast tissue.
At the time of implantation, the inner cell mass forms a flat disk with an upper layer of
cells, the epiblast, and a lower layer, the hypoblast.
As in birds, the human embryo develops almost entirely from the epiblast.
As implantation is completed, gastrulation begins.
Cells move inward from the epiblast through the primitive streak to form mesoderm
and endoderm.
At the same time, extraembryonic membranes develop.
The trophoblast continues to expand into the endometrium.
The invading trophoblast, mesodermal cells derived from the epiblast, and adjacent
endometrial tissue all contribute to the formation of the placenta.
The embryonic membranes of mammals are homologous with those of birds and other
mammals.
The chorion, which completely surrounds the embryo and other embryonic
membranes, functions in gas exchange.
The amnion encloses the embryo in a fluid-filled amniotic cavity.
The yolk sac encloses another fluid-filled cavity, which contains no yolk.
The yolk sac membrane of mammals is the site of early formation of blood cells,
which later migrate to the embryo.
The fourth extraembryonic membrane, the allantois, is incorporated into the umbilical
cord, where it forms blood vessels that transport oxygen and nutrients from the
placenta to the embryo and rid the embryo of carbon dioxide and nitrogenous wastes.
Biology Chapter Notes
The extraembryonic membranes of reptiles, where embryos are nourished with yolk, were
conserved as mammals diverged in the course of evolution but with modifications adapted
to development within the reproductive tract of the mother.
The completion of gastrulation is followed by the first events of organogenesis: the
formation of the neural tube, notochord, and somites.
Concept 47.2 Morphogenesis in animals involves specific changes in cell shape, position,
and adhesion
Morphogenesis is a major aspect of development in plants and animals, but only in
animals does it involve cell movement.
Movement of parts of a cell can bring about changes in cell shape.
It can also enable a cell to migrate from one place to another within the embryo.
Changes in cell shape and cell position are involved in cleavage, gastrulation, and
organogenesis.
Changes in the shape of a cell usually involve the reorganization of the cytoskeleton.
Consider how the cells of the neural plate form the neural tube.
First, the microtubules oriented parallel to the dorsal-ventral axis of the embryo help to
lengthen the cells in that direction.
At the dorsal end of each cell is a parallel array of actin filaments oriented crosswise.
These contract, giving the cells a wedge shape that bends the ectoderm inward.
Similar changes in cell shape occur during other invaginations and evaginations of
tissue layers throughout development.
The cytoskeleton is also drives cell migration.
Cells “crawl” within the embryo by extending cytoplasmic fibers to form cellular
protrusions, in a manner akin to amoeboid movement.
The cellular protrusions of migrating embryonic cells are usually flat sheets
(lamellipodia) or spikes (filopodia).
During gastrulation, invagination is initiated by the wedging of cells on the surface of
the blastula, but the movement of cells deeper into the embryo involves the extension
of filopodia by cells at the leading edge of the migrating tissue.
The cells that first move through the blastopore and along the inside of the
blastocoel drag others along behind them as a sheet of cells.
This involuted sheet of cells forms the endoderm and mesoderm of the embryo.
Cell crawling is also involved in convergent extension, a type of morphogenetic
movement in which the cells of a tissue layer rearrange themselves so the sheet
converges and extends, becoming narrower but longer.
Convergent extension allows the archenteron to elongate in the sea urchin and frog
and is responsible for the change in shape of a frog embryo from spherical to
submarine shaped.
The movements of convergent extension probably involve the extracellular matrix
(ECM), a mixture of secreted glycoproteins lying outside the plasma membrane.
ECM fibers may direct cell movement by functioning as tracks, directing migrating
cells along particular routes.
Some ECM substances, such as fibronectins, help cells migrate by providing
anchorage for crawling.
Other ECM substances may inhibit migration in certain directions.
Biology Chapter Notes
In frog gastrulation, fibronectin fibers line the roof of the blastocoel.
As the future mesoderm moves into the interior of the embryo, cells at the free edge of
the mesodermal sheet migrate along these fibers.
Researchers can prevent the attachment of cells to fibronectin (and prevent inward
movement of the mesoderm) by injecting embryos with antifibronectin antibodies.
As migrating cells move along specific paths through the embryo, receptor proteins on
their surfaces pick up directional cues from the immediate environment.
Such signals from the ECM can direct the orientation of cytoskeletal elements to
propel the cell in the proper direction.
Cell adhesion molecules (CAMs), located on cell surfaces, bind to CAMs on other cells.
CAMs vary in amount and chemical identity with cell type.
These differences help to regulate morphogenetic movement and tissue binding.
Cadherins are also involved in cell-to-cell adhesion.
Cadherins require the presence of calcium for proper function.
There are many cadherins, and the gene for each cadherin is expressed in specific
locations at specific times during embryonic development.
Concept 47.3 The developmental fate of cells depends on their history and on inductive
signals
Development requires the timely differentiation of cells in specific locations.
Two general principles integrate the current understanding of the genetic and cellular
mechanisms that underlie differentiation during embryonic development.
First, during early cleavage divisions, embryonic cells must somehow become different
from one another.
In many animal species, initial differences result from uneven distribution of
cytoplasmic determinants (mRNAs, proteins, and other molecules) in the unfertilized
egg.
The resulting differences in the cytoplasmic composition of cells help specify body
axes and influence the expression of genes that affect the developmental fate of cells.
For example, the cells of the inner cell mass are located internally in the early
human embryo, while trophoblast cells are located on the outer surface of the
blastocyst.
The difference in cell environment determines the fate of these cells.
Second, once initial cell asymmetries are set up, subsequent interactions among the
embryonic cells influence their fate, usually by causing changes in gene expression.
This mechanism is termed induction.
Induction, which brings about the differentiation of many specialized cell types, is
mediated by diffusible chemical signals or by cell-surface interactions.
Fate mapping can reveal cell genealogies in chordate embryos.
Fate maps illustrate the developmental history of cells.
In classic experiments in the 1920s, German embryologist Vogt charted fate maps for
different regions of early amphibian embryos.
His work provided evidence that the lineage of cells making up the three germ layers
created by gastrulation is traceable to cells in the blastula, before gastrulation begins.
Biology Chapter Notes
Developmental biologists have combined fate-mapping studies with experimental
manipulation of parts of embryos.
Two important conclusions have emerged.
“Founder cells” give rise to specific tissues in older embryos.
As development proceeds, a cell’s developmental potential (the range of structures it
can form) becomes restricted.
The eggs of most vertebrates have cytoplasmic determinants that help establish the body
axes.
A bilaterally symmetrical animal has an anterior-posterior axis, a dorsal-ventral axis, and
left and right sides.
Establishing this basic body plan is a first step in morphogenesis and a prerequisite for
the development of tissues and organs.
In frogs, locations of melanin and yolk define the animal and vegetal hemispheres
respectively.
The animal-vegetal axis indirectly determines the anterior-posterior body axis.
Fertilization in frogs triggers cortical rotation, which establishes the dorsal-ventral axis
and leads to the appearance of the gray crescent, whose position marks the dorsal side.
Once any two axes are established, the third (right-left) is specified by default.
Molecular mechanisms then carry out the program associated with that axis.
In amniotes, body axes are not fully established until later.
In chicks, gravity is involved in establishing the anterior-posterior axis as the egg
travels down the oviduct before being laid.
Later, pH differences between the two sides of the blastoderm establish the dorsal-
ventral axis.
In mammals, no polarity is obvious until after cleavage, although recent research suggests
that the orientation of the egg and sperm nuclei before fusion may play a role in
determining the axes.
In many species with cytoplasmic determinants, only the zygote is totipotent, capable of
developing into all cell types found in the adult.
The fate of embryonic cells is affected by both the distribution of cytoplasmic
determinants and cleavage pattern.
In frogs, the first cleavage occurs along an axis that produces two identical blastomeres
with identical developmental potential.
The cells of the mammalian embryo remain totipotent until the 16-cell stage, when they
become arranged into the precursors of the trophoblast and inner cell mass of the
blastocyst.
At that time, location determines cell fate.
At the 8-cell stage, each of the blastomeres of the mammalian embryo can form a
complete embryo if isolated.
The progressive restriction of potency is a general feature of development in animals.
In some species, the cells of the early gastrula retain the capacity to give rise to more
than one kind of cell, although they are no longer totipotent.
In general, the tissue-specific fates of cells in the late gastrula are fixed.
Even if manipulated experimentally, they will give rise to the same type of cells as
in a normal embryo.
Biology Chapter Notes
Inductive signals play an important role in cell fate determination and pattern formation.
Once embryonic cell division creates cells that are different from one another, the cells
begin to influence each other’s fates by induction.
At the molecular level, the effect of induction is usually the switching on of a set of
genes that make the receiving cells differentiate into a specific tissue.
In the 1920s, Hans Spemann and Hilde Mangold carried out a set of transplantation
experiments.
These experiments showed that the dorsal lip of the blastopore in an early gastrula
serves as an organizer of the embryo by initiating a chain of inductions that results in
the formation of the notochord, neural tube, and other organs.
Developmental biologists are working to identify the molecular basis of induction by
Spemann’s organizer (also called the gastrula organizer or simply the organizer).
A growth factor called bone morphogenetic protein 4 (BMP-4) is active exclusively in
cells on the ventral side of the amphibian gastrula.
BMP-4 induces those cells to form ventral structures.
Organizer cells inactivate BMP-4 on the dorsal side of the embryo by producing
proteins that bind to BMP-4, rendering it unable to signal.
This allows formation of dorsal structures such as the notochord and neural tube.
Proteins related to BMP-4 and its inhibitors are also found in other animals, suggesting
that they evolved long ago and may participate in development in many different
organisms.
Many inductions involve a sequence of inductive steps that progressively determine the
fate of cells.
In late gastrula of the frog, ectoderm cells destined to form the lenses of the eyes
receive inductive signals from the ectodermal cells that will form the neural plate.
Later, inductive signals from the optic cup, an outgrowth of the developing brain,
complete the determination of lens-forming cells.
Inductive signals play a major role in pattern formation, the development of an animal’s
spatial information.
Positional information, supplied by molecular cues, tells a cell where it is relative to
the animal’s body axes.
Limb development in chicks serves as a model of pattern formation.
Wings and legs of chicks begin as bumps of tissue called limb buds.
Each component of a chick limb develops with a precise location and orientation
relative to three axes, the proximal-distal axis (shoulder-to-fingertip), the anterior-
posterior axis (thumb-to-little-finger), and the dorsal-ventral axis (knuckle-to-palm).
A limb bud consists of a core of mesodermal tissue covered by a layer of ectoderm.
