Sunteți pe pagina 1din 32

Chapter 9 Cellular Respiration: Harvesting Chemical Energy

Lecture Outline
Overview: Life Is Work
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 9.1 Catabolic pathways yield energy by oxidizing organic fuels
The arrangement of atoms of organic molecules represents potential energy.
Enzymes catalyze the systematic degradation of organic molecules that are rich in energy to simpler waste products
with less energy.
Some of the released energy is used to do work; the rest is dissipated as heat.
Catabolic metabolic pathways release the energy stored in complex organic molecules.
One type of catabolic process, fermentation, leads to the partial degradation of sugars in the absence of oxygen.
A more efficient and widespread catabolic process, cellular respiration, consumes oxygen as a reactant to complete
the breakdown of a variety of organic molecules.
In eukaryotic cells, mitochondria are the site of most of the processes of cellular respiration.
Cellular respiration is similar in broad principle to the combustion of gasoline in an automobile engine after oxygen
is mixed with hydrocarbon fuel.
Food is the fuel for respiration. The exhaust is carbon dioxide and water.
The overall process is:
organic compounds + O2 --> CO2 + H2O + energy (ATP + heat).
Carbohydrates, fats, and proteins can all be used as the fuel, but it is most useful to consider glucose.
C6H12O6 + 6O2 --> 6CO2 + 6H2O + Energy (ATP + heat)
The catabolism of glucose is exergonic with a ? G of ?686 kcal per mole of glucose.
Some of this energy is used to produce ATP, which can perform cellular work.
Redox reactions release energy when electrons move closer to electronegative atoms.
Catabolic pathways transfer the electrons stored in food molecules, releasing energy that is used to synthesize ATP.
Reactions that result in the transfer of one or more electrons from one reactant to another are oxidation-reduction
reactions, or redox reactions.
The loss of electrons is called oxidation.
The addition of electrons is called reduction.
The formation of table salt from sodium and chloride is a redox reaction.
Na + Cl --> Na+ + Cl?
Here sodium is oxidized and chlorine is reduced (its charge drops from 0 to ?1).
More generally: Xe? + Y --> X + Ye?
X, the electron donor, is the reducing agent and reduces Y.
Y, the electron recipient, is the oxidizing agent and oxidizes X.
Redox reactions require both a donor and acceptor.
Redox reactions also occur when the transfer of electrons is not complete but involves a change in the degree of
electron sharing in covalent bonds.
In the combustion of methane to form water and carbon dioxide, the nonpolar covalent bonds of methane (CH)
and oxygen (O=O) are converted to polar covalent bonds (C=O and OH).
When methane reacts with oxygen to form carbon dioxide, electrons end up farther away from the carbon atom and
closer to their new covalent partners, the oxygen atoms, which are very electronegative.
In effect, the carbon atom has partially lost its shared electrons. Thus, methane has been oxidized.
The two atoms of the oxygen molecule share their electrons equally. When oxygen reacts with the hydrogen from
methane to form water, the electrons of the covalent bonds are drawn closer to the oxygen.
In effect, each oxygen atom has partially gained electrons, and so the oxygen molecule has been reduced.
Oxygen is very electronegative, and is one of the most potent of all oxidizing agents.
Energy must be added to pull an electron away from an atom.
The more electronegative the atom, the more energy is required to take an electron away from it.
An electron loses potential energy when it shifts from a less electronegative atom toward a more electronegative
one.
A redox reaction that relocates electrons closer to oxygen, such as the burning of methane, releases chemical energy
that can do work.
The fall of electrons during respiration is stepwise, via NAD+ and an electron transport chain.
Cellular respiration does not oxidize glucose in a single step that transfers all the hydrogen in the fuel to oxygen at
one time.
Rather, glucose and other fuels are broken down in a series of steps, each catalyzed by a specific enzyme.
At key steps, electrons are stripped from the glucose.
In many oxidation reactions, the electron is transferred with a proton, as a hydrogen atom.
The hydrogen atoms are not transferred directly to oxygen but are passed first to a coenzyme called NAD+
(nicotinamide adenine dinucleotide).
How does NAD+ trap electrons from glucose?
Dehydrogenase enzymes strip two hydrogen atoms from the fuel (e.g., glucose), oxidizing it.
The enzyme passes two electrons and one proton to NAD+.
The other proton is released as H+ to the surrounding solution.
By receiving two electrons and only one proton, NAD+ has its charge neutralized when it is reduced to NADH.
NAD+ functions as the oxidizing agent in many of the redox steps during the catabolism of glucose.
The electrons carried by NADH have lost very little of their potential energy in this process.
Each NADH molecule formed during respiration represents stored energy. This energy is tapped to synthesize ATP
as electrons fall from NADH to oxygen.
How are electrons extracted from food and stored by NADH finally transferred to oxygen?
Unlike the explosive release of heat energy that occurs when H2 and O2 are combined (with a spark for activation
energy), cellular respiration uses an electron transport chain to break the fall of electrons to O2 into several steps.
The electron transport chain consists of several molecules (primarily proteins) built into the inner membrane of a
mitochondrion.
Electrons released from food are shuttled by NADH to the top higher-energy end of the chain.
At the bottom lower-energy end, oxygen captures the electrons along with H+ to form water.
Electron transfer from NADH to oxygen is an exergonic reaction with a free energy change of ?53 kcal/mol.
Electrons are passed to increasingly electronegative molecules in the chain until they reduce oxygen, the most
electronegative receptor.
In summary, during cellular respiration, most electrons travel the following downhill route: food --> NADH -->
electron transport chain --> oxygen.
These are the stages of cellular respiration: a preview.
Respiration occurs in three metabolic stages: glycolysis, the citric acid cycle, and the electron transport chain and
oxidative phosphorylation.
Glycolysis occurs in the cytoplasm.
It begins catabolism by breaking glucose into two molecules of pyruvate.
The citric acid cycle occurs in the mitochondrial matrix.
It completes the breakdown of glucose by oxidizing a derivative of pyruvate to carbon dioxide.
Several steps in glycolysis and the citric acid cycle are redox reactions in which dehydrogenase enzymes transfer
electrons from substrates to NAD+, forming NADH.
NADH passes these electrons to the electron transport chain.
In the electron transport chain, the electrons move from molecule to molecule until they combine with molecular
oxygen and hydrogen ions to form water.
As they are passed along the chain, the energy carried by these electrons is transformed in the mitochondrion into a
form that can be used to synthesize ATP via oxidative phosphorylation.
The inner membrane of the mitochondrion is the site of electron transport and chemiosmosis, processes that together
constitute oxidative phosphorylation.
Oxidative phosphorylation produces almost 90% of the ATP generated by respiration.
Some ATP is also formed directly during glycolysis and the citric acid cycle by substrate-level phosphorylation.
Here an enzyme transfers a phosphate group from an organic substrate to ADP, forming ATP.
For each molecule of glucose degraded to carbon dioxide and water by respiration, the cell makes up to 38 ATP,
each with 7.3 kcal/mol of free energy.
Respiration uses the small steps in the respiratory pathway to break the large denomination of energy contained in
glucose into the small change of ATP.
The quantity of energy in ATP is more appropriate for the level of work required in the cell.
Concept 9.2 Glycolysis harvests chemical energy by oxidizing glucose to pyruvate
During glycolysis, glucose, a six carbon-sugar, is split into two three-carbon sugars.
These smaller sugars are oxidized and rearranged to form two molecules of pyruvate, the ionized form of pyruvic
acid.
Each of the ten steps in glycolysis is catalyzed by a specific enzyme.
These steps can be divided into two phases: an energy investment phase and an energy payoff phase.
In the energy investment phase, the cell invests ATP to provide activation energy by phosphorylating glucose.
This requires 2 ATP per glucose.
In the energy payoff phase, ATP is produced by substrate-level phosphorylation and NAD+ is reduced to NADH by
electrons released by the oxidation of glucose.
The net yield from glycolysis is 2 ATP and 2 NADH per glucose.
No CO2 is produced during glycolysis.
Glycolysis can occur whether O2 is present or not.
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.
After pyruvate enters the mitochondrion via active transport, it is converted to a compound called acetyl coenzyme
A or acetyl CoA.
This step is accomplished by a multienzyme complex that catalyzes three reactions:
1. A carboxyl group is removed as CO2.
2. The remaining two-carbon fragment is oxidized to form acetate. An enzyme transfers the pair of electrons to NAD+
to form NADH.
3. Acetate combines with coenzyme A to form the very reactive molecule acetyl CoA.
Acetyl CoA is now ready to feed its acetyl group into the citric acid cycle for further oxidation.
The citric acid cycle is also called the Krebs cycle in honor of Hans Krebs, who was largely responsible for
elucidating its pathways in the 1930s.
The citric acid cycle oxidizes organic fuel derived from pyruvate.
The citric acid cycle has eight steps, each catalyzed by a specific enzyme.
The acetyl group of acetyl CoA joins the cycle by combining with the compound oxaloacetate, forming citrate.
The next seven steps decompose the citrate back to oxaloacetate. It is the regeneration of oxaloacetate that makes
this process a cycle.
Three CO2 molecules are released, including the one released during the conversion of pyruvate to acetyl CoA.
The cycle generates one ATP per turn by substrate-level phosphorylation.
A GTP molecule is formed by substrate-level phosphorylation.
The GTP is then used to synthesize an ATP, the only ATP generated directly by the citric acid cycle.
Most of the chemical energy is transferred to NAD+ and FAD during the redox reactions.
The reduced coenzymes NADH and FADH2 then transfer high-energy electrons to the electron transport chain.