Two critical organizer regions are present in all vertebrate limb buds.
The cells of these regions secrete proteins that provide key positional information to
the other cells of the bud.
One limb-bud organizer region is the apical ectodermal ridge (AER), a thickened area of
ectoderm at the tip of the bud.
The AER is required for the outgrowth of the limb along the proximal-distal axis and
for patterning along this axis.
The cells of the AER produce several secreted protein signals, belonging to the
fibroblast growth factor (FGF) family.
Biology Chapter Notes
These signals promote limb-bud outgrowth.
If the AER is surgically removed and beads soaked in FGF are put in its place, a nearly
normal limb will develop.
The AER (and other limb-bud ectoderms) also appears to guide pattern formation along
the limb’s dorsal-ventral axis.
If the ectoderm of the limb bud, including the AER, is detached from the mesoderm
and rotated 180° back-to-front, the limb elements that form have reversed dorsal-
ventral orientation.
The second major limb-bud organizer region is the zone of polarizing activity (ZPA), a
block of mesodermal tissue located underneath the ectoderm where the posterior side of
the bud is attached to the body.
The ZPA is necessary for proper pattern formation along the anterior-posterior axis of
the limb.
Cells nearest the ZPA give rise to posterior structures (such as our little finger); cells
farthest from the ZPA form anterior structures (such as our thumb).
Tissue transplantation experiments support the hypothesis that the ZPA produces an
inductive signal that conveys positional information indicating “posterior.”
The cells of the ZPA secrete a protein growth factor called Sonic hedgehog.
If cells genetically engineered to produce large amounts of Sonic hedgehog are
implanted in the anterior region of a normal limb bud, a mirror-image limb bud
results.
Extra toes and fingers in mice (and maybe humans) result from the production
of Sonic hedgehog in the wrong part of the limb bud.
We can conclude from these experiments that pattern formation requires cells to receive
and interpret environmental cues that vary with location.
These cues tell cells where they are in the 3-D realm of a developing organ.
Organizers such as the AER and the ZPA function as signaling centers.
The AER and ZPA also interact with each other via signaling molecules and signaling
pathways, to influence each other’s developmental fates.
What determines whether a limb bud develops into a forelimb or a hindlimb?
The cells receiving signals from the AER and ZPA respond according to their own
developmental histories.
Earlier developmental signals have set up patterns of gene expression that distinguish
future forelimbs from future hindlimbs.
Construction of a fully formed animal involves a sequence of events that include many
steps of signaling and differentiation.
Initial cell asymmetries allow different types of cells to influence each other to express
specific sets of genes.
The products of these genes direct cells to differentiate into specific types.
Coordinated with morphogenesis, various pathways of pattern formation occur in all
the different parts of the developing embryo.
These processes produce a complex arrangement of multiple tissues and organs, each
functioning in the appropriate location to form a coordinated organism.
Concept 48.1 Nervous systems consist of circuits of neurons and supporting cells
Nervous systems show diverse patterns of organization.
All animals except sponges have some type of nervous system.
What distinguishes nervous systems of different animal groups is how the neurons are
organized into circuits.
Cnidarians have radially symmetrical bodies organized around a gastrovascular cavity.
In hydras, neurons controlling the contraction and expansion of the gastrovascular
cavity are arranged in diffuse nerve nets.
The nervous systems of more complex animals contain nerve nets, as well as nerves,
which are bundles of fiberlike extensions of neurons.
With cephalization come more complex nervous systems.
Neurons are clustered in a brain near the anterior end in animals with elongated,
bilaterally symmetrical bodies.
In simple cephalized animals such as the planarian, a small brain and longitudinal nerve
cords form a simple central nervous system (CNS).
In more complex invertebrates, such as annelids and arthropods, behavior is regulated by
more complicated brains and ventral nerve cords containing segmentally arranged clusters
of neurons called ganglia.
Concept 48.2 Ion pumps and ion channels maintain the resting potential of a neuron
Every cell has a voltage, or membrane potential, across its plasma membrane.
All cells have an electrical potential difference (voltage) across their plasma membrane).
This voltage is called the membrane potential.
In neurons, the membrane potential is typically between −60 and −80 mV when the
cell is not transmitting signals.
The membrane potential of a neuron that is not transmitting signals is called the resting
potential.
In all neurons, the resting potential depends on the ionic gradients that exist across the
plasma membrane.
In mammals,+
the extracellular fluid has a Na+ concentration of 150 millimolar (mM)
and a K of 5 mM.
In the cytosol, Na+ concentration is 15 mM, and K+ concentration is 150 mM.
These gradients are maintained by the sodium-potassium pump.
The magnitude of the membrane voltage at equilibrium, called the equilibrium potential
(Eion), is given by a formula called the Nernst equation.
For an ion with a net charge of +1, the Nernst equation is:
Eion = 62mV (log [ion]outside/[ion]inside)
The Nernst equation applies to any membrane that is permeable to a single type of ion.
In our model, the membrane is only permeable to K+ +, and the Nernst equation can be
used to calculate EK, the equilibrium potential for K .
With this K+ concentration gradient, K+ is at equilibrium when the inside of the
membrane is 92 mV more negative than the outside.
Assume that the membrane is only permeable to Na+.
ENa, the equilibrium potential+
for Na+, is +62 mV, indicating that, with this Na+
concentration gradient, Na is at equilibrium when the inside of the membrane is 62
mV more positive than the outside.
How does a real mammalian neuron differ from these model neurons?
The plasma membrane of a real neuron at rest has many open potassium channels, but it
also has a relatively small number of open sodium channels.
Biology Chapter Notes
Consequently, the resting potential is around −60 to −80 mV, between EK and ENa.
Neither K+ nor Na + is at equilibrium, and there is a net flow of each ion (a current)
across the membrane at rest.
The resting membrane potential remains steady, which means that the K + and Na+ currents
are equal and opposite.
The reason the +resting potential is closer to E K than to ENa is that the membrane is more
permeable to K than to Na+.
If something causes the membrane’s permeability to Na + to increase, the membrane
potential will move toward ENa and away from EK.
This is the basis of nearly all electrical signals in the nervous system.
The membrane potential can change from its resting value when the membrane’s
permeability to particular ions changes.
Sodium and potassium play major roles, but there are also important roles for chloride and
calcium ions.
The resting potential results from the diffusion of K + and Na+ through ion channels that
are always open.
These channels are ungated.
Neurons also have gated ion channels, which open or close in response to one of three
types of stimuli.
Stretch-gated ion channels are found in cells that sense stretch, and open when the
membrane is mechanically deformed.
Ligand-gated ion channels are found at synapses and open or close when a specific
chemical, such as a neurotransmitter, binds to the channel.
Voltage-gated ion channels are found in axons (and in the dendrites and cell bodies of
some neurons, as well as in some other types of cells) and open or close in response to
a change in membrane potential.
Concept 48.6 The cerebral cortex controls voluntary movement and cognitive functions
The cerebrum is divided into frontal, temporal, occipital, and parietal lobes.
Researchers have identified a number of functional areas within each lobe.
These areas include primary sensory areas, each of which receives and processes a
specific type of sensory information, and association areas, which integrate the
information from various parts of the brain.
The major increase in the size of the neocortex that occurred during mammalian evolution
was mostly an expansion of the association areas that integrate higher cognitive functions
and make more complex behavior and learning possible.
Most sensory information coming into the cortex is directed via the thalamus to primary
sensory areas within the lobes: visual information to the occipital lobe; auditory input to
the temporal lobe; and somatosensory information about touch, pain, pressure,
temperature, and position of limbs and muscles to the parietal lobe.
In mammals, olfactory information is first sent to regions in the cortex that are similar
in mammals and reptiles, and then via the thalamus to an interior part of the frontal
lobe.
Based on the integrated sensory information, the cerebral cortex can generate motor
commands that cause specific behaviors.
These commands consist of action potentials produced by neurons in the primary
motor cortex, which lies at the rear of the frontal lobe.
The action potentials travel along axons to the brainstem and spinal cord, where they
excite motor neurons, which in turn excite skeletal muscle cells.
In both the somatosensory cortex and the motor cortex, neurons are distributed in an
orderly fashion according to the part of the body that generates the sensory input or
receives the motor command.
The cortical surface area devoted to each body part is not related to the size of the part.
Concept 48.7 CNS injuries and diseases are the focus of much research
Unlike the PNS, the mammalian CNS does not have the ability to repair itself when
damaged or injured by disease.
Surviving neurons in the brain can make new connections and sometimes compensate for
damage.
Generally speaking, brain and spinal cord injuries, strokes, and diseases that destroy
CNS neurons have devastating effects.
Research on nerve cell development and neural stem cells may be the future of treatment
for damage to the CNS.
Researchers are investigating how neurons “find their way” during CNS development.
To reach their target cells, axons must elongate from a few micrometers to a meter or
more.
Molecular signposts along the way direct and redirect the growing axon in a series of
mid-course connections that result in a meandering, but not random, elongation.
The responsive region at the leading edge of the neuron is called the growth cone.
Signal molecules released by cells along the growth route bind to receptors on the
plasma membrane of the growth cone, triggering a signal transduction pathway.
The axon may respond by growing toward the source of the signal molecules
(attraction) or away from it (repulsion).
Cell adhesion molecules on the axon’s growth cone also play a role by attaching to
complementary molecules on surrounding cells that provide tracks for the growing
axon to follow.
Nerve growth factor released by astrocytes and growth-promoting proteins produced
by the neurons themselves contribute to the process by simulating axonal elongation.
The growing axon expresses different genes as it develops, and it is influenced by
surrounding cells that it moves away from.
This complex process has been conserved during millions of years of evolution, for
the genes, gene products, and mechanisms of axon guidance are remarkably similar
in humans, nematode worms, and insects.
In 1998, it was discovered that a adult human brain does produce new neurons.
New neurons have been found in the hippocampus.
The origins of sensing date back to the appearance in prokaryotes of cellular structures
that sense pressure and chemicals in the environment and direct movement in an
appropriate direction.
These structures have been transformed during the course of evolution into diverse
mechanisms that sense various types of energy and generate many different levels of
physical movement in response.
The detection and processing of sensory information and the generation of motor output
provide the physiological basis for all animal behavior.
Concept 49.1 Sensory receptors transduce stimulus energy and transmit signals to the
central nervous system
o The brain’s processing of sensory input and motor output is cyclical rather than
linear.
Sensing, brain analysis, and action are ongoing and overlapping processes.
Information is transmitted through the nervous system in the form of all-or-nothing action
potentials.
What matters is where action potentials go.
Sensations begin as different forms of energy detected by sensory receptors.
This energy is converted to action potentials that travel to appropriate regions of the
brain.
Once the brain is aware of sensations, it interprets them, giving the perception of
stimuli.
Perceptions such as colors, smells, sounds, and tastes are constructions formed in the
brain and do not exist outside of it.
o Sensory receptors transduce stimulus energy and transmit signals to the nervous
system.
Sensory reception begins with the detection of stimulus energy by sensory receptors.
Most sensory receptors are specialized neurons or epithelial cells that exist singly or in
groups with other cell types in sensory organs, such as eyes or ears.
Exteroreceptors detect stimuli originating outside the body, such as heat, light,
pressure, and chemicals.
Interoreceptors detect stimuli originating inside the body, such as blood pressure and
body position.
Sensory receptors convey the energy of stimuli into membrane potentials and transmit
signals to the nervous system.
Sensory receptors perform four functions in this process: sensory transduction,
amplification, transmission, and integration.
Biology Chapter Notes
The conversion of stimulus energy into a change in membrane potential of a sensory
receptor is sensory transduction.
The change in membrane potential itself is receptor potential.
Receptor potentials are graded potentials; their magnitude varies with the strength
of the stimulus.
All receptor potentials result from the opening or closing of ion channels in the
sensory receptor’s plasma membrane.
Many sensory receptors are extremely sensitive.
Most light receptors can detect a single photon of light.
Hair cells of the inner ear can detect motion of only a fraction of a nanometer.
Chemical receptors can detect a single molecule.
The strengthening of stimulus energy by cells in sensory pathways is called
amplification.
An action potential conducted from the eye to the brain has about 100,000 times as
much energy as the few photons of light that triggered it.
Some amplification occurs in the sensory receptors, and signal transduction
pathways involving second messengers often contribute to it.
Amplification can also take place in the accessory structures of a complex sense
organ.
The conduction of sensory impulses to the CNS is transmission.
Some sensory receptors transmit chemical signals to sensory neurons.
The strength of the stimulus and receptor potential affects the amount of excitatory
neurotransmitter released by the sensory receptor.
Some sensory receptors are sensory neurons.
The intensity of the receptor potential affects the frequency of action potentials.
Many sensory neurons spontaneously generate action potentials at a low rate.
Therefore, a stimulus does not switch the production of action potentials on or off
in these neurons.
Rather, it modulates action potential frequency.
The processing, or integration, of sensory information begins as soon as the information
is received.
Receptor potentials produced by stimuli delivered to different parts of a sensory
receptor are integrated through summation, as are postsynaptic potentials in sensory
neurons that synapse with multiple receptors.
Another type of integration by receptors is sensory adaptation, a decrease in
responsiveness to continued stimulation.
The integration of sensory information occurs at all levels within the nervous system.
Complex sensory structures such as eyes have higher levels of integration, and the
CNS further processes all incoming signals.
o Sensory receptors are categorized by the type of energy they transduce.
Mechanoreceptors respond to mechanical energy such as pressure, touch, stretch,
motion, and sound.
Bending or stretching of a mechanoreceptor’s plasma membrane increases the
membrane’s permeability to sodium and potassium ions.
The crayfish stretch receptor, the vertebrate hair cell, and the vertebrate stretch receptor
are examples of mechanoreceptors.
Biology Chapter Notes
Muscle spindles respond to the stretching of skeletal muscle, depolarizing sensory
neurons and triggering action potentials that are transmitted to the spinal cord.
Muscle spindles and the sensory neurons that innervate them are part of the nerve
circuits that underlie reflexes.
The mammalian sense of touch also relies on mechanoreceptors that are the dendrites
of sensory neurons, embedded in layers of connective tissue.
Receptors that detect light touch are close to the surface of the skin, while receptors
responding to strong pressure and vibrations are in deep skin layers.
Chemoreceptors respond to chemical stimuli.
General chemoreceptors transmit information about total solute concentration of a
solution, while specific chemoreceptors respond to specific types of molecules.
Osmoreceptors in the mammalian brain are general receptors that detect changes in
solute concentration of the blood and stimulate thirst when osmolarity increases.
Internal chemoreceptors respond to glucose, O2, CO2, and amino acids.
Two of the most sensitive and specific chemoreceptors known are present in the
antennae of the male silkworm moth.
They detect the two chemical components of the female moth sex pheromone.
In each example, the stimulus molecule binds to a specific site on the membrane of the
receptor cell and initiates changes in membrane permeability.
Electromagnetic receptors respond to various forms of electromagnetic energy such as
visible light, electricity, and magnetism.
Photoreceptors respond to the radiation we know as visible light and are often
organized into eyes.
Some snakes have infrared detectors that detect body heat of prey.
Some fishes generate electric currents and use electroreceptors to locate prey that
disrupt those currents.
Many animals use Earth’s magnetic field lines to orient themselves as they migrate.
The iron-containing mineral magnetite is found in the skulls of many vertebrates, in
the abdomen of bees, in the teeth of some molluscs, and in certain protists and
prokaryotes that orient to Earth’s magnetic field.
Thermoreceptors respond to heat or cold and help regulate body temperature by
signaling surface and body core temperature.
Thermoreceptors in the skin and in the anterior hypothalamus send information to the
body’s thermostat, located in the posterior hypothalamus.
Pain receptors, or nociceptors, are a class of naked dendrites in the epidermis.
Most animals experience pain, although we cannot say what perceptions other animals
associate with stimulation of pain receptors.
Pain is an important sensation, because the stimulus leads to a defensive reaction.
Different types of pain receptors respond to different types of pain, such as excess
heat, pressure, or chemicals released from damaged or inflamed tissues.
Prostaglandins increase pain by sensitizing receptors, lowering their threshold.
Aspirin and ibuprofen reduce pain by inhibiting prostaglandin synthesis.
Concept 49.4 Similar mechanisms underlie vision throughout the animal kingdom
Many types of light detectors have evolved in the animal kingdom, from simple clusters
of cells that detect only direction and intensity of light to complex image-forming eyes.
Biology Chapter Notes
All photoreceptors contain similar pigment molecules that absorb light.
Most, if not all, animal photoreceptors may be homologous.
All animals with vision share genes associated with the embryonic development of
photoreceptors.
The genetic underpinnings of all photoreceptors may have evolved in the earliest
bilateral animals.
The specific types of eyes that form in an animal depend on developmental patterns
regulated by genetic mechanisms that evolved later, superimposed on the common
ancestral mechanism.
o A diversity of photoreceptors has evolved among invertebrates.
The eye cups of planarians are among the simplest photoreceptors.
These structures detect light intensity and direction, but do not provide image
formation.
The movement of a planarian is integrated with photoreception.
Two major types of image-forming eyes have evolved in invertebrates.
One type of eye is the compound eye of insects and crustaceans.
Each eye consists of ommatidia, each with its own light-focusing lens.
Each ommatidium detects light from a tiny portion of the visual field.
This type of eye is very good at detecting movement.
Insects have excellent color vision, and some can see ultraviolet light.
Single-lens eyes are found in some invertebrates such as jellies, polychaetes, spiders, and
molluscs.
The eye of an octopus works much like a camera and is similar to the vertebrate eye.
Light enters through the pupil, with the iris changing the diameter.
Behind the pupil, a single lens focuses light on a layer of photoreceptors.
Muscles in an invertebrate’s single-lens eye move the lens to focus at different
distances.
o Vertebrates have single-lens eyes.
Vertebrate eyes are structurally analogous to the invertebrate single-lens eye.
The globe of the vertebrate eye (the eyeball) consists of a tough, white outer layer of
connective tissue called the sclera and a thin, pigmented inner layer called the choroid.
A delicate layer of epithelial cells forms a mucus membrane, the conjunctiva, which
covers the external cover of the sclera and keeps the eye moist.
At the front of the eye is a transparent cornea, which lets light into the eye and acts as
a fixed lens.
The conjunctiva does not cover the cornea.
The anterior choroid prevents light rays from scattering and distorting the image.
Anteriorly, it forms the iris, which gives the eye its color.
The iris regulates the size of the pupil.
Inside the choroid, the retina lines the interior surface of the choroid.
The retina contains photoreceptors, except at the optic disk (where the optic nerve
attaches).
The lens (a transparent protein disk) and ciliary body divide the eye into two cavities.
Biology Chapter Notes
The anterior cavity is filled with aqueous humor produced by the ciliary body.
Glaucoma results when the ducts that drain aqueous humor are blocked.
The posterior cavity is filled with vitreous humor.
The lens, the aqueous humor, and the vitreous humor all play a role in focusing light onto
the retina.
In squids, octopuses, and many fishes, this is accomplished by moving the lens
forward and backward.
In mammals, focus is accomplished by changing the shape of the lens.
The lens is flattened for focusing on distant objects.
The lens is rounded for focusing on near objects.
When focusing on a close object, the lens becomes almost spherical, a change called
accommodation.
The retina contains about 125 million rod cells, which are light sensitive but do not
distinguish colors, and about 6 million cone cells, which are not as light sensitive as rods
but provide color vision.
These cells account for 70% of the sensory receptors in the body.
Rods are most highly concentrated at the peripheral regions of the retina and are
completely absent from the fovea, the center of the visual field.
Cones are most dense at the fovea, which has 150,000 cones per square millimeter.
o The light-absorbing pigment rhodopsin triggers a signal-transduction pathway.
Each rod or cone in the vertebrate retina contains visual pigments consisting of light-
absorbing molecules called retinal bonded to membrane proteins called opsin.
Rhodopsin (retinal + opsin) is the visual pigment of rods.
The absorption of light by rhodopsin initiates a signal-transduction pathway.
Color reception is more complex than the rhodopsin mechanism.
There are three subclasses of cone cells, each with its own type of photopsin.
Color perception is based on the brain’s analysis of the relative responses of each
type of cone.
In humans, colorblindness is due to a deficiency, or absence, of one or more
photopsins.
It is inherited as an X-linked trait and is more common in males than females.
o The retina assists the cerebral cortex in processing visual information.
Visual processing begins with rods and cones synapsing with neurons called bipolar cells.
In the dark, rods and cones are depolarized, and they continually release the
neurotransmitter glutamate at these synapses.
This steady glutamate release depolarizes some bipolar cells and hyperpolarizes others.
In the light, rods and cones hyperpolarize, shutting off the release of glutamate.
In response, the bipolar cells that are depolarized by glutamate hyperpolarize, and
those that are hyperpolarized by glutamate depolarize.
Three other types of neurons contribute to information processing in the retina: ganglion
cells, horizontal cells, and amacrine cells.
Bipolar cells synapse with ganglion cells and transmit action potentials to the brain via
axons in the optic nerve.