Each cycle produces one ATP by substrate-level phosphorylation, three NADH, and one FADH2 per acetyl CoA.
Concept 9.4 During oxidative phosphorylation, chemiosmosis couples electron transport to ATP synthesis
The inner mitochondrial membrane couples electron transport to ATP synthesis.
Only 4 of 38 ATP ultimately produced by respiration of glucose are produced by substrate-level phosphorylation.
Two are produced during glycolysis, and 2 are produced during the citric acid cycle.
NADH and FADH2 account for the vast majority of the energy extracted from the food.
These reduced coenzymes link glycolysis and the citric acid cycle to oxidative phosphorylation, which uses energy
released by the electron transport chain to power ATP synthesis.
The electron transport chain is a collection of molecules embedded in the cristae, the folded inner membrane of the
mitochondrion.
The folding of the cristae increases its surface area, providing space for thousands of copies of the chain in each
mitochondrion.
Most components of the chain are proteins bound to prosthetic groups, nonprotein components essential for
catalysis.
Electrons drop in free energy as they pass down the electron transport chain.
During electron transport along the chain, electron carriers alternate between reduced and oxidized states as they
accept and donate electrons.
Each component of the chain becomes reduced when it accepts electrons from its uphill neighbor, which is less
electronegative.
It then returns to its oxidized form as it passes electrons to its more electronegative downhill neighbor.
Electrons carried by NADH are transferred to the first molecule in the electron transport chain, a flavoprotein.
The electrons continue along the chain that includes several cytochrome proteins and one lipid carrier.
The prosthetic group of each cytochrome is a heme group with an iron atom that accepts and donates electrons.
The last cytochrome of the chain, cyt a3, passes its electrons to oxygen, which is very electronegative.
Each oxygen atom also picks up a pair of hydrogen ions from the aqueous solution to form water.
For every two electron carriers (four electrons), one O2 molecule is reduced to two molecules of water.
The electrons carried by FADH2 have lower free energy and are added at a lower energy level than those carried by
NADH.
The electron transport chain provides about one-third less energy for ATP synthesis when the electron donor is
FADH2 rather than NADH.
The electron transport chain generates no ATP directly.
Its function is to break the large free energy drop from food to oxygen into a series of smaller steps that release
energy in manageable amounts.
How does the mitochondrion couple electron transport and energy release to ATP synthesis?
The answer is a mechanism called chemiosmosis.
A protein complex, ATP synthase, in the cristae actually makes ATP from ADP and Pi.
ATP uses the energy of an existing proton gradient to power ATP synthesis.
The proton gradient develops between the intermembrane space and the matrix.
The proton gradient is produced by the movement of electrons along the electron transport chain.
The chain is an energy converter that uses the exergonic flow of electrons to pump H+ from the matrix into the
intermembrane space.
The protons pass back to the matrix through a channel in ATP synthase, using the exergonic flow of H+ to drive the
phosphorylation of ADP.
Thus, the energy stored in a H+ gradient across a membrane couples the redox reactions of the electron transport
chain to ATP synthesis.
From studying the structure of ATP synthase, scientists have learned how the flow of H+ through this large enzyme
powers ATP generation.
ATP synthase is a multisubunit complex with four main parts, each made up of multiple polypeptides:
1. A rotor in the inner mitochondrial membrane.
2. A knob that protrudes into the mitochondrial matrix.
3. An internal rod extending from the rotor into the knob.
4. A stator, anchored next to the rotor, which holds the knob stationary.
Protons flow down a narrow space between the stator and rotor, causing the rotor and its attached rod to rotate.
The spinning rod causes conformational changes in the stationary knob, activating three catalytic sites in the knob
where ADP and inorganic phosphate combine to make ATP.
How does the inner mitochondrial membrane generate and maintain the H+ gradient that drives ATP synthesis in the
ATP synthase protein complex?
Creating the H+ gradient is the function of the electron transport chain.
The ETC is an energy converter that uses the exergonic flow of electrons to pump H+ across the membrane from the
mitochondrial matrix to the intermembrane space.
The H+ has a tendency to diffuse down its gradient.
The ATP synthase molecules are the only place that H+ can diffuse back to the matrix.
The exergonic flow of H+ is used by the enzyme to generate ATP.
This coupling of the redox reactions of the electron transport chain to ATP synthesis is called chemiosmosis.
How does the electron transport chain pump protons?
Certain members of the electron transport chain accept and release H+ along with electrons.
At certain steps along the chain, electron transfers cause H+ to be taken up and released into the surrounding
solution.
The electron carriers are spatially arranged in the membrane in such a way that protons are accepted from the
mitochondrial matrix and deposited in the intermembrane space.
The H+ gradient that results is the proton-motive force.
The gradient has the capacity to do work.
Chemiosmosis is an energy-coupling mechanism that uses energy stored in the form of an H+ gradient across a
membrane to drive cellular work.
In mitochondria, the energy for proton gradient formation comes from exergonic redox reactions, and ATP synthesis
is the work performed.
Chemiosmosis in chloroplasts also generates ATP, but light drives the electron flow down an electron transport
chain and H+ gradient formation.
Prokaryotes generate H+ gradients across their plasma membrane.
They can use this proton-motive force not only to generate ATP, but also to pump nutrients and waste products
across the membrane and to rotate their flagella.
Here is an accounting of ATP production by cellular respiration.
During cellular respiration, most energy flows from glucose --> NADH --> electron transport chain --> proton-
motive force --> ATP.
Lets consider the products generated when cellular respiration oxidizes a molecule of glucose to six CO2
molecules.
Four ATP molecules are produced by substrate-level phosphorylation during glycolysis and the citric acid cycle.
Many more ATP molecules are generated by oxidative phosphorylation.
Each NADH from the citric acid cycle and the conversion of pyruvate contributes enough energy to the proton-
motive force to generate a maximum of 3 ATP.
The NADH from glycolysis may also yield 3 ATP.
Each FADH2 from the citric acid cycle can be used to generate about 2 ATP.
Why is our accounting so inexact?
There are three reasons that we cannot state an exact number of ATP molecules generated by one molecule of
glucose.
1. Phosphorylation and the redox reactions are not directly coupled to each other, so the ratio of number of NADH to
number of ATP is not a whole number.
One NADH results in 10 H+ being transported across the inner mitochondrial membrane.
Between 3 and 4 H+ must reenter the mitochondrial matrix via ATP synthase to generate 1 ATP.
Therefore, 1 NADH generates enough proton-motive force for synthesis of 2.5 to 3.3 ATP.
We round off and say that 1 NADH generates 3 ATP.
2. The ATP yield varies slightly depending on the type of shuttle used to transport electrons from the cytosol into the
mitochondrion.
The mitochondrial inner membrane is impermeable to NADH, so the two electrons of the NADH produced in
glycolysis must be conveyed into the mitochondrion by one of several electron shuttle systems.
In some shuttle systems, the electrons are passed to NAD+, which generates 3 ATP. In others, the electrons are
passed to FAD, which generates only 2 ATP.
3. The proton-motive force generated by the redox reactions of respiration may drive other kinds of work, such as
mitochondrial uptake of pyruvate from the cytosol.
If all the proton-motive force generated by the electron transport chain were used to drive ATP synthesis, one
glucose molecule could generate a maximum of 34 ATP by oxidative phosphorylation plus 4 ATP (net) from
substrate-level phosphorylation to give a total yield of 3638 ATP (depending on the efficiency of the shuttle).
How efficient is respiration in generating ATP?
Complete oxidation of glucose releases 686 kcal/mol.
Phosphorylation of ADP to form ATP requires at least 7.3 kcal/mol.
Efficiency of respiration is 7.3 kcal/mol times 38 ATP/glucose divided by 686 kcal/mol glucose, which equals 0.4 or
40%.
Approximately 60% of the energy from glucose is lost as heat.
Some of that heat is used to maintain our high body temperature (37C).
Cellular respiration is remarkably efficient in energy conversion.
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.
However, fermentation provides a mechanism by which some cells can oxidize organic fuel and generate ATP
without the use of oxygen.
In glycolysis, glucose is oxidized to two pyruvate molecules with NAD+ as the oxidizing agent.
Glycolysis is exergonic and produces 2 ATP (net).
If oxygen is present, additional ATP can be generated when NADH delivers its electrons to the electron transport
chain.
Glycolysis generates 2 ATP whether oxygen is present (aerobic) or not (anaerobic).
Anaerobic catabolism of sugars can occur by fermentation.
Fermentation can generate ATP from glucose by substrate-level phosphorylation as long as there is a supply of
NAD+ to accept electrons.
If the NAD+ pool is exhausted, glycolysis shuts down.
Under aerobic conditions, NADH transfers its electrons to the electron transfer chain, recycling NAD+.
Under anaerobic conditions, various fermentation pathways generate ATP by glycolysis and recycle NAD+ by
transferring electrons from NADH to pyruvate or derivatives of pyruvate.
In alcohol fermentation, pyruvate is converted to ethanol in two steps.
First, pyruvate is converted to a two-carbon compound, acetaldehyde, by the removal of CO2.
Second, acetaldehyde is reduced by NADH to ethanol.
Alcohol fermentation by yeast is used in brewing and winemaking.
During lactic acid fermentation, pyruvate is reduced directly by NADH to form lactate (the ionized form of lactic
acid) without release of CO2.
Lactic acid fermentation by some fungi and bacteria is used to make cheese and yogurt.
Human muscle cells switch from aerobic respiration to lactic acid fermentation to generate ATP when O2 is scarce.
The waste product, lactate, may cause muscle fatigue, but ultimately it is converted back to pyruvate in the liver.