Ecology is the scientific study of the interactions between organisms and their
environment.
Concept 50.1 Ecology is the study of interactions between organisms and the
environment
Ecologists ask questions about factors affecting the distribution and abundance of
organisms.
Ecologists might study how interactions between organisms and the environment affect
the number of species living in an area, the cycling of nutrients, or the growth of
populations.
Ecology and evolutionary biology are closely related sciences.
Ecology has a long history as a descriptive science.
Modern ecology is also a rigorous experimental science.
Ecology and evolutionary biology are closely related sciences.
Events that occur over ecological time (minutes to years) translate into effects over
evolutionary time (decades to millennia).
For example, hawks feeding on field mice kill certain individuals (over ecological
time), reducing population size (an ecological effect), altering the gene pool (an
evolutionary effect), and selecting for mice with fur color that camouflages them in
their environment (over evolutionary time).
Ecological research ranges from the adaptations of individual organisms to the dynamics
of the biosphere.
The environment of any organism includes the following components:
Abiotic components: nonliving chemical and physical factors such as temperature,
light, water, and nutrients.
Biotic components: all living organisms in the individual’s environment.
Ecology can be divided into a number of areas of study.
Organismal ecology is concerned with the behavioral, physiological, and morphological
ways individuals interact with the environment.
A population is a group of individuals of the same species living in a particular
geographic area. Population ecology examines factors that affect population size and
composition.
A community consists of all the organisms of all the species that inhabit a particular area.
Community ecology examines the interactions between species and considers how
Biology Chapter Notes
factors such as predation, competition, disease, and disturbance affect community
structure and organization.
An ecosystem consists of all the abiotic factors in addition to the entire community of
species that exist in a certain area. Ecosystem ecology studies energy flow and cycling of
chemicals among the various abiotic and biotic components.
A landscape or seascape consists of several different ecosystems linked by exchanges of
energy, materials, and organisms. Landscape ecology deals with arrays of ecosystems
and their arrangement in a geographic region.
Each landscape or seascape consists of a mosaic of different types of patches, an
environmental characteristic ecologists refer to as patchiness. Landscape ecological
research focuses on the factors controlling exchanges of energy, materials, and
organisms among ecosystem patches.
The biosphere is the global ecosystem, the sum of all of the planet’s ecosystems. The
biosphere includes the entire portion of Earth inhabited by life, ranging from the
atmosphere to a height of several kilometers to the oceans and water bearing rocks to a
depth of several kilometers.
Ecology provides a scientific context for evaluating environmental issues.
It is important to clarify the difference between ecology, the scientific study of the
distribution and abundance of organisms, and environmentalism, advocacy for the
protection or preservation of the natural environment.
To address environmental problems, we need to understand the interactions of organisms
and the environment.
The science of ecology provides that understanding.
In 1962, Rachel Carson’s book Silent Spring warned that the use of pesticides such as
DDT was causing population declines in many nontarget organisms.
Today, acid precipitation, land misuse, toxic wastes, habitat destruction, and the growing
list of endangered or extinct species are just a few of the problems that threaten the Earth.
Many influential ecologists feel a responsibility to educate legislators and the general
public about decisions that affect the environment.
It is important to communicate the scientific complexity of environmental issues.
Our ecological information is always incomplete. The precautionary principle
(essentially “an ounce of prevention is worth a pound of cure”) can guide decision making
on environmental issues.
Concept 50.2 Interactions between organisms and the environment limit the
distribution of species
Ecologists have long recognized distinct global and regional patterns in the distribution of
organisms.
Biogeography is the study of past and present distributions of individual species in the
context of evolutionary theory.
Ecologists ask a series of questions to determine what limits the geographical distribution
of any species.
Species dispersal contributes to the distribution of organisms.
Concept 50.3 Abiotic and biotic factors influence the structure and dynamics of aquatic
biomes
Varying combinations of biotic and abiotic factors determine the nature of the Earth’s
biomes, major types of ecological associations that occupy broad geographic regions of
land or water.
Aquatic biomes occupy the largest part of the biosphere.
Ecologists distinguish between freshwater and marine biomes on the basis of physical and
chemical differences.
Marine biomes generally have salt concentrations that average 3%, while freshwater
biomes have salt concentrations of less than 1%.
Marine biomes cover approximately 75% of the earth’s surface and have an enormous
effect on the biosphere.
The evaporation of water from the oceans provides most of the planet’s rainfall.
Ocean temperatures have a major effect on world climate and wind patterns.
Photosynthesis by marine algae and photosynthetic bacteria produce a substantial
proportion of the world’s oxygen. Respiration by these organisms consumes huge
amounts of atmospheric carbon dioxide.
Freshwater biomes are closely linked to the soils and biotic components of the terrestrial
biomes through which they pass.
The pattern and speed of water flow and the surrounding climate are also important.
Most aquatic biomes are physically and chemically stratified.
Light is absorbed by the water and by photosynthetic organisms, so light intensity
decreases rapidly with depth.
There is sufficient light for photosynthesis in the upper photic zone.
Very little light penetrates to the lower aphotic zone.
The substrate at the bottom of an aquatic biome is the benthic zone.
This zone is made up of sand and sediments and is occupied by communities of
organisms called benthos.
A major food source for benthos is dead organic material or detritus, which rains
down from the productive surface waters of the photic zone.
Sunlight warms surface waters, while deeper waters remain cold.
Biology Chapter Notes
As a result, water temperature in lakes is stratified, especially in summer and winter.
In the ocean and most lakes, a narrow stratum of rapid temperature change called a
thermocline separates the more uniformly warm upper layer from more uniformly
cold deeper waters.
In aquatic biomes, community distribution is determined by depth of the water, distance
from shore, and open water versus bottom.
In marine communities, phytoplankton, zooplankton, and many fish species live in the
relatively shallow photic zone.
The aphotic zone contains little life, except for microorganisms and relatively sparse
populations of luminescent fishes and invertebrates.
The major aquatic biomes include lakes, wetlands, streams, rivers, estuaries, intertidal
biomes, oceanic pelagic biomes, coral reefs, and marine benthic biomes.
Freshwater lakes vary greatly in oxygen and nutrient content.
Oligotrophic lakes are deep, nutrient poor, oxygen rich, and contain little life.
Eutrophic lakes are shallow, nutrient rich, and oxygen poor.
In lakes, the littoral zone is the shallow, well-lit water close to shore.
The limnetic zone is the open surface water.
Wetlands are areas covered with sufficient water to support aquatic plants.
They can be saturated or periodically flooded.
Wetlands include marshes, bogs, and swamps.
They are among the most productive biomes on Earth and are home to a diverse
community of invertebrates and birds.
Because of the high organic production and decomposition in wetlands, their water and
soil are low in dissolved oxygen.
Wetlands have a high capacity to filter dissolved nutrients and chemical pollutants.
Humans have destroyed many wetlands, but some are now protected.
Streams and rivers are bodies of water moving continuously in one direction.
Headwaters are cold, clear, turbulent, and swift.
They carry little sediment and relatively few mineral nutrients.
As water travels downstream, it picks up O2 and nutrients on the way.
Nutrient content is largely determined by the terrain and vegetation of the area.
Many streams and rivers have been polluted by humans, degrading water quality
and killing aquatic organisms.
Damming and flood control impairs the natural functioning of streams and rivers
and threatens migratory species such as salmon.
Estuaries are areas of transition between river and sea.
The salinity of these areas can vary greatly.
Estuaries have complex flow patterns, with networks of tidal channels, islands, levees,
and mudflats.
They support an abundance of fish and invertebrate species and are crucial feeding
areas for many species of waterfowl.
An intertidal zone is a marine biome that is periodically submerged and exposed by the
tides.
The upper intertidal zone experiences longer exposure to air and greater variation in
salinity and temperature than do the lower intertidal areas.
Biology Chapter Notes
Many organisms live only at a particular stratum in the intertidal.
The oceanic pelagic biome is the open blue water, mixed by wind-driven oceanic
currents.
The surface waters of temperate oceans turn over during fall through spring.
The open ocean has high oxygen levels and low nutrient levels.
This biome covers 70% of the Earth’s surface and has an average depth of 4,000
meters.
Coral reefs are limited to the photic zone of stable tropic marine environments with high
water clarity. They are found at temperatures between 18°C and 30°C.
They are formed by the calcium carbonate skeletons of coral animals.
Mutualistic dinoflagellate algae live within the tissues of the corals.
Coral reefs are home to a very diverse assortment of vertebrates and invertebrates.
Collecting of coral skeletons and overfishing for food and the aquarium trade have
reduced populations of corals and reef fishes.
Global warming and pollution contribute to large-scale coral mortality.
The marine benthic zone consists of the seafloor below the surface waters of the coastal
or neritic zone and the offshore pelagic zone.
Most of the ocean’s benthic zone receives no sunlight.
Organisms in the very deep abyssal zone are adapted to continuous cold (about 3°C)
and extremely high pressure.
Unique assemblages of organisms are associated with deep-sea hydrothermal vents
of volcanic origin on mid-ocean ridges.
The primary producers in these communities are chemoautotrophic prokaryotes that
obtain energy by oxidizing2−H2S formed by a reaction of volcanically heated water
with dissolved sulfate (SO4 ).
Concept 50.4 Climate largely determines the distribution and structure of terrestrial
biomes
Because there are latitudinal patterns of climate over the Earth’s surface, there are also
latitudinal patterns of biome distribution.
A climograph denotes the annual mean temperature and precipitation of a region.
Temperature and rainfall are well correlated with different terrestrial biomes, and each
biome has a characteristic climograph.
Most terrestrial biomes are named for major physical or climatic features or for their
predominant vegetation.
Vertical stratification is an important feature of terrestrial biomes.
The canopy of the tropical rain forest is the top layer, covering the low-tree stratum,
shrub understory, ground layer, litter layer, and root layer.
Grasslands have a canopy formed by grass, a litter layer, and a root layer.
Stratification of vegetation provides many different habitats for animals.
Terrestrial biomes usually grade into each other without sharp boundaries. The area of
intergradation, called the ecotone, may be narrow or wide.
The species composition of any biome differs from location to location.
Biomes are dynamic, and natural disturbance rather than stability tends to be the rule.
Biology Chapter Notes
Hurricanes create openings for new species in tropical and temperate forests.
In northern coniferous forests, snowfall may break branches and small trees, producing
gaps that allow deciduous species to grow.
As a result, biomes exhibit patchiness, with several different communities represented
in any particular area.
In many biomes, the dominant plants depend on periodic disturbance.
For example, natural wildfires are an integral component of grasslands, savannas,
chaparral, and many coniferous forests.