Fermentation and cellular respiration are anaerobic and aerobic alternatives, respectively, for producing ATP from
sugars.
Both use glycolysis to oxidize sugars to pyruvate with a net production of 2 ATP by substrate-level phosphorylation.
Both use NAD+ as an oxidizing agent to accept electrons from food during glycolysis.
The two processes differ in their mechanism for oxidizing NADH to NAD+.
In fermentation, the electrons of NADH are passed to an organic molecule to regenerate NAD+.
In respiration, the electrons of NADH are ultimately passed to O2, generating ATP by oxidative phosphorylation.
More ATP is generated from the oxidation of pyruvate in the citric acid cycle.
Without oxygen, the energy still stored in pyruvate is unavailable to the cell.
Under aerobic respiration, a molecule of glucose yields 38 ATP, but the same molecule of glucose yields only 2
ATP under anaerobic respiration.
Yeast and many bacteria are facultative anaerobes that can survive using either fermentation or respiration.
At a cellular level, human muscle cells can behave as facultative anaerobes.
For facultative anaerobes, pyruvate is a fork in the metabolic road that leads to two alternative routes.
Under aerobic conditions, pyruvate is converted to acetyl CoA and oxidation continues in the citric acid cycle.
Under anaerobic conditions, pyruvate serves as an electron acceptor to recycle NAD+.
The oldest bacterial fossils are more than 3.5 billion years old, appearing long before appreciable quantities of O2
accumulated in the atmosphere.
Therefore, the first prokaryotes may have generated ATP exclusively from glycolysis.
The fact that glycolysis is a ubiquitous metabolic pathway and occurs in the cytosol without membrane-enclosed
organelles suggests that glycolysis evolved early in the history of life.
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.
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.

Lecture Outline for Campbell/Reece Biology, 7th Edition, Pearson Education, Inc. 9-1
Chapter 12 The Cell Cycle
Lecture Outline
Overview: The Key Roles of Cell Division
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 12.1 Cell division results in genetically identical daughter cells
Cell division requires the distribution of identical genetic materialDNAto two daughter cells.
What is remarkable is the fidelity with which DNA is passed along, without dilution, from one generation to the
next.
A dividing cell duplicates its DNA, allocates the two copies to opposite ends of the cell, and then splits into two
daughter cells.
A cells genetic information, packaged as DNA, is called its genome.
In prokaryotes, the genome is often a single long DNA molecule.
In eukaryotes, the genome consists of several DNA molecules.
A human cell must duplicate about 2 m of DNA and separate the two copies such that each daughter cell ends up
with a complete genome.
DNA molecules are packaged into chromosomes.
Every eukaryotic species has a characteristic number of chromosomes in each cell nucleus.
Human somatic cells (body cells) have 46 chromosomes, made up of two sets of 23 (one from each parent).
Human gametes (sperm or eggs) have one set of 23 chromosomes, half the number in a somatic cell.
Eukaryotic chromosomes are made of chromatin, a complex of DNA and associated protein.
Each single chromosome contains one long, linear DNA molecule carrying hundreds or thousands of genes, the
units that specify an organisms inherited traits.
The associated proteins maintain the structure of the chromosome and help control gene activity.
When a cell is not dividing, each chromosome is in the form of a long, thin chromatin fiber.
Before cell division, chromatin condenses, coiling and folding to make a smaller package.
Each duplicated chromosome consists of two sister chromatids, which contain identical copies of the chromosomes
DNA.
The chromatids are initially attached by adhesive proteins along their lengths.
As the chromosomes condense, the region where the chromatids connect shrinks to a narrow area, the centromere.
Later in cell division, the sister chromatids are pulled apart and repackaged into two new nuclei at opposite ends of
the parent cell.
Once the sister chromatids separate, they are considered individual chromosomes.
Mitosis, the formation of the two daughter nuclei, is usually followed by division of the cytoplasm, cytokinesis.
These processes start with one cell and produce two cells that are genetically identical to the original parent cell.
Each of us inherited 23 chromosomes from each parent: one set in an egg and one set in sperm.
The fertilized egg, or zygote, underwent cycles of mitosis and cytokinesis to produce a fully developed multicellular
human made up of 200 trillion somatic cells.
These processes continue every day to replace dead and damaged cells.
Essentially, these processes produce clonescells with identical genetic information.
In contrast, gametes (eggs or sperm) are produced only in gonads (ovaries or testes) by a variation of cell division
called meiosis.
Meiosis yields four nonidentical daughter cells, each with half the chromosomes of the parent.
In humans, meiosis reduces the number of chromosomes from 46 to 23.
Fertilization fuses two gametes together and doubles the number of chromosomes to 46 again.
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.
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 1012 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 cells 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.
The mitotic spindle, fibers composed of microtubules and associated proteins, is a major driving force in mitosis.
As the spindle assembles during prophase, the elements come from partial disassembly of the cytoskeleton.
The spindle fibers elongate by incorporating more subunits of the protein tubulin.
Assembly of the spindle microtubules starts in the centrosome.
The centrosome (microtubule-organizing center) is a nonmembranous organelle that organizes the cells
microtubules.
In animal cells, the centrosome has a pair of centrioles at the center, but the centrioles are not essential for cell
division.
During interphase, the single centrosome replicates to form two centrosomes.
As mitosis starts, the two centrosomes are located near the nucleus.
As the spindle microtubules grow from them, the centrioles are pushed apart.
By the end of prometaphase, they are at opposite ends of the cell.
An aster, a radial array of short microtubules, extends from each centrosome.
The spindle includes the centrosomes, the spindle microtubules, and the asters.
Each sister chromatid has a kinetochore of proteins and chromosomal DNA at the centromere.
The kinetochores of the joined sister chromatids face in opposite directions.
During prometaphase, some spindle microtubules (called kinetochore microtubules) attach to the kinetochores.
When a chromosomes kinetochore is captured by microtubules, the chromosome moves toward the pole from
which those microtubules come.
When microtubules attach to the other pole, this movement stops and a tug-of-war ensues.
Eventually, the chromosome settles midway between the two poles of the cell, on the metaphase plate.
Nonkinetochore microtubules from opposite poles overlap and interact with each other.
By metaphase, the microtubules of the asters have grown and are in contact with the plasma membrane.
The spindle is now complete.
Anaphase commences when the proteins holding the sister chromatids together are inactivated.
Once the chromosomes are separate, full-fledged chromosomes, they move toward opposite poles of the cell.
How do the kinetochore microtubules function into the poleward movement of chromosomes?
One hypothesis is that the chromosomes are reeled in by the shortening of microtubules at the spindle poles.
Experimental evidence supports the hypothesis that motor proteins on the kinetochore walk the attached
chromosome along the microtubule toward the nearest pole.
Meanwhile, the excess microtubule sections depolymerize at their kinetochore ends.
What is the function of the nonkinetochore microtubules?
Nonkinetochore microtubules are responsible for lengthening the cell along the axis defined by the poles.
These microtubules interdigitate and overlap across the metaphase plate.
During anaphase, the area of overlap is reduced as motor proteins attached to the microtubules walk them away from
one another, using energy from ATP.
As microtubules push apart, the microtubules lengthen by the addition of new tubulin monomers to their overlapping
ends, allowing continued overlap.
Cytokinesis divides the cytoplasm: a closer look.
Cytokinesis, division of the cytoplasm, typically follows mitosis.
In animal cells, cytokinesis occurs by a process called cleavage.
The first sign of cleavage is the appearance of a cleavage furrow in the cell surface near the old metaphase plate.
On the cytoplasmic side of the cleavage furrow is a contractile ring of actin microfilaments associated with
molecules of the motor protein myosin.
Contraction of the ring pinches the cell in two.
Cytokinesis in plants, which have cell walls, involves a completely different mechanism.
During telophase, vesicles from the Golgi coalesce at the metaphase plate, forming a cell plate.
The plate enlarges until its membranes fuse with the plasma membrane at the perimeter.
The contents of the vesicles form new cell wall material between the daughter cells.
Mitosis in eukaryotes may have evolved from binary fission in bacteria.
Prokaryotes reproduce by binary fission, not mitosis.
Most bacterial genes are located on a single bacterial chromosome that consists of a circular DNA molecule and
associated proteins.
While bacteria are smaller and simpler than eukaryotic cells, they still have large amounts of DNA that must be
copied and distributed equally to two daughter cells.
The circular bacterial chromosome is highly folded and coiled in the cell.
In binary fission, chromosome replication begins at one point in the circular chromosome, the origin of replication
site, producing two origins.
As the chromosome continues to replicate, one origin moves toward each end of the cell.
While the chromosome is replicating, the cell elongates.
When replication is complete, its plasma membrane grows inward to divide the parent cell into two daughter cells,
each with a complete genome.
Researchers have developed methods to allow them to observe the movement of bacterial chromosomes.
The movement is similar to the poleward movements of the centromere regions of eukaryotic chromosomes.
However, bacterial chromosomes lack visible mitotic spindles or even microtubules.
The mechanism behind the movement of the bacterial chromosome is becoming clearer but is still not fully
understood.
Several proteins have been identified and play important roles.
How did mitosis evolve?
There is evidence that mitosis had its origins in bacterial binary fission.
Some of the proteins involved in binary fission are related to eukaryotic proteins.
Two of these are related to eukaryotic tubulin and actin proteins.
As eukaryotes evolved, the ancestral process of binary fission gave rise to mitosis.
Possible intermediate evolutionary steps are seen in the division of two types of unicellular algae.
In dinoflagellates, replicated chromosomes are attached to the nuclear envelope.
In diatoms, the spindle develops within the nucleus.
In most eukaryotic cells, the nuclear envelope breaks down and a spindle separates the chromosomes.
Concept 12.3 The cell cycle is regulated by a molecular control system
The timing and rates of cell division in different parts of an animal or plant are crucial for normal growth,
development, and maintenance.
The frequency of cell division varies with cell type.
Some human cells divide frequently throughout life (skin cells).
Others have the ability to divide, but keep it in reserve (liver cells).
Mature nerve and muscle cells do not appear to divide at all after maturity.
Investigation of the molecular mechanisms regulating these differences provide important insights into the operation
of normal cells, and may also explain cancer cells escape controls.
Cytoplasmic signals drive the cell cycle.
The cell cycle appears to be driven by specific chemical signals present in the cytoplasm.
Some of the initial evidence for this hypothesis came from experiments in which cultured mammalian cells at
different phases of the cell cycle were fused to form a single cell with two nuclei.
Fusion of an S phase cell and a G1 phase cell induces the G1 nucleus to start S phase.
This suggests that chemicals present in the S phase nucleus stimulated the fused cell.
Fusion of a cell in mitosis (M phase) with one in interphase (even G1 phase) induces the second cell to enter mitosis.
The sequential events of the cell cycle are directed by a distinct cell cycle control system.
Cyclically operating molecules trigger and coordinate key events in the cell cycle.
The control cycle has a built-in clock, but it is also regulated by external adjustments and internal controls.
A checkpoint in the cell cycle is a critical control point where stop and go-ahead signals regulate the cycle.
The signals are transmitted within the cell by signal transduction pathways.
Animal cells generally have built-in stop signals that halt the cell cycle at checkpoints until overridden by go-ahead
signals.
Many signals registered at checkpoints come from cellular surveillance mechanisms.
These indicate whether key cellular processes have been completed correctly.
Checkpoints also register signals from outside the cell.
Three major checkpoints are found in the G1, G2, and M phases.
For many cells, the G1 checkpoint, the restriction point in mammalian cells, is the most important.
If the cell receives a go-ahead signal at the G1 checkpoint, it usually completes the cell cycle and divides.
If it does not receive a go-ahead signal, the cell exits the cycle and switches to a nondividing state, the G0 phase.
Most cells in the human body are in this phase.
Liver cells can be called back to the cell cycle by external cues, such as growth factors released during injury.
Highly specialized nerve and muscle cells never divide.
Rhythmic fluctuations in the abundance and activity of cell cycle control molecules pace the events of the cell cycle.
These regulatory molecules include protein kinases that activate or deactivate other proteins by phosphorylating
them.
These kinases are present in constant amounts but require attachment of a second protein, a cyclin, to become
activated.
Levels of cyclin proteins fluctuate cyclically.
Because of the requirement for binding of a cyclin, the kinases are called cyclin-dependent kinases, or Cdks.
Cyclin levels rise sharply throughout interphase, and then fall abruptly during mitosis.
Peaks in the activity of one cyclin-Cdk complex, MPF, correspond to peaks in cyclin concentration.
MPF (maturation-promoting factor or M-phase-promoting-factor) triggers the cells passage past the G2
checkpoint to the M phase.
MPF promotes mitosis by phosphorylating a variety of other protein kinases.
MPF stimulates fragmentation of the nuclear envelope by phosphorylation of various proteins of the nuclear lamina.
It also triggers the breakdown of cyclin, dropping cyclin and MPF levels during mitosis and inactivating MPF.
The noncyclin part of MPF, the Cdk, persists in the cell in inactive form until it associates with new cyclin
molecules synthesized during the S and G2 phases of the next round of the cycle.
At least three Cdk proteins and several cyclins regulate the key G1 checkpoint.
Similar mechanisms are also involved in driving the cell cycle past the M phase checkpoint.
Internal and external cues help regulate the cell cycle.
While research scientists know that active Cdks function by phosphorylating proteins, the identity of all these
proteins is still under investigation.
Scientists do not yet know what Cdks actually do in most cases.
Some steps in the signaling pathways that regulate the cell cycle are clear.
Some signals originate inside the cell, others outside.
The M phase checkpoint ensures that all the chromosomes are properly attached to the spindle at the metaphase
plate before anaphase.
This ensures that daughter cells do not end up with missing or extra chromosomes.
A signal to delay anaphase originates at kinetochores that have not yet attached to spindle microtubules.
This keeps the anaphase-promoting complex (APC) in an inactive state.
When all kinetochores are attached, the APC activates, triggering breakdown of cyclin and inactivation of proteins
holding sister chromatids together.
A variety of external chemical and physical factors can influence cell division.
For example, cells fail to divide if an essential nutrient is left out of the culture medium.
Particularly important for mammalian cells are growth factors, proteins released by one group of cells that stimulate
other cells to divide.
For example, platelet-derived growth factors (PDGF), produced by platelet blood cells, bind to tyrosine-kinase
receptors of fibroblasts, a type of connective tissue cell.
This triggers a signal-transduction pathway that allows cells to pass the G1 checkpoint and divide.
Each cell type probably responds specifically to a certain growth factor or combination of factors.
The role of PDGF is easily seen in cell culture.
Fibroblasts in culture will only divide in the presence of a medium that also contains PDGF.
In a living organism, platelets release PDGF in the vicinity of an injury.
The resulting proliferation of fibroblasts helps heal the wound.
At least 50 different growth factors can trigger specific cells to divide.
The effect of an external physical factor on cell division can be seen in density-dependent inhibition of cell division.
Cultured cells normally divide until they form a single layer on the inner surface of the culture container.
If a gap is created, the cells will grow to fill the gap.
At high densities, the amount of growth factors and nutrients is insufficient to allow continued cell growth.
Most animal cells also exhibit anchorage dependence for cell division.
To divide, they must be anchored to a substratum, typically the extracellular matrix of a tissue.
Control appears to be mediated by pathways involving plasma membrane proteins and elements of the cytoskeleton
linked to them.
Cancer cells exhibit neither density-dependent inhibition nor anchorage dependence.
Cancer cells have escaped from cell cycle controls.
Cancer cells divide excessively and invade other tissues because they are free of the bodys control mechanisms.
Cancer cells do not stop dividing when growth factors are depleted.
This is either because a cancer cell manufactures its own growth factors, has an abnormality in the signaling
pathway, or has an abnormal cell cycle control system.
If and when cancer cells stop dividing, they do so at random points, not at the normal checkpoints in the cell cycle.
Cancer cells may divide indefinitely if they have a continual supply of nutrients.
In contrast, nearly all mammalian cells divide 20 to 50 times under culture conditions before they stop, age, and die.
Cancer cells may be immortal.
HeLa cells from a tumor removed from a woman (Henrietta Lacks) in 1951 are still reproducing in culture.
The abnormal behavior of cancer cells begins when a single cell in a tissue undergoes a transformation that converts
it from a normal cell to a cancer cell.
Normally, the immune system recognizes and destroys transformed cells.
However, cells that evade destruction proliferate to form a tumor, a mass of abnormal cells.
If the abnormal cells remain at the originating site, the lump is called a benign tumor.
Most do not cause serious problems and can be fully removed by surgery.
In a malignant tumor, the cells become invasive enough to impair the functions of one or more organs.
In addition to chromosomal and metabolic abnormalities, cancer cells often lose attachment to nearby cells, are
carried by the blood and lymph system to other tissues, and start more tumors in an event called metastasis.
Cancer cells are abnormal in many ways.
They may have an unusual number of chromosomes, their metabolism may be disabled, and they may cease to
function in any constructive way.
Cancer cells may secrete signal molecules that cause blood vessels to grow toward the tumor.
Treatments for metastasizing cancers include high-energy radiation and chemotherapy with toxic drugs.
These treatments target actively dividing cells.
Chemotherapeutic drugs interfere with specific steps in the cell cycle.
For example, Taxol prevents mitotic depolymerization, preventing cells from proceeding past metaphase.
The side effects of chemotherapy are due to the drugs effects on normal cells.
Researchers are beginning to understand how a normal cell is transformed into a cancer cell.
The causes are diverse, but cellular transformation always involves the alteration of genes that influence the cell
cycle control system.
Lecture Outline for Campbell/Reece Biology, 7th Edition, Pearson Education, Inc. 12-1
Chapter 13 Meiosis and Sexual Life Cycles
Lecture Outline
Overview: Hereditary Similarity and Variation
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 13.1 Offspring acquire genes from parents by inheriting chromosomes
Parents endow their offspring with coded information in the form of genes.
Your genome is comprised of the tens of thousands of genes that you inherited from your mother and your father.
Genes program specific traits that emerge as we develop from fertilized eggs into adults.
Genes are segments of DNA. Genetic information is transmitted as specific sequences of the four
deoxyribonucleotides in DNA.
This is analogous to the symbolic information of language in which words and sentences are translated into mental
images.
Cells translate genetic sentences into freckles and other features with no resemblance to genes.
Most genes program cells to synthesize specific enzymes and other proteins whose cumulative action produces an
organisms inherited traits.
The transmission of hereditary traits has its molecular basis in the precise replication of DNA.
This produces copies of genes that can be passed from parents to offspring.
In plants and animals, sperm and ova (unfertilized eggs) transmit genes from one generation to the next.
After fertilization (fusion of a sperm cell and an ovum), genes from both parents are present in the nucleus of the
fertilized egg, or zygote.
Almost all the DNA in a eukaryotic cell is subdivided into chromosomes in the nucleus.
Tiny amounts of DNA are also found in mitochondria and chloroplasts.
Every living species has a characteristic number of chromosomes.
Humans have 46 chromosomes in almost all of their cells.
Each chromosome consists of a single DNA molecule associated with various proteins.