Human activity has radically altered the natural patterns of periodic physical disturbance.
Fires are now controlled for the sake of agricultural land use.
Humans have altered much of the Earth’s surface, replacing original biomes with urban or
agricultural ones.
The major terrestrial biomes include tropical forest, desert, savanna, chaparral, temperate
grassland, coniferous forest, temperate broadleaf forest, and tundra.
Tropical forests are found close to the equator.
Tropical rain forests receive constant high amounts of rainfall (200 to 400 cm
annually).
In tropical dry forests, precipitation is highly seasonal.
In both, air temperatures range between 25°C and 29°C year round.
Tropical forests are stratified, and competition for light is intense.
Animal diversity is higher in tropical forests than in any other terrestrial biome.
Deserts occur in a band near 30° north and south latitudes and in the interior of
continents.
Deserts have low and highly variable rainfall, generally less than 30 cm per year.
Temperature varies greatly seasonally and daily.
Desert vegetation is usually sparse and includes succulents such as cacti and deeply
rooted shrubs.
Many desert animals are nocturnal, so they can avoid the heat.
Desert organisms display adaptations to allow them to resist or survive desiccation.
Savanna is found in equatorial and subequatorial regions.
Rainfall is seasonal, averaging 30–50 cm per year.
The savanna is warm year-round, averaging 24–29°C with some seasonal variation.
Savanna vegetation is grassland with scattered trees.
Large herbivorous mammals are common inhabitants.
The dominant herbivores are insects, especially termites.
Fire is important in maintaining savanna biomes.
Chaparrals have highly seasonal precipitation with mild, wet winters and dry, hot
summers.
Annual precipitation ranges from 30 to 50 cm.
Chaparral is dominated by shrubs and small trees, with a high diversity of grasses and
herbs.
Plant and animal diversity is high.
Adaptations to fire and drought are common.
Humans have studied animal behavior for as long as we have lived on Earth.
As hunter and hunted, knowledge of animal behavior was essential to human behavior.
The modern scientific discipline of behavioral ecology studies how behavior develops,
evolves, and contributes to survival and reproductive success.
Concept 51.1 Behavioral ecologists distinguish between proximate and ultimate causes
of behavior
Scientific questions that can be posed about any behavior can be divided into two classes:
those that focus on the immediate stimulus and mechanism for the behavior and those that
explore how the behavior contributes to survival and reproduction.
What is behavior?
Behavioral traits are an important part of an animal’s phenotype.
Many behaviors result from an animal’s muscular activity, such as a predator chasing a
prey.
In some behaviors, muscular activity is less obvious, as in bird song.
Some nonmuscular activities are also behaviors, as when an animal secretes a
pheromone to attract a member of the opposite sex.
Learning is also a behavioral process.
Put simply, behavior is everything an animal does and how it does it.
Proximate questions are mechanistic, concerned with the environmental stimuli that
trigger a behavior, as well as the genetic, physiological, and anatomical mechanisms
underlying a behavioral act.
Proximate questions are referred to as “how?” questions.
Ultimate questions address the evolutionary significance of a behavior and why natural
selection favors this behavior.
Ultimate questions are referred to as “why?” questions.
Red-crowned cranes breed in spring and early summer.
A proximate question about the timing of breeding by this species might ask, “How
does day length influence breeding by red-crowned cranes?”
A reasonable hypothesis for the proximate cause of this behavior is that breeding is
triggered by the effect of increased day length on the crane’s production of and
response to particular hormones.
An ultimate hypothesis might be that red-crowned cranes reproduce in spring and early
summer because that is when breeding is most productive.
At that time of year, parent birds can find an ample supply of food for rapidly
growing offspring, providing an advantage in reproductive success compared to
birds that breed in other seasons.
These two levels of causation are related.
Biology Chapter Notes
Proximate mechanisms produce behaviors that evolved because they increase fitness in
some way.
For example, increased day length has little adaptive significance for red-crowned
cranes, but because it corresponds to seasonal conditions that increase reproductive
success, such as the availability of food for feeding young birds, breeding when days
are long is a proximate mechanism that has evolved in cranes.
Classical ethology presaged an evolutionary approach to behavioral biology.
In the mid-20th century, a number of pioneering behavioral biologists developed the
discipline of ethology, the scientific study of animal behavior.
In 1963, Niko Tinbergen suggested four questions that must be answered to fully
understand any behavior.
o What is the mechanistic basis of the behavior, including chemical, anatomical, and
physiological mechanisms?
o How does development of the animal, from zygote to mature individual, influence the
behavior?
o What is the evolutionary history of the behavior?
o How does the behavior contribute to survival and reproduction (fitness)?
Tinbergen’s list includes both proximate and ultimate questions.
The first two, which concern mechanism and development, are proximate questions,
while the second two are ultimate, or evolutionary, questions.
A fixed action pattern (FAP) is a sequence of unlearned behavioral acts that is
essentially unchangeable and, once initiated, is usually carried to completion.
A FAP is triggered by an external sensory stimulus called a sign stimulus.
In the red-spined stickleback, the male attacks other males that invade his nesting
territory.
The stimulus for the attack is the red underside of the intruder.
A male stickleback will attack any model that has some red visible on it.
A proximate explanation for this aggressive behavior is that the red belly of the intruding
male acts as a sign stimulus that releases aggression in a male stickleback.
An ultimate explanation is that by chasing away other male sticklebacks, a male decreases
the chance that eggs laid in his nesting territory will be fertilized by another male.
Imprinting is a type of behavior that includes learning and innate components and is
generally irreversible.
Imprinting has a sensitive period, a limited phase in an animal’s behavior that is the
only time that certain behaviors can be learned.
An example of imprinting is young geese following their mother.
In species that provide parental care, parent-offspring bonding is a critical time in the
life cycle.
During the period of bonding, the young imprint on their parent and learn the basic
behavior of the species, while the parent learns to recognize its offspring.
Among gulls, the sensitive period for parental bonding on young lasts one or two days.
If bonding does not occur, the parent will not initiate care of the infant, leading to
certain death of the offspring and decreasing the parent’s reproductive success.
How do young gulls know on whom—or what—to imprint?
The tendency to respond is innate in birds.
Biology Chapter Notes
The world provides the imprinting stimulus, and young gulls respond to and identify
with the first object they encounter that has certain key characteristics.
In greylag geese, the key stimulus is movement of the object away from the young.
A proximate explanation for young geese following and imprinting on their mother is that
during an early, critical developmental stage, the young geese observe their mother
moving away from them and calling.
An ultimate explanation is that, on average, geese that follow and imprint on their mother
receive more care and learn necessary skills, and thus have a greater chance of surviving,
than those that do not follow.
Early study of imprinting and fixed action patterns helped make the distinction between
proximate and ultimate causes of behavior.
They also helped to establish a strong tradition of experimental approaches in
behavioral ecology.
Concept 51.3 Environment, interacting with an animal’s genetic makeup, influences the
development of behaviors
Concept 51.5 Natural selection favors behaviors that increase survival and reproductive
success
The genetic components of behavior evolve through natural selection favoring traits that
enhance survival and reproductive success in a population.
Two of the most direct ways that behavior can affect fitness are through influences on
foraging and mate choice.
Foraging includes not only eating, but also any mechanisms that an animal uses to
recognize, search for, and capture food items.
Optimal foraging theory views foraging behavior as a compromise between the benefits
of nutrition and the costs of obtaining food, such as the energy expenditure and risk of
predation while foraging.
Biology Chapter Notes
Natural selection should favor foraging behavior that minimizes the costs of foraging
and maximizes the benefits.
Behavioral ecologists apply cost-benefit analysis to study the proximate and ultimate
causes of diverse foraging strategies.
Reta Zach of the University of British Columbia carried out a cost-benefit analysis of
feeding behavior in crows.
Crows search the tide pools of Mandarte Island, B.C., for snails called whelks.
A crow flies up and drops the whelk onto the rocks to break its shell.
If the drop is successful, the crow eats the snail’s soft body.
If it is not successful, the crow flies higher and tries again.
Zach predicted—and found—that crows would, on average, fly to a height that would
provide the most food relative to the total amount of energy required to break the
whelk shells.
Bluegill sunfish feed on small crustaceans called Daphnia, selecting larger individuals
that supply the most energy per unit time.
Smaller individuals will be selected if larger prey are too far away.
Optimal foraging theory predicts that the proportion of small to large prey captured will
vary with prey density.
At high densities, it is efficient for bluegill sunfish to feed only on large crustaceans.
At low densities, bluegill sunfish should exhibit little size selectivity because all prey
are needed to meet energy requirements.
In experiments, young bluegill sunfish forage efficiently but not as close to optimum as
older individuals.
Perhaps younger fish do not judge size and distance as accurately because their vision
is not yet completely developed.
Learning may also improve the foraging efficiency of bluegill sunfish as they age.
Risk of predation is one of the most significant potential costs to a forager.
Mule deer are preyed on by mountain lions throughout their range.
Researchers studied mule deer populations in Idaho to determine if they forage in a
way that reduces their risk of falling prey to mountain lions.
The researchers found that food available to mule deer was fairly uniform across the
potential foraging area.
Risk of predation varied greatly, however.
Mountain lions killed most mule deer at forest edges.
Few were killed in open areas and forest interiors.
How does mule deer feeding behavior respond to the differences in feeding risk?
Mule deer feed predominantly in open areas, avoiding forest edges and forest interiors.
When deer are at the forest edge, they spend significantly more time scanning their
surroundings than when they are in other areas.
Mating behavior, which includes seeking and attracting mates, choosing among potential
mates, and competing for mates, is the product of a form of natural selection called sexual
selection.
The mating relationship between males and females varies a great deal from species to
species.
Concept 51.6 The concept of inclusive fitness can account for most altruistic social
behavior
Most social behaviors are selfish, meaning that they benefit the individual at the expense
of others, especially competitors.
Behavior that maximizes an individual’s survival and reproductive success is favored by
selection, regardless of its effect on other individuals.
How do we account for behaviors that help others?
Altruism is defined as behavior that appears to decrease individual fitness but
increases the fitness of others.
Belding’s ground squirrel lives in some mountainous regions of the western United States.
The squirrel is vulnerable to predators such as coyotes and hawks.
Biology Chapter Notes
If a squirrel sees a predator approach, it often gives a high-pitched alarm call, which
alerts unaware individuals.
The alerted squirrels then retreat to their burrows.
This conspicuous alarm behavior calls attention to the caller, who has a greater risk of
being killed.
In honeybees, workers are sterile but labor on behalf of a single fertile queen.