Each chromosome has hundreds or thousands of genes, each at a specific location, its locus.
Like begets like, more or less: a comparison of asexual and sexual reproduction.
Only organisms that reproduce asexually can produce offspring that are exact copies of themselves.
In asexual reproduction, a single individual is the sole parent to donate genes to its offspring.
Single-celled eukaryotes can reproduce asexually by mitotic cell division to produce two genetically identical
daughter cells.
Some multicellular eukaryotes, like Hydra, can reproduce by budding, producing a mass of cells by mitosis.
An individual that reproduces asexually gives rise to a clone, a group of genetically identical individuals.
Members of a clone may be genetically different as a result of mutation.
In sexual reproduction, two parents produce offspring that have unique combinations of genes inherited from the
two parents.
Unlike a clone, offspring produced by sexual reproduction vary genetically from their siblings and their parents.
Concept 13.2 Fertilization and meiosis alternate in sexual life cycles
A life cycle is the generation-to-generation sequence of stages in the reproductive history of an organism.
It starts at the conception of an organism and continues until the organism produces its own offspring.
Human cells contain sets of chromosomes.
In humans, each somatic cell (all cells other than sperm or ovum) has 46 chromosomes.
Each chromosome can be distinguished by size, position of the centromere, and pattern of staining with certain dyes.
Images of the 46 human chromosomes can be arranged in pairs in order of size to produce a karyotype display.
The two chromosomes comprising a pair have the same length, centromere position, and staining pattern.
These homologous chromosome pairs carry genes that control the same inherited characters.
Two distinct sex chromosomes, the X and the Y, are an exception to the general pattern of homologous
chromosomes in human somatic cells.
The other 22 pairs are called autosomes.
The pattern of inheritance of the sex chromosomes determines an individuals sex.
Human females have a homologous pair of X chromosomes (XX).
Human males have an X and a Y chromosome (XY).
Only small parts of the X and Y are homologous.
Most of the genes carried on the X chromosome do not have counterparts on the tiny Y.
The Y chromosome also has genes not present on the X.
The occurrence of homologous pairs of chromosomes is a consequence of sexual reproduction.
We inherit one chromosome of each homologous pair from each parent.
The 46 chromosomes in each somatic cell are two sets of 23, a maternal set (from your mother) and a paternal set
(from your father).
The number of chromosomes in a single set is represented by n.
Any cell with two sets of chromosomes is called a diploid cell and has a diploid number of chromosomes,
abbreviated as 2n.
Sperm cells or ova (gametes) have only one set of chromosomes22 autosomes and an X (in an ovum) and 22
autosomes and an X or a Y (in a sperm cell).
A gamete with a single chromosome set is haploid, abbreviated as n.
Any sexually reproducing species has a characteristic haploid and diploid number of chromosomes.
For humans, the haploid number of chromosomes is 23 (n = 23), and the diploid number is 46 (2n = 46).
Lets discuss the role of meiosis in the human life cycle.
The human life cycle begins when a haploid sperm cell fuses with a haploid ovum.
These cells fuse (syngamy), resulting in fertilization.
The fertilized egg (zygote) is diploid because it contains two haploid sets of chromosomes bearing genes from the
maternal and paternal family lines.
As an organism develops from a zygote to a sexually mature adult, mitosis generates all the somatic cells of the
body.
Each somatic cell contains a full diploid set of chromosomes.
Gametes, which develop in the gonads (testes or ovaries), are not produced by mitosis.
If gametes were produced by mitosis, the fusion of gametes would produce offspring with four sets of chromosomes
after one generation, eight after a second, and so on.
Instead, gametes undergo the process of meiosis in which the chromosome number is halved.
Human sperm or ova have a haploid set of 23 different chromosomes, one from each homologous pair.
Fertilization restores the diploid condition by combining two haploid sets of chromosomes.
Organisms display a variety of sexual life cycles.
Fertilization and meiosis alternate in all sexual life cycles.
However, the timing of meiosis and fertilization does vary among species.
These variations can be grouped into three main types of life cycles.
In most animals, including humans, gametes are the only haploid cells.
Gametes do not divide but fuse to form a diploid zygote that divides by mitosis to produce a multicellular organism.
Plants and some algae have a second type of life cycle called alternation of generations.
This life cycle includes two multicellular stages, one haploid and one diploid.
The multicellular diploid stage is called the sporophyte.
Meiosis in the sporophyte produces haploid spores that develop by mitosis into the haploid gametophyte stage.
Gametes produced via mitosis by the gametophyte fuse to form the zygote, which grows into the sporophyte by
mitosis.
Most fungi and some protists have a third type of life cycle.
Gametes fuse to form a zygote, which is the only diploid phase.
The zygote undergoes meiosis to produce haploid cells.
These haploid cells grow by mitosis to form the haploid multicellular adult organism.
The haploid adult produces gametes by mitosis.
Note that either haploid or diploid cells can divide by mitosis, depending on the type of life cycle. However, only
diploid cells can undergo meiosis.
Although the three types of sexual life cycles differ in the timing of meiosis and fertilization, they share a
fundamental feature: each cycle of chromosome halving and doubling contributes to genetic variation among
offspring.
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.
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.
The chromosome number is reduced from diploid to haploid in meiosis but is conserved in mitosis.
Mitosis produces daughter cells that are genetically identical to the parent and to each other.
Meiosis produces cells that are genetically distinct from the parent cell and from each other.
Three events, unique to meiosis, occur during the first division cycle.
1. During prophase I of meiosis, replicated homologous chromosomes line up and become physically connected along
their lengths by a zipperlike protein complex, the synaptonemal complex, in a process called synapsis. Genetic
rearrangement between nonsister chromatids called crossing over also occurs. Once the synaptonemal complex is
disassembled, the joined homologous chromosomes are visible as a tetrad. X-shaped regions called chiasmata are
visible as the physical manifestation of crossing over. Synapsis and crossing over do not occur in mitosis.
2. At metaphase I of meiosis, homologous pairs of chromosomes align along the metaphase plate. In mitosis,
individual replicated chromosomes line up along the metaphase plate.
3. At anaphase I of meiosis, it is homologous chromosomes, not sister chromatids, that separate and are carried to
opposite poles of the cell. Sister chromatids of each replicated chromosome remain attached. In mitosis, sister
chromatids separate to become individual chromosomes.
Meiosis I is called the reductional division because it halves the number of chromosome sets per cella reduction
from the diploid to the haploid state.
The sister chromatids separate during the second meiosis division, meiosis II.
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:
1. Independent assortment of chromosomes.
2. Crossing over.
3. 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.
Therefore, the first meiotic division results in independent assortment of maternal and paternal chromosomes into
daughter cells.
The number of combinations possible when chromosomes assort independently into gametes is 2n, 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:
0. Independent assortment of homologous chromosomes during meiosis I and of nonidentical sister chromatids during
meiosis II.
1. Crossing over between homologous chromosomes during prophase I.
2. 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 populations 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.
If the environment changes or a population moves to a new environment, new genetic combinations that work best
in the new conditions will produce more offspring, and these genes will increase.
The formerly favored genes will decrease.
Sex and mutation continually generate new genetic variability.
Although Darwin realized that heritable variation makes evolution possible, he did not have a theory of inheritance.
Gregor Mendel, a contemporary of Darwins, published a theory of inheritance that supported Darwins theory.
However, this work was largely unknown until 1900, after Darwin and Mendel had both been dead for more than 15
years.
Lecture Outline for Campbell/Reece Biology, 7th Edition, Pearson Education, Inc. 13-1
Chapter 16 The Molecular Basis of Inheritance
Lecture Outline
Overview: Lifes Operating Instructions
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 16.1 DNA is the genetic material
The search for genetic material led to DNA.
Once T. H. Morgans group showed that genes are located on chromosomes, the two constituents of chromosomes
proteins and DNAwere the candidates for the genetic material.
Until the 1940s, the great heterogeneity and specificity of function of proteins seemed to indicate that proteins were
the genetic material.
However, this was not consistent with experiments with microorganisms, such as bacteria and viruses.
The discovery of the genetic role of DNA began with research by Frederick Griffith in 1928.
He studied Streptococcus pneumoniae, a bacterium that causes pneumonia in mammals.
One strain, the R strain, was harmless.
The other strain, the S strain, was pathogenic.
Griffith mixed heat-killed S strain with live R strain bacteria and injected this into a mouse.
The mouse died, and he recovered the pathogenic strain from the mouses blood.
Griffith called this phenomenon transformation, a phenomenon now defined as a change in genotype and phenotype
due to the assimilation of foreign DNA by a cell.
For the next 14 years, scientists tried to identify the transforming substance.
Finally in 1944, Oswald Avery, Maclyn McCarty, and Colin MacLeod announced that the transforming substance
was DNA.
Still, many biologists were skeptical.
Proteins were considered better candidates for the genetic material.
There was also a belief that the genes of bacteria could not be similar in composition and function to those of more
complex organisms.
Further evidence that DNA was the genetic material was derived from studies that tracked the infection of bacteria
by viruses.
Viruses consist of DNA (or sometimes RNA) enclosed by a protective coat of protein.
To replicate, a virus infects a host cell and takes over the cells metabolic machinery.
Viruses that specifically attack bacteria are called bacteriophages or just phages.
In 1952, Alfred Hershey and Martha Chase showed that DNA was the genetic material of the phage T2.
The T2 phage, consisting almost entirely of DNA and protein, attacks Escherichia coli (E. coli), a common intestinal
bacteria of mammals.
This phage can quickly turn an E. coli cell into a T2-producing factory that releases phages when the cell ruptures.
To determine the source of genetic material in the phage, Hershey and Chase designed an experiment in which they
could label protein or DNA and then track which entered the E. coli cell during infection.
They grew one batch of T2 phage in the presence of radioactive sulfur, marking the proteins but not DNA.
They grew another batch in the presence of radioactive phosphorus, marking the DNA but not proteins.
They allowed each batch to infect separate E. coli cultures.
Shortly after the onset of infection, they spun the cultured infected cells in a blender, shaking loose any parts of the
phage that remained outside the bacteria.
The mixtures were spun in a centrifuge, which separated the heavier bacterial cells in the pellet from lighter free
phages and parts of phage in the liquid supernatant.
They then tested the pellet and supernatant of the separate treatments for the presence of radioactivity.
Hershey and Chase found that when the bacteria had been infected with T2 phages that contained radiolabeled
proteins, most of the radioactivity was in the supernatant that contained phage particles, not in the pellet with the
bacteria.
When they examined the bacterial cultures with T2 phage that had radiolabeled DNA, most of the radioactivity was
in the pellet with the bacteria.
Hershey and Chase concluded that the injected DNA of the phage provides the genetic information that makes the
infected cells produce new viral DNA and proteins to assemble into new viruses.
The fact that cells double the amount of DNA in a cell prior to mitosis and then distribute the DNA equally to each
daughter cell provided some circumstantial evidence that DNA was the genetic material in eukaryotes.
Similar circumstantial evidence came from the observation that diploid sets of chromosomes have twice as much
DNA as the haploid sets in gametes of the same organism.
By 1947, Erwin Chargaff had developed a series of rules based on a survey of DNA composition in organisms.
He already knew that DNA was a polymer of nucleotides consisting of a nitrogenous base, deoxyribose, and a
phosphate group.
The bases could be adenine (A), thymine (T), guanine (G), or cytosine (C).
Chargaff noted that the DNA composition varies from species to species.
In any one species, the four bases are found in characteristic, but not necessarily equal, ratios.
He also found a peculiar regularity in the ratios of nucleotide bases that are known as Chargaffs rules.
In all organisms, the number of adenines was approximately equal to the number of thymines (%T = %A).
The number of guanines was approximately equal to the number of cytosines (%G = %C).
Human DNA is 30.9% adenine, 29.4% thymine, 19.9% guanine, and 19.8% cytosine.
The basis for these rules remained unexplained until the discovery of the double helix.
Watson and Crick discovered the double helix by building models to conform to X-ray data.
By the beginnings of the 1950s, the race was on to move from the structure of a single DNA strand to the three-
dimensional structure of DNA.
Among the scientists working on the problem were Linus Pauling in California and Maurice Wilkins and Rosalind
Franklin in London.
Maurice Wilkins and Rosalind Franklin used X-ray crystallography to study the structure of DNA.
In this technique, X-rays are diffracted as they passed through aligned fibers of purified DNA.
The diffraction pattern can be used to deduce the three-dimensional shape of molecules.
James Watson learned from their research that DNA was helical in shape, and he deduced the width of the helix and
the spacing of nitrogenous bases.
The width of the helix suggested that it was made up of two strands, contrary to a three-stranded model that Linus
Pauling had recently proposed.
Watson and his colleague Francis Crick began to work on a model of DNA with two strands, the double helix.
Using molecular models made of wire, they placed the sugar-phosphate chains on the outside and the nitrogenous
bases on the inside of the double helix.
This arrangement put the relatively hydrophobic nitrogenous bases in the molecules interior.
The sugar-phosphate chains of each strand are like the side ropes of a rope ladder.
Pairs of nitrogenous bases, one from each strand, form rungs.
The ladder forms a twist every ten bases.
The nitrogenous bases are paired in specific combinations: adenine with thymine and guanine with cytosine.
Pairing like nucleotides did not fit the uniform diameter indicated by the X-ray data.
A purine-purine pair is too wide, and a pyrimidine-pyrimidine pairing is too short.
Only a pyrimidine-purine pairing produces the 2-nm diameter indicated by the X-ray data.
In addition, Watson and Crick determined that chemical side groups of the nitrogenous bases would form hydrogen
bonds, connecting the two strands.
Based on details of their structure, adenine would form two hydrogen bonds only with thymine, and guanine would
form three hydrogen bonds only with cytosine.
This finding explained Chargaffs rules.
The base-pairing rules dictate the combinations of nitrogenous bases that form the rungs of DNA.
However, this does not restrict the sequence of nucleotides along each DNA strand.
The linear sequence of the four bases can be varied in countless ways.
Each gene has a unique order of nitrogenous bases.
In April 1953, Watson and Crick published a succinct, one-page paper in Nature reporting their double helix model
of DNA.
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 Cricks 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, they labeled the nucleotides of the old strands with a heavy isotope of nitrogen (15N), while
any new nucleotides were indicated by a lighter isotope (14N).
Replicated strands could be separated by density in a centrifuge.
Each modelthe semiconservative model, the conservative model, and the dispersive modelmade 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.
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.
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 1,0002,000 nucleotides long in E. coli and 100200 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 510 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.
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.
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 cells 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.

Lecture Outline for Campbell/Reece Biology, 7th Edition, Pearson Education, Inc. 16-1
Chapter 17 From Gene to Protein
Lecture Outline
Overview: The Flow of Genetic Information
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, Mendels 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 17.1 Genes specify proteins via transcription and translation
The study of metabolic defects provided evidence that genes specify proteins.
In 1909, Archibald Gerrod was the first to suggest that genes dictate phenotype through enzymes that catalyze
specific chemical reactions in the cell.
He suggested that the symptoms of an inherited disease reflect a persons inability to synthesize a particular enzyme.
He referred to such diseases as inborn errors of metabolism.
Gerrod speculated that alkaptonuria, a hereditary disease, was caused by the absence of an enzyme that breaks down
a specific substrate, alkapton.
Research conducted several decades later supported Gerrods hypothesis.
Progress in linking genes and enzymes rested on the growing understanding that cells synthesize and degrade most
organic molecules in a series of steps, a metabolic pathway.
In the 1930s, George Beadle and Boris Ephrussi speculated that each mutation affecting eye color in Drosophila
blocks pigment synthesis at a specific step by preventing production of the enzyme that catalyzes that step.
However, neither the chemical reactions nor the enzymes that catalyze them were known at the time.
Beadle and Edward Tatum were finally able to establish the link between genes and enzymes in their exploration of
the metabolism of a bread mold, Neurospora crassa.
They bombarded Neurospora with X-rays and screened the survivors for mutants that differed in their nutritional
needs.
Wild-type Neurospora can grow on a minimal medium of agar, inorganic salts, glucose, and the vitamin biotin.
Beadle and Tatum identified mutants that could not survive on minimal medium, because they were unable to
synthesize certain essential molecules from the minimal ingredients.
However, most of these nutritional mutants can survive on a complete growth medium that includes all 20 amino
acids and a few other nutrients.
One type of mutant required only the addition of arginine to the minimal growth medium.
Beadle and Tatum concluded that this mutant was defective somewhere in the biochemical pathway that normally
synthesizes arginine.
They identified three classes of arginine-deficient mutants, each apparently lacking a key enzyme at a different step
in the synthesis of arginine.
They demonstrated this by growing these mutant strains in media that provided different intermediate molecules.
Their results provided strong evidence for the one geneone enzyme hypothesis.
Later research refined the one geneone enzyme hypothesis.
First, not all proteins are enzymes.
Keratin, the structural protein of hair, and insulin, a hormone, both are proteins and gene products.
This tweaked the hypothesis to one geneone protein.
Later research demonstrated that many proteins are composed of several polypeptides, each of which has its own
gene.
Therefore, Beadle and Tatums idea has been restated as the one geneone polypeptide hypothesis.
Some genes code for RNA molecules that play important roles in cells although they are never translated into
protein.
Transcription and translation are the two main processes linking gene to protein.
Genes provide the instructions for making specific proteins.
The bridge between DNA and protein synthesis is the nucleic acid RNA.
RNA is chemically similar to DNA, except that it contains ribose as its sugar and substitutes the nitrogenous base
uracil for thymine.
An RNA molecule almost always consists of a single strand.
In DNA or RNA, the four nucleotide monomers act like the letters of the alphabet to communicate information.
The specific sequence of hundreds or thousands of nucleotides in each gene carries the information for the primary
structure of proteins, the linear order of the 20 possible amino acids.
To get from DNA, written in one chemical language, to protein, written in another, requires two major stages:
transcription and translation.
During transcription, a DNA strand provides a template for the synthesis of a complementary RNA strand.
Just as a DNA strand provides a template for the synthesis of each new complementary strand during DNA
replication, it provides a template for assembling a sequence of RNA nucleotides.
Transcription of many genes produces a messenger RNA (mRNA) molecule.
During translation, there is a change of language.
The site of translation is the ribosome, complex particles that facilitate the orderly assembly of amino acids into
polypeptide chains.
Why cant proteins be translated directly from DNA?
The use of an RNA intermediate provides protection for DNA and its genetic information.
Using an RNA intermediate allows more copies of a protein to be made simultaneously, since many RNA transcripts
can be made from one gene.
Also, each gene transcript can be translated repeatedly.
The basic mechanics of transcription and translation are similar in eukaryotes and prokaryotes.
Because bacteria lack nuclei, their DNA is not segregated from ribosomes and other protein-synthesizing equipment.
This allows the coupling of transcription and translation.
Ribosomes attach to the leading end of an mRNA molecule while transcription is still in progress.
In a eukaryotic cell, transcription occurs in the nucleus, and translation occurs at ribosomes in the cytoplasm.