Workers will sacrifice themselves to sting intruders in defense of the hive.
Naked mole rats are highly social rodents that live in underground chambers and tunnels
in Africa.
These rodents are hairless and nearly blind and live in colonies of 75–250 individuals.
Each colony has only one reproducing female, the queen, who mates with one to three
males, called kings.
The rest of the colony consists of nonreproductive females and males who forage for
underground roots and tubers and care for the kings, queen, and young rats.
How can a naked mole rat (or a honeybee or a ground squirrel) enhance its fitness by
helping other members of the population?
How is altruistic behavior maintained by evolution?
If related individuals help each other, they are, in effect, helping keep their own genes
in the population.
Inclusive fitness is defined as the effect an individual has on proliferating its own genes
by reproducing and by helping relatives raise offspring.
William Hamilton proposed a quantitative measure for predicting when natural selection
should favor altruistic acts.
Hamilton’s rule states the conditions under which altruistic acts will be favored by
natural selection.
The three key variables are as follows:
o The benefit to the recipient is B.
o The cost to the altruist is C.
o The coefficient of relatedness is r, which equals the probability that a particular gene
present in one individual will also be inherited from a common parent or ancestor in a
second individual.
The rule is as follows:
rB > C
The more closely related two individuals are, the greater the value of altruism.
Kin selection is the mechanism of inclusive fitness, where individuals help relatives raise
young.
Some animals behave altruistically toward others who are not close relatives.
Such behavior can be adaptive if the aided individual can be counted on to return the
favor in the future.
This exchange of aid is called reciprocal altruism and is commonly used to explain
altruism between unrelated humans.
Reciprocal altruism is limited to species with stable social groups in which individuals
have many opportunities to exchange aid and where there would be negative social
consequences for those who “cheat” and refuse to return favors to those who have helped
them in the past.
Biology Chapter Notes
However, because cheating may provide a large benefit to cheaters, behavioral ecologists
have questioned how reciprocal altruism could arise.
To answer this question, behavioral ecologists have turned to game theory.
Axelrod and Hamilton found that reciprocal altruism can evolve and persist in a
population where individuals adopt a behavioral strategy called tit for tat.
In this strategy, an individual treats another individual the same way it was treated the
last time they met.
Individuals are always altruistic, or cooperative, on the first encounter, and will remain
so as long as their altruism is reciprocated.
When it is not, they will retaliate immediately but will return to cooperative
behavior as soon as the other individual becomes cooperative.
Animals learn by observing others.
Social learning is learning through observing others.
Social learning forms the roots of culture, which can be defined as a system of
information transfer through social learning or teaching.
Cultural transfer of information has the potential to alter behavioral phenotypes and
influence the fitness of individuals.
Social learning is not restricted to humans.
In many species, mate choice is strongly influenced by social learning.
Mate choice copying, a behavior in which individuals in a population copy the mate
choices of others, has been extensively studied in the guppy Poecilia reticulata.
Female guppies prefer to mate with males having a high percentage of orange coloration.
However, if a female sees another female engaging in courtship with a male with
relatively little orange, she will choose a male with little orange herself.
Below a certain threshold of difference in mate color, mate choice copying by female
guppies can mask genetically controlled female preference for orange males.
What is the advantage for females?
A female that mates with males that are attractive to other females may increase the
probability that her male offspring will also be attractive and have high reproductive
success.
In their studies of vervet monkeys in Amboseli National Park, Kenya, Dorothy Cheny and
Richard Seyfarth found that performance of a behavior can improve through learning.
Vervet monkeys (Cercopithecus aethiops) produce a complex set of alarm calls.
Distinct alarm calls warn of leopards, eagles, or snakes, all of which prey on the small
vervets.
Vervets react to each alarm differently, depending on the threat.
Infant vervets give alarm calls but in an undiscriminating way.
For example, they call “eagle” for any bird.
With age, they improve their accuracy.
Vervets learn how to give the right call by observing other members of the group
and by receiving social confirmation for accurate calls.
Sociobiology places social behavior in an evolutionary context.
Concept 52.4 The logistic growth model includes the concept of carrying capacity
Typically, resources are limited.
As population density increases, each individual has access to an increasingly smaller
share of available resources.
Ultimately, there is a limit to the number of individuals that can occupy a habitat.
Ecologists define carrying capacity (K) as the maximum stable population size that a
particular environment can support.
Carrying capacity is not fixed but varies over space and time with the abundance of
limiting resources.
Energy limitation often determines carrying capacity, although other factors, such as
shelters, refuges from predators, soil nutrients, water, and suitable nesting sites can be
limiting.
If individuals cannot obtain sufficient resources to reproduce, the per capita birth rate b
will decline.
If they cannot find and consume enough energy to maintain themselves, the per capita
death rate m may increase.
A decrease in b or an increase in m results in a lower per capita rate of increase r.
We can modify our mathematical model to incorporate changes in growth rate as the
population size nears the carrying capacity.
In the logistic population growth model, the per capita rate of increase declines as
carrying capacity is reached.
Mathematically, we start with the equation for exponential growth, adding an expression
that reduces the rate of increase as N increases.
Concept 52.5 Populations are regulated by a complex interaction of biotic and abiotic
influences
Why do all populations eventually strop growing?
What environmental factors stop a population from growing?
Why do some populations show radical fluctuations in size over time, while others remain
relatively stable?
These questions have practical applications at the core of management programs for
agricultural pests or endangered species.
The first step to answering these questions is to examine the effects of increased
population density on rates of birth, death, immigration, and emigration.
Density-dependent factors have an increased effect on a population as population density
increases.
This is a type of negative feedback.
Density-independent factors are unrelated to population density.
Biology Chapter Notes
Negative feedback prevents unlimited population growth.
A variety of factors can cause negative feedback on population growth.
Resource limitation in crowded populations can reduce population growth by reducing
reproductive output.
Intraspecific competition for food can lead to declining birth rates.
In animal populations, territoriality may limit density.
In this case, territory space becomes the resource for which individuals compete.
The presence of nonbreeding individuals in a population is an indication that
territoriality is restricting population growth.
Population density can also influence the health and thus the survival of organisms.
As crowding increases, the transmission rate of a disease may increase.
Tuberculosis, caused by bacteria that spread through the air when an infected person
coughs or sneezes, affects a higher percentage of people living in high-density cities
than in rural areas.
Predation may be an important cause of density-dependent mortality for a prey species if a
predator encounters and captures more food as the population density of the prey
increases.
As a prey population builds up, predators may feed preferentially on that species,
consuming a higher percentage of individuals.
The accumulation of toxic wastes can contribute to density-dependent regulation of
population size.
In wine, as yeast population increases, they accumulate alcohol during fermentation.
However, yeast can only withstand an alcohol percentage of approximately 13%
before they begin to die.
For some animal species, intrinsic factors appear to regulate population size.
White-footed mice individuals become more aggressive as population size increases,
even when food and shelter are provided in abundance.
Eventually the population ceases to grow.
These behavioral changes may be due to hormonal changes, which delay sexual
maturation and depress the immune system.
All populations for which we have data show some fluctuation in numbers.
The study of population dynamics focuses on the complex interactions between biotic
and abiotic factors that cause variation in population size.
Populations of large mammals, such as deer and moose, were once thought to remain
relatively stable over time.
A long-term population study of a moose population on Isle Royale has challenged that
view.
The population has had two major increases and collapses over the past 40 years.
Large mammal populations do show much more stability than other populations.
Dungeness crab populations fluctuated hugely over a 40-year period.
One key factor causing these fluctuations is cannibalism.
Large numbers of juveniles are eaten by older juveniles and older crabs.
In addition, successful settlement of crabs is dependent on water temperatures and
ocean currents.
Concept 52.6 Human population growth has slowed after centuries of exponential
increase
The concepts of population dynamics can be applied to the specific case of the human
population.
It is unlikely that any other population of large animals has ever sustained so much
population growth for so long.
One prey species may gain protection by mimicking the appearance of another prey
species.
In Batesian mimicry a harmless, palatable species mimics a harmful, unpalatable
model.
In Müllerian mimicry, two or more unpalatable species resemble each other.
Each species gains an additional advantage because predators are more likely to
encounter an unpalatable prey and learn to avoid prey with that appearance.
Predators may also use mimicry.
Some snapping turtles have tongues resembling wiggling worms to lure small fish.
Herbivory is a +/− interaction in which an herbivore eats parts of a plant or alga.
Herbivores include large mammals and small invertebrates.
Herbivores have specialized adaptations.
Many herbivorous insects have chemical sensors on their feet to recognize
appropriate food plants.
Mammalian herbivores have specialized dentition and digestive systems to process
vegetation.
Plants may produce chemical toxins, which may act in combination with spines and
thorns to prevent herbivory.
Concept 53.5 Contrasting views of community structure are the subject of continuing
debate
Biology Chapter Notes
The integrated hypothesis of community structure depicts a community as an
assemblage of closely linked species locked into association by mandatory biotic
interactions.
The community functions as an integrated unit, as a superorganism.
The individualistic hypothesis of community structure depicts a community as a chance
assemblage of species found in the same area because they happen to have similar abiotic
requirements for rainfall, temperature, or soil type.
These two very different hypotheses suggest different priorities in studying biological
communities.
The integrated hypothesis emphasizes assemblages of species as the essential units for
understanding the interactions and distributions of species.
The individualistic hypothesis emphasizes single species.
The hypotheses make contrasting predictions about how plant species should be
distributed along an environmental gradient.
The integrated hypothesis predicts that species should be clustered into discrete
communities with noticeable boundaries because the presence or absence of a
particular species is largely governed by the presence or absence of other species.
The individualistic hypothesis predicts that communities should generally lack discrete
geographic boundaries because each species has an independent, individualistic,
distribution along the environmental gradient.
In most cases where there are broad regions characterized by gradients of environmental
variation, the composition of plant communities does seem to change continuously, with
each species more or less independently distributed.
The debate continues with the rivet and redundancy models.
The individualistic hypothesis is generally accepted by plant ecologists.
Further debate arises when these ideas are applied to animals.
American ecologists Anne and Paul Ehrlich proposed the rivet model of communities.
This hypothesis is a reincarnation of the interactive model and suggests that most animal
species are associated with particular other species in the community.
Reducing or increasing the abundance of one species in a community will affect many
other species.
Australian ecologist Brian Walker’s redundancy model proposes that most animal
species in a community are not closely associated with one another.
Species operate independently, and an increase or decrease in one species in a
community has little effect on other species.
In this sense, species in a community are redundant.
If a predator disappears, another predatory species will take its place as a consumer of
specific prey.
The rivet and redundancy models represent extremes; most communities have some
features of each model.