The transcription of a protein-coding eukaryotic gene results in pre-mRNA.
The initial RNA transcript of any gene is called a primary transcript.
RNA processing yields the finished mRNA.
To summarize, genes program protein synthesis via genetic messages in the form of messenger RNA.
The molecular chain of command in a cell is DNA --> RNA --> protein.
In the genetic code, nucleotide triplets specify amino acids.
If the genetic code consisted of a single nucleotide or even pairs of nucleotides per amino acid, there would not be
enough combinations (4 and 16, respectively) to code for all 20 amino acids.
Triplets of nucleotide bases are the smallest units of uniform length that can code for all the amino acids.
With a triplet code, three consecutive bases specify an amino acid, creating 43 (64) possible code words.
The genetic instructions for a polypeptide chain are written in DNA as a series of nonoverlapping three-nucleotide
words.
During transcription, one DNA strand, the template strand, provides a template for ordering the sequence of
nucleotides in an RNA transcript.
A given DNA strand can be the template strand for some genes along a DNA molecule, while for other genes in
other regions, the complementary strand may function as the template.
The complementary RNA molecule is synthesized according to base-pairing rules, except that uracil is the
complementary base to adenine.
Like a new strand of DNA, the RNA molecule is synthesized in an antiparallel direction to the template strand of
DNA.
The mRNA base triplets are called codons, and they are written in the 5 --> 3 direction.
During translation, the sequence of codons along an mRNA molecule is translated into a sequence of amino acids
making up the polypeptide chain.
During translation, the codons are read in the 5 --> 3 direction along the mRNA.
Each codon specifies which one of the 20 amino acids will be incorporated at the corresponding position along a
polypeptide.
Because codons are base triplets, the number of nucleotides making up a genetic message must be three times the
number of amino acids making up the protein product.
It takes at least 300 nucleotides to code for a polypeptide that is 100 amino acids long.
The task of matching each codon to its amino acid counterpart began in the early 1960s.
Marshall Nirenberg determined the first match: UUU coded for the amino acid phenylalanine.
He created an artificial mRNA molecule entirely of uracil and added it to a test tube mixture of amino acids,
ribosomes, and other components for protein synthesis.
This poly-U translated into a polypeptide containing a single amino acid, phenylalanine, in a long chain.
AAA, GGG, and CCC were solved in the same way.
Other more elaborate techniques were required to decode mixed triplets such as AUA and CGA.
By the mid-1960s the entire code was deciphered.
Sixty-one of 64 triplets code for amino acids.
The codon AUG not only codes for the amino acid methionine, but also indicates the start of translation.
Three codons do not indicate amino acids but are stop signals marking the termination of translation.
There is redundancy in the genetic code but no ambiguity.
Several codons may specify the same amino acid, but no codon specifies more than one amino acid.
The redundancy in the code is not random. In many cases, codons that are synonyms for a particular amino acid
differ only in the third base of the triplet.
To extract the message from the genetic code requires specifying the correct starting point.
This establishes the reading frame; subsequent codons are read in groups of three nucleotides.
The cells protein-synthesizing machinery reads the message as a series of nonoverlapping three-letter words.
In summary, genetic information is encoded as a sequence of nonoverlapping base triplets, or codons, each of which
is translated into a specific amino acid during protein synthesis.
The genetic code must have evolved very early in the history of life.
The genetic code is nearly universal, shared by organisms from the simplest bacteria to the most complex plants and
animals.
In laboratory experiments, genes can be transcribed and translated after they are transplanted from one species to
another.
This has permitted bacteria to be programmed to synthesize certain human proteins after insertion of the appropriate
human genes.
Such applications are exciting developments in biotechnology.
Exceptions to the universality of the genetic code exist in certain unicellular eukaryotes and in the organelle genes of
some species.
Some prokaryotes can translate stop codons into one of two amino acids not found in most organisms.
The evolutionary significance of the near universality of the genetic code is clear.
A language shared by all living things arose very early in the history of lifeearly enough to be present in the
common ancestors of all modern organisms.
A shared genetic vocabulary is a reminder of the kinship that bonds all life on Earth.
Concept 17.2 Transcription is the DNA-directed synthesis of RNA: a closer look
Messenger RNA, the carrier of information from DNA to the cells protein-synthesizing machinery, is transcribed
from the template strand of a gene.
RNA polymerase separates the DNA strands at the appropriate point and bonds the RNA nucleotides as they base-
pair along the DNA template.
Like DNA polymerases, RNA polymerases can only assemble a polynucleotide in its 5 --> 3 direction.
Unlike DNA polymerases, RNA polymerases are able to start a chain from scratch; they dont need a primer.
Specific sequences of nucleotides along the DNA mark where gene transcription begins and ends.
RNA polymerase attaches and initiates transcription at the promoter.
In prokaryotes, the sequence that signals the end of transcription is called the terminator.
Molecular biologists refer to the direction of transcription as downstream and the other direction as upstream.
The stretch of DNA that is transcribed into an RNA molecule is called a transcription unit.
Bacteria have a single type of RNA polymerase that synthesizes all RNA molecules.
In contrast, eukaryotes have three RNA polymerases (I, II, and III) in their nuclei.
RNA polymerase II is used for mRNA synthesis.
Transcription can be separated into three stages: initiation, elongation, and termination of the RNA chain.
The presence of a promoter sequence determines which strand of the DNA helix is the template.
Within the promoter is the starting point for the transcription of a gene.
The promoter also includes a binding site for RNA polymerase several dozen nucleotides upstream of the start
point.
In prokaryotes, RNA polymerase can recognize and bind directly to the promoter region.
In eukaryotes, proteins called transcription factors mediate the binding of RNA polymerase and the initiation of
transcription.
Only after certain transcription factors are attached to the promoter does RNA polymerase II bind to it.
The completed assembly of transcription factors and RNA polymerase II bound to a promoter is called a
transcription initiation complex.
A crucial promoter DNA sequence is called a TATA box.
RNA polymerase then starts transcription.
As RNA polymerase moves along the DNA, it untwists the double helix, 10 to 20 bases at time.
The enzyme adds nucleotides to the 3 end of the growing strand.
Behind the point of RNA synthesis, the double helix re-forms and the RNA molecule peels away.
Transcription progresses at a rate of 60 nucleotides per second in eukaryotes.
A single gene can be transcribed simultaneously by several RNA polymerases at a time.
A growing strand of RNA trails off from each polymerase.
The length of each new strand reflects how far along the template the enzyme has traveled from the start point.
The congregation of many polymerase molecules simultaneously transcribing a single gene increases the amount of
mRNA transcribed from it.
This helps the cell make the encoded protein in large amounts.
Transcription proceeds until after the RNA polymerase transcribes a terminator sequence in the DNA.
In prokaryotes, RNA polymerase stops transcription right at the end of the terminator.
Both the RNA and DNA are then released.
In eukaryotes, the pre-mRNA is cleaved from the growing RNA chain while RNA polymerase II continues to
transcribe the DNA.
Specifically, the polymerase transcribes a DNA sequence called the polyadenylation signal sequence that codes for a
polyadenylation sequence (AAUAAA) in the pre-mRNA.
At a point about 10 to 35 nucleotides past this sequence, the pre-mRNA is cut from the enzyme.
The polymerase continues transcribing for hundreds of nucleotides.
Transcription is terminated when the polymerase eventually falls off the DNA.
Concept 17.3 Eukaryotic cells modify RNA after transcription
Enzymes in the eukaryotic nucleus modify pre-mRNA before the genetic messages are dispatched to the cytoplasm.
During RNA processing, both ends of the primary transcript are usually altered.
Certain interior parts of the molecule are cut out and the remaining parts spliced together.
At the 5 end of the pre-mRNA molecule, a modified form of guanine is added, the 5 cap.
At the 3 end, an enzyme adds 50 to 250 adenine nucleotides, the poly-A tail.
These modifications share several important functions.
They seem to facilitate the export of mRNA from the nucleus.
They help protect mRNA from hydrolytic enzymes.
They help the ribosomes attach to the 5 end of the mRNA.
The most remarkable stage of RNA processing occurs during the removal of a large portion of the RNA molecule in
a cut-and-paste job of RNA splicing.
Most eukaryotic genes and their RNA transcripts have long noncoding stretches of nucleotides.
Noncoding segments of nucleotides called intervening regions, or introns, lie between coding regions.
The final mRNA transcript includes coding regions, exons, which are translated into amino acid sequences, plus the
leader and trailer sequences.
RNA splicing removes introns and joins exons to create an mRNA molecule with a continuous coding sequence.
This splicing is accomplished by a spliceosome.
Spliceosomes consist of a variety of proteins and several small nuclear ribonucleoproteins (snRNPs) that recognize
the splice sites.
snRNPs are located in the cell nucleus and are composed of RNA and protein molecules.
Each snRNP has several protein molecules and a small nuclear RNA molecule (snRNA).
Each snRNA is about 150 nucleotides long.
The spliceosome interacts with certain sites along an intron, releasing the introns and joining together the two exons
that flanked the introns.
snRNAs appear to play a major role in catalytic processes, as well as spliceosome assembly and splice site
recognition.
The idea of a catalytic role for snRNA arose from the discovery of ribozymes, RNA molecules that function as
enzymes.
In some organisms, splicing occurs without proteins or additional RNA molecules.
The intron RNA functions as a ribozyme and catalyzes its own excision.
For example, in the protozoan Tetrahymena, self-splicing occurs in the production of ribosomal RNA (rRNA), a
component of the organisms ribosomes.
The pre-rRNA actually removes its own introns.
The discovery of ribozymes rendered obsolete the statement, All biological catalysts are proteins.