We still do not have enough information to answer the fundamental questions raised by
these models: Are communities loose associations of species or highly integrated
units?
To fully assess these models, we need to study how species interact in communities
and how tight these interactions are.
An ecosystem consists of all the organisms living in a community as well as all the
abiotic factors with which they interact.
The dynamics of an ecosystem involve two processes that cannot be fully described by
population or community processes and phenomena: energy flow and chemical cycling.
Energy enters most ecosystems in the form of sunlight.
It is converted to chemical energy by autotrophs, passed to heterotrophs in the organic
compounds of food, and dissipated as heat.
Chemical elements are cycled among abiotic and biotic components of the ecosystem.
Energy, unlike matter, cannot be recycled.
An ecosystem must be powered by a continuous influx of energy from an external
source, usually the sun.
Energy flows through ecosystems, while matter cycles within them.
Concept 54.1 Ecosystem ecology emphasizes energy flow and chemical cycling
Ecosystem ecologists view ecosystems as transformers of energy and processors of
matter.
We can follow the transformation of energy by grouping the species in a community into
trophic levels of feeding relationships.
Ecosystems obey physical laws.
The law of conservation of energy states that energy cannot be created or destroyed but
only transformed.
Plants and other photosynthetic organisms convert solar energy to chemical energy,
but the total amount of energy does not change.
The total amount of energy stored in organic molecules plus the amounts reflected and
dissipated as heat must equal the total solar energy intercepted by the plant.
The second law of thermodynamics states that some energy is lost as heat in any
conversion process.
We can measure the efficiency of ecological energy conversions.
Chemical elements are continually recycled.
A carbon or nitrogen atom moves from one trophic level to another and eventually to
the decomposers and back again.
Trophic relationships determine the routes of energy flow and chemical cycling in
ecosystems.
Autotrophs, the primary producers of the ecosystem, ultimately support all other
organisms.
Concept 54.2 Physical and chemical factors limit primary production in ecosystems
The amount of light energy converted to chemical energy by an ecosystem’s autotrophs in
a given time period is an ecosystem’s primary production.
An ecosystem’s energy budget depends on primary production.
Most primary producers use light energy to synthesize organic molecules, which can be
broken down to produce ATP.
The amount of photosynthetic production sets the spending limit of the entire ecosystem.
A global energy budget can be analyzed.
Every day, Earth is bombarded by approximately 10 23 joules of solar radiation.
The intensity of solar energy striking Earth varies with latitude, with the tropics
receiving the greatest input.
Most of this radiation is scattered, absorbed, or reflected by the atmosphere.
Much of the solar radiation that reaches Earth’s surface lands on bare ground or
bodies of water that either absorb or reflect the energy.
Only a small fraction actually strikes algae, photosynthetic prokaryotes, or plants,
and only some of this is of wavelengths suitable for photosynthesis.
Of the visible light that reaches photosynthetic organisms, only about 1% is
converted to chemical energy.
Concept 54.3 Energy transfer between trophic levels is usually less than 20% efficient
The amount of chemical energy in consumers’ food that is converted to their own new
biomass during a given time period is called the secondary production of an ecosystem.
We can measure the efficiency of animals as energy transformers using the following
equation:
production efficiency = net secondary production / assimilation of primary production
Net secondary production is the energy stored in biomass represented by growth and
reproduction.
Assimilation consists of the total energy taken in and used for growth, reproduction, and
respiration.
Production efficiency is thus the fraction of food energy that is not used for respiration.
This differs among organisms.
Birds and mammals generally have low production efficiencies of between 1% and
3% because they use so much energy to maintain a constant body temperature.
Fishes have production efficiencies of around 10%.
Insects are even more efficient, with production efficiencies averaging 40%.
Trophic efficiency is the percentage of production transferred from one trophic level to
the next.
Trophic efficiencies must always be less than production efficiencies because they take
into account not only the energy lost through respiration and contained in feces, but
also the energy in organic material at lower trophic levels that is not consumed.
Trophic efficiencies usually range from 5% to 20%.
In other words, 80–95% of the energy available at one trophic level is not transferred
to the next.
This loss is multiplied over the length of a food chain.
If 10% of energy is transferred from primary producers to primary consumers, and
10% of that energy is transferred to secondary consumers, then only 1% of net primary
production is available to secondary consumers.
Pyramids of net production represent the multiplicative loss of energy in a food chain.
Concept 54.4 Biological and geochemical processes move nutrients between organic and
inorganic parts of the ecosystem
Chemical elements are available to ecosystems only in limited amounts.
Life on Earth depends on the recycling of essential chemical elements.
Nutrient circuits involve both biotic and abiotic components of ecosystems and are called
biogeochemical cycles.
There are two general categories of biogeochemical cycles: global and regional.
Gaseous forms of carbon, oxygen, sulfur, and nitrogen occur in the atmosphere, and
cycles of these elements are global.
Elements that are less mobile in the environment, such as phosphorus, potassium,
calcium, and trace elements generally cycle on a more localized scale in the short term.
Soil is the main abiotic reservoir for these elements.
We will consider a general model of chemical cycling that includes the main reservoirs of
elements and the processes that transfer elements between reservoirs.
Each reservoir is defined by two characteristics: whether it contains organic or
inorganic materials and whether or not the materials are directly available for use by
organisms.
Reservoir a. The nutrients in living organisms and in detritus are available to other
organisms when consumers feed and when detritivores consume nonliving organic
material.
Reservoir b. Some materials move to the fossilized organic reservoir as dead organisms
and are buried by sedimentation over millions of years. Nutrients in fossilized deposits
cannot be assimilated directly.
Reservoir c. Inorganic elements and compounds that are dissolved in water or present in
soil or air are available for use by organisms.
Reservoir d. Inorganic elements present in rocks are not directly available for use by
organisms. These nutrients may gradually become available through erosion and
weathering.
Describing biogeochemical cycles in general terms is much simpler than trying to trace
elements through these cycles.
Ecologists study chemical cycling by adding tiny amounts of radioactive isotopes to
the elements they are tracing.
There are a number of important biogeochemical cycles.
We will consider the cycling of water, carbon, nitrogen, and phosphorus.
The water cycle
Biological importance
Biology Chapter Notes
Water is essential to all organisms and its availability influences rates of ecosystem
processes.
Biologically available forms
Liquid water is the primary form in which water is used.
Reservoirs
The oceans contain 97% of the water in the biosphere.
2% is bound as ice, and 1% is in lakes, rivers, and groundwater.
A negligible amount is in the atmosphere.
Key processes
The main processes driving the water cycle are evaporation of liquid water by solar
energy, condensation of water vapor into clouds, and precipitation.
Transpiration by terrestrial plants moves significant amounts of water.
Surface and groundwater flow returns water to the oceans.
The carbon cycle
Biological importance
Organic molecules have a carbon framework.
Biologically available forms
Autotrophs convert carbon dioxide to organic molecules that are used by heterotrophs.
Reservoirs
The major reservoirs of carbon include fossil fuels, soils, aquatic sediments, the
oceans, plant and animal biomass, and the atmosphere (CO2).
Key processes
Photosynthesis by plants and phytoplankton fixes atmospheric CO 2.
CO2 is added to the atmosphere by cellular respiration of producers and consumers.
Volcanoes and the burning of fossil fuels add CO2 to the atmosphere.
The nitrogen cycle
Biological importance
Nitrogen is a component of amino acids, proteins, and nucleic acids.
It may be a limiting plant nutrient.
Biologically available forms
Plants and algae can use ammonium (NH4+) or nitrate (NO3−).
Various bacteria can also use NH4+, NO3−, or NO2.
Animals can use only organic forms of nitrogen.
Reservoirs
The major reservoir of nitrogen is the atmosphere, which is 80% nitrogen gas (N2).
Nitrogen is also bound in soils and the sediments of lakes, rivers, and oceans.
Some nitrogen is dissolved in surface water and groundwater.
Nitrogen is stored in living biomass.
Key processes
Nitrogen enters ecosystems primarily through bacterial nitrogen fixation.
Some nitrogen is fixed by lightning and industrial fertilizer production.
Biology Chapter Notes
Ammonification by bacteria decomposes organic nitrogen.
In nitrification, bacteria convert NH4+ to NO3−.
In denitrification, bacteria use NO3− for metabolism instead of O2, releasing N2.
The phosphorus cycle
Biological importance
Phosphorus is a component of nucleic acids, phospholipids, and ATP and other
energy-storing molecules.
It is a mineral constituent of bones and teeth.
Biologically available forms
The only biologically important inorganic form of phosphorus is phosphate (PO 43−),
which plants absorb and use to synthesize organic compounds.
Reservoirs
The major reservoir of phosphorus is sedimentary rocks of marine origin.
There are also large quantities of phosphorus in soils, dissolved in the oceans, and in
organisms.
Key processes
Weathering of rocks gradually adds phosphate to soil.
Some phosphate leaches into groundwater and surface water and moves to the sea.
Phosphate may be taken up by producers and incorporated into organic material.
It is returned to soil or water through decomposition of biomass or excretion by
consumers.
Decomposition rates largely determine the rates of nutrient cycling.
The rates at which nutrients cycle in different ecosystems are extremely variable as a
result of variable rates of decomposition.
Decomposition takes an average of four to six years in temperate forests, while in a
tropical rain forest, most organic material decomposes in a few months to a few years.
The difference is largely the result of warmer temperatures and more abundant
precipitation in tropical rain forests.
Like net primary production, the rate of decomposition increases with actual
evapotranspiration.
In tropical rain forests, relatively little organic material accumulates as leaf litter on the
forest floor.
75% of the nutrients in the ecosystem are present in the woody trunks of trees.
10% of the nutrients are concentrated in the soil.
In temperate forests, where decomposition is slower, the soil may contain 50% of the
organic material.
In aquatic ecosystems, decomposition in anaerobic mud of bottom sediments can take 50
years or more.
However, algae and aquatic plants usually assimilate nutrients directly from the water.
Aquatic sediments may constitute a nutrient sink.
Nutrient cycling is strongly regulated by vegetation.
Long-term ecological research (LTER) monitors the dynamics of ecosystems over long
periods of time.
Biology Chapter Notes
The Hubbard Brook Experimental Forest has been studied since 1963.
The study site is a deciduous forest with several valleys, each drained by a small creek
that is a tributary of Hubbard Brook.
Preliminary studies confirmed that internal cycling within a terrestrial ecosystem
conserves most of the mineral nutrients.
Some areas were completely logged and then sprayed with herbicides for three years to
prevent regrowth of plants.
All the original plant material was left in place to decompose.
Water runoff from the altered watershed increased by 30–40%, apparently because there
were no plants to absorb and transpire water from the soil.
The concentration of Ca2+ in the creek increased four-fold, while concentration of K+
increased by a factor of 15.
Nitrate loss was increased by a factor of 60.
This study demonstrates that the amount of nutrients leaving an intact forest ecosystem is
controlled by the plants.
Results of the Hubbard Brook studies assess natural ecosystem dynamics and provide
insight into the mechanisms by which human activities affect ecosystem processes.
Concept 54.5 The human population is disrupting chemical cycles throughout the
biosphere
Human activities and technologies have disrupted the trophic structure, energy flow, and
chemical cycling of ecosystems worldwide.
The human population moves nutrients from one part of the biosphere to another.
Human activity intrudes in nutrient cycles.
Nutrients from farm soil may run off into streams and lakes, depleting nutrients in one
area, causing excesses in another, and disrupting chemical cycles in both places.
Humans also add entirely new materials—many toxic—to ecosystems.
In agricultural ecosystems, a large amount of nutrients are removed from the area as crop
biomass.
After a while, the natural store of nutrients can become exhausted.
The soil cannot be used to grow crops without nutrient supplementation.
Nitrogen is the main nutrient lost through agriculture.
Plowing and mixing the soil increase the decomposition rate of organic matter,
releasing usable nitrogen that is then removed from the ecosystem when crops are
harvested.
Recent studies indicate that human activities have approximately doubled the worldwide
supply of fixed nitrogen, due to the use of fertilizers, cultivation of legumes, and burning.
This may increase the amount of nitrogen oxides in the atmosphere and contribute to
atmospheric warming, depletion of ozone, and possibly acid precipitation.
The key problem with excess nitrogen seems to be critical load, the amount of added
nitrogen that can be absorbed by plants without damaging the ecosystem.
Nitrogenous minerals in the soil that exceed the critical load eventually leach into
groundwater or run off into freshwater and marine ecosystems, contaminating water
supplies, choking waterways, and killing fish.
Biology Chapter Notes
Lakes are classified by nutrient availability as oligotrophic or eutrophic.
In an oligotrophic lake, primary productivity is relatively low because the mineral
nutrients required by phytoplankton are scarce.
Overall productivity is higher in eutrophic lakes.
Human intrusion has disrupted freshwater ecosystems by cultural eutrophication.
Sewage and factory wastes and runoff of animal wastes from pastures and stockyards
have overloaded many freshwater streams and lakes with nitrogen.
This results in an explosive increase in the density of photosynthetic organisms,
released from nutrient limitation.
Shallow areas become choked with weeds and algae.
As photosynthetic organisms die and organic materials accumulate at the lake bottom,
detritivores use all the available oxygen in the deeper waters.
This can eliminate fish species.
Combustion of fossil fuels is the main cause of acid precipitation.
The burning of fossil fuels releases oxides of sulfur and nitrogen that react with water in
the atmosphere to produce sulfuric and nitric acids.
These acids fall back to earth as acid precipitation—rain, snow, sleet or fog with a pH less
than 5.6.
Acid precipitation is a regional or global problem, rather than a local one.
The tall exhaust stacks built for smelters and generating plans export the problem far
downwind.
Acid precipitation lowers the pH of soil and water and affects the soil chemistry of
terrestrial ecosystems.
With decreased pH, calcium and other nutrients leach from the soil.
The resulting nutrient deficiencies affect the health of plants and limit their growth.
Freshwater ecosystems are very sensitive to acid precipitation.
Lakes underlain by granite bedrock have poor buffering capacity because of low
bicarbonate levels.
Fish populations have declined in many lakes in Norway, Sweden, and Canada as pH
levels fall.
Lake trout are keystone predators in many Canadian lakes.
When they are replaced by acid-tolerant species, the dynamics of food webs in the
lakes change dramatically.
Environmental regulations and new industrial technologies have led to reduced sulfur
dioxide emissions in many developed countries.
The water chemistry of many streams and freshwater lakes is slowly improving as a
result.
Ecologists estimate that it will take another 10 to 20 years for these ecosystems to
recover, even if emissions continue to decline.
Massive emissions of sulfur dioxide and acid precipitation continue in parts of central and
eastern Europe.
Toxins can become concentrated in successive trophic levels of food webs.
Humans introduce many toxic chemicals into ecosystems.
Overexploitation refers to the human harvesting of wild plants and animals at rates that
exceed the ability of those populations to rebound.
It is possible for overexploitation to endanger certain plant species, such as rare trees
that are harvested for their wood.
However, the term usually applies to commercially hunted or fished animal species.
Large organisms with low intrinsic reproductive rates are especially susceptible to
overexploitation.
The African elephant has been overhunted largely due to the ivory trade.
Elephant populations have declined dramatically over the past 50 years.
Despite a ban on the sale of new ivory, poaching continues in central and east
Africa.
The great auk was overhunted for its feathers, eggs, and meat.
It became extinct in the 1840s.
The bluefin tuna is another example of an overharvested species.
This big tuna brings $100 per pound in Japan, where it is used for sushi and
sashimi.
Biology Chapter Notes
With this demand, it took just ten years to reduce North American bluefin
populations to 20% of their 1980 levels.
The collapse of the northern cod fishery off Newfoundland in the 1990s shows that it is
possible to overharvest what had been a very common species.
Ecosystem dynamics depend on networks of interspecific interactions within biological
communities.
The extinction of one species can doom others, especially if the extinction involves a
keystone species, an ecosystem engineer, or a species with a highly specialized
relationship with other species.
Sea otters are a keystone species whose elimination over most of their historic range
led to major changes in the structure of shallow-water benthic communities along the
west coast of North America.
The extermination of beavers, one of the best-known ecosystem engineers, resulted in
a large reduction in wetland and pond habitats across much of North America.
Concept 55.2 Population conservation focuses on population size, genetic diversity, and
critical habitat
Biologists focusing on conservation at the population and species levels follow two main
approaches—the small-population approach and the declining-population approach.
The small-population approach studies the processes that can cause very small
populations to become extinct.
The extinction vortex is a downward spiral unique to small populations.
A small population is prone to positive-feedback loops of inbreeding and genetic drift
that draw it into a vortex toward smaller and smaller numbers until extinction is
inevitable.
The key factor driving the vortex is the loss of genetic diversity necessary to enable
evolutionary responses to environmental change, such as new strains of pathogens.
Not all populations are doomed by low genetic diversity.
Overhunting of northern elephant seals in the 1890s reduced the species to only 20
individuals—clearly a bottleneck that reduced genetic variation.
Since that time, northern elephant seal populations have rebounded to 150,000
individuals, although the genetic variation of the species remains low.
A number of plant species have inherently low genetic variation.
Species of cord grass, which thrive in salt marshes, are genetically uniform at many
loci.
Having spread by cloning, this species dominates large areas of tidal mudflats in
Europe and Asia.
How small is too small for a population? How small does a population have to be before it
starts down the extinction vortex?
The answer depends on the type of organism and its environment, and must be
determined case by case.
The greater prairie chicken (Tympanuchus cupido) was common in large areas of North
America a century ago.
Agriculture fragmented the population of the greater prairie chicken in the central and
western states and provinces.
Concept 55.3 Landscape and regional conservation aim to sustain entire biotas
On a broad scale, the principles of community, ecosystem, and landscape ecology can be
brought to bear on studies of the biodiversity of entire landscapes.
Human population dynamics and economics are also considered.
Landscape ecology is the application of ecological principles to the study of human land-
use patterns.
A landscape is a regional assemblage of interacting ecosystems.
This type of ecology is important in conservation biology because many species use
more than one type of ecosystem and many live on the borders between ecosystems.
Edges and corridors can strongly influence landscape biodiversity.
Boundaries, or edges, between ecosystems and within ecosystems are defining features of
landscapes.
An edge has its own set of physical conditions, which differ from those on either side
of it.
Edges have their own communities of organisms.
Some organisms thrive in edge communities because they have access to the resources of
both adjacent areas.
For example, the ruffled grouse (Bonasa umbellatus) requires forest habitat for nesting,
winter food, and shelter.
It also needs forest openings with dense shrubs and herbs for summer food.
The proliferation of edge species can have positive or negative effects on a community’s
biodiversity.
For example, a 1997 study in Cameroon suggested that forest edges may be important
sites of speciation.
On the other hand, communities in which edges have resulted from human alterations
often have reduced biodiversity because of domination by edge-adapted species.
Cowbirds flourish in areas where forests are heavily cut and fragmented, creating
more edge habitat and open land.
Increasing cowbird parasitism and loss of habitat are correlated with declining
populations of cowbird’s host species.
The influence of fragmentation on the structure of communities has been explored for two
decades in the long-term Biological Dynamics of Forest Fragments Project in the Amazon
River basin.
Researchers are clearly documenting the physical and biological effects of forest
fragmentation in taxa ranging from bryophytes to beetles to birds.
Concept 55.5 Sustainable development seeks to improve the human condition while
conserving biodiversity
Many have embraced the concept of sustainable development, the long-term prosperity
of human societies and the ecosystems that support them.
The Sustainable Biosphere Initiative is a research agenda endorsed by the Ecological
Society of America.
The goal is to obtain the basic ecological information necessary for responsible
development, management, and conservation of Earth’s resources.
The research agenda includes studies of global change, including interactions between
climate and ecological processes, biological diversity and its role in maintaining
ecological processes, and the ways in which the productivity of natural and artificial
ecosystems can be sustained.
This initiative requires a strong commitment of human and economic resources.
Sustainable development is not just about science.
It must include life sciences, social sciences, economics, and humanities.
Equally important, it requires a reassessment of our values.
The success of conservation in Costa Rica has involved leadership by the national
government as well as an essential partnership with nongovernmental organizations and
private citizens.
How have living conditions of Costa Ricans fared as the country pursued conservation
goals?
Infant mortality rate in Costa Rica declined sharply during the 20th century, and life
expectancy at birth increased.
The 2003 literacy rate in Costa Rica was 96%.
Such statistics show that living conditions in Costa Rica improved greatly over the period
in which the country dedicated itself to conservation and restoration.
One of the challenges the country faces is maintaining its commitment to conservation in
the face of a growing population.
Costa Rica’s population, currently 4 million, is predicted to grow to 6 million people
over the next 50 years.
It is likely that the Costa Rican people will confront the remaining challenges of
sustainable development with success.
The future of the biosphere may depend on our biophilia.
Not many people live in truly wild environments or even visit such places.
Biophilia includes our sense of connection to diverse organisms and our attachment to
pristine landscapes.
Most biologists have embraced this idea.