The fact that RNA is single-stranded plays an important role in allowing certain RNA molecules to function as
ribozymes.
A region of the RNA molecule may base-pair with a complementary region elsewhere in the same molecule, thus
giving the RNA a specific 3-D structure that is key to its ability to catalyze reactions.
Introns and RNA splicing appear to have several functions.
Some introns play a regulatory role in the cell. These introns contain sequences that control gene activity in some
way.
Splicing itself may regulate the passage of mRNA from the nucleus to the cytoplasm.
One clear benefit of split genes is to enable one gene to encode for more than one polypeptide.
Alternative RNA splicing gives rise to two or more different polypeptides, depending on which segments are treated
as exons.
Sex differences in fruit flies may be due to differences in splicing RNA transcribed from certain genes.
Early results of the Human Genome Project indicate that this phenomenon may be common in humans, and may
explain why we have a relatively small number of genes.
Proteins often have a modular architecture with discrete structural and functional regions called domains.
The presence of introns in a gene may facilitate the evolution of new and potentially useful proteins as a result of a
process known as exon shuffling.
In many cases, different exons code for different domains of a protein.
The presence of introns increases the probability of potentially beneficial crossing over between genes.
Introns increase the opportunity for recombination between two alleles of a gene.
This raises the probability that a crossover will switch one version of an exon for another version found on the
homologous chromosome.
There may also be occasional mixing and matching of exons between completely different genes.
Either way, exon shuffling can lead to new proteins through novel combinations of functions.
Concept 17.4 Translation is the RNA-directed synthesis of a polypeptide: a closer look
In the process of translation, a cell interprets a series of codons along an mRNA molecule and builds a polypeptide.
The interpreter is transfer RNA (tRNA), which transfers amino acids from the cytoplasmic pool to a ribosome.
A cell has all 20 amino acids available in its cytoplasm, either by synthesizing them from scratch or by taking them
up from the surrounding solution.
The ribosome adds each amino acid carried by tRNA to the growing end of the polypeptide chain.
During translation, each type of tRNA links an mRNA codon with the appropriate amino acid.
Each tRNA arriving at the ribosome carries a specific amino acid at one end and has a specific nucleotide triplet, an
anticodon, at the other.
The anticodon base-pairs with a complementary codon on mRNA.
If the codon on mRNA is UUU, a tRNA with an AAA anticodon and carrying phenylalanine will bind to it.
Codon by codon, tRNAs deposit amino acids in the prescribed order, and the ribosome joins them into a polypeptide
chain.
The tRNA molecule is a translator, because it can read a nucleic acid word (the mRNA codon) and translate it to a
protein word (the amino acid).
Like other types of RNA, tRNA molecules are transcribed from DNA templates in the nucleus.
Once it reaches the cytoplasm, each tRNA is used repeatedly, picking up its designated amino acid in the cytosol,
depositing the amino acid at the ribosome, and returning to the cytosol to pick up another copy of that amino acid.
A tRNA molecule consists of a strand of about 80 nucleotides that folds back on itself to form a three-dimensional
structure.
It includes a loop containing the anticodon and an attachment site at the 3 end for an amino acid.
If each anticodon had to be a perfect match to each codon, we would expect to find 61 types of tRNA, but the actual
number is about 45.
The anticodons of some tRNAs recognize more than one codon.
This is possible because the rules for base pairing between the third base of the codon and anticodon are relaxed
(called wobble).
At the wobble position, U on the anticodon can bind with A or G in the third position of a codon.
Wobble explains why the synonymous codons for a given amino acid can differ in their third base, but not usually in
their other bases.
Each amino acid is joined to the correct tRNA by aminoacyl-tRNA synthetase.
The 20 different synthetases match the 20 different amino acids.
Each has active sites for only a specific tRNA-and-amino-acid combination.
The synthetase catalyzes a covalent bond between them in a process driven by ATP hydrolysis.
The result is an aminoacyl-tRNA or activated amino acid.
Ribosomes facilitate the specific coupling of the tRNA anticodons with mRNA codons during protein synthesis.
Each ribosome is made up of a large and a small subunit.
The subunits are composed of proteins and ribosomal RNA (rRNA), the most abundant RNA in the cell.
In eukaryotes, the subunits are made in the nucleolus.
After rRNA genes are transcribed to rRNA in the nucleus, the rRNA and proteins are assembled to form the subunits
with proteins from the cytoplasm.
The subunits exit the nucleus via nuclear pores.
The large and small subunits join to form a functional ribosome only when they attach to an mRNA molecule.
While very similar in structure and function, prokaryotic and eukaryotic ribosomes have enough differences that
certain antibiotic drugs (like tetracycline) can paralyze prokaryotic ribosomes without inhibiting eukaryotic
ribosomes.
Each ribosome has a binding site for mRNA and three binding sites for tRNA molecules.
The P site holds the tRNA carrying the growing polypeptide chain.
The A site carries the tRNA with the next amino acid to be added to the chain.
Discharged tRNAs leave the ribosome at the E (exit) site.
The ribosome holds the tRNA and mRNA in close proximity and positions the new amino acid for addition to the
carboxyl end of the growing polypeptide.
It then catalyzes the formation of the peptide bond.
As the polypeptide becomes longer, it passes through an exit tunnel in the ribosomes large unit and is released to
the cytosol.
Recent advances in our understanding of the structure of the ribosome strongly support the hypothesis that rRNA,
not protein, carries out the ribosomes functions.
RNA is the main constituent at the interphase between the two subunits and of the A and P sites.
It is the catalyst for peptide bond formation.
A ribosome can be regarded as one colossal ribozyme.
Translation can be divided into three stages: initiation, elongation, and termination.
All three phases require protein factors that aid in the translation process.
Both initiation and chain elongation require energy provided by the hydrolysis of GTP.
Initiation brings together mRNA, a tRNA with the first amino acid, and the two ribosomal subunits.
First, a small ribosomal subunit binds with mRNA and a special initiator tRNA, which carries methionine and
attaches to the start codon.
The small subunit then moves downstream along the mRNA until it reaches the start codon, AUG, which signals the
start of translation.
This establishes the reading frame for the mRNA.
The initiator tRNA, already associated with the complex, then hydrogen-bonds with the start codon.
Proteins called initiation factors bring in the large subunit so that the initiator tRNA occupies the P site.
Elongation involves the participation of several protein elongation factors, and consists of a series of three-step
cycles as each amino acid is added to the proceeding one.
During codon recognition, an elongation factor assists hydrogen bonding between the mRNA codon under the A site
with the corresponding anticodon of tRNA carrying the appropriate amino acid.
This step requires the hydrolysis of two GTP.
During peptide bond formation, an rRNA molecule catalyzes the formation of a peptide bond between the
polypeptide in the P site with the new amino acid in the A site.
This step separates the tRNA at the P site from the growing polypeptide chain and transfers the chain, now one
amino acid longer, to the tRNA at the A site.
During translocation, the ribosome moves the tRNA with the attached polypeptide from the A site to the P site.
Because the anticodon remains bonded to the mRNA codon, the mRNA moves along with it.
The next codon is now available at the A site.
The tRNA that had been in the P site is moved to the E site and then leaves the ribosome.
Translocation is fueled by the hydrolysis of GTP.
Effectively, translocation ensures that the mRNA is read 5 --> 3 codon by codon.
The three steps of elongation continue to add amino acids codon by codon until the polypeptide chain is completed.
Termination occurs when one of the three stop codons reaches the A site.
A release factor binds to the stop codon and hydrolyzes the bond between the polypeptide and its tRNA in the P site.
This frees the polypeptide, and the translation complex disassembles.
Typically a single mRNA is used to make many copies of a polypeptide simultaneously.
Multiple ribosomes, polyribosomes, may trail along the same mRNA.
Polyribosomes can be found in prokaryotic and eukaryotic cells.
A ribosome requires less than a minute to translate an average-sized mRNA into a polypeptide.
During and after synthesis, a polypeptide coils and folds to its three-dimensional shape spontaneously.
The primary structure, the order of amino acids, determines the secondary and tertiary structure.
Chaperone proteins may aid correct folding.
In addition, proteins may require posttranslational modifications before doing their particular job.
This may require additions such as sugars, lipids, or phosphate groups to amino acids.
Enzymes may remove some amino acids or cleave whole polypeptide chains.
Two or more polypeptides may join to form a protein.
Signal peptides target some eukaryotic polypeptides to specific destinations in the cell.
Two populations of ribosomes, free and bound, are active participants in protein synthesis.
Free ribosomes are suspended in the cytosol and synthesize proteins that reside in the cytosol.
Bound ribosomes are attached to the cytosolic side of the endoplasmic reticulum.
They synthesize proteins of the endomembrane system as well as proteins secreted from the cell.
While bound and free ribosomes are identical in structure, their location depends on the type of protein that they are
synthesizing.
Translation in all ribosomes begins in the cytosol, but a polypeptide destined for the endomembrane system or for
export has a specific signal peptide region at or near the leading end.
This consists of a sequence of about 20 amino acids.
A signal recognition particle (SRP) binds to the signal peptide and attaches it and its ribosome to a receptor protein
in the ER membrane.
The SRP consists of a protein-RNA complex.
After binding, the SRP leaves and protein synthesis resumes with the growing polypeptide snaking across the
membrane into the cisternal space via a protein pore.
An enzyme usually cleaves the signal polypeptide.
Secretory proteins are released entirely into the cisternal space, but membrane proteins remain partially embedded in
the ER membrane.
Other kinds of signal peptides are used to target polypeptides to mitochondria, chloroplasts, the nucleus, and other
organelles that are not part of the endomembrane system.
In these cases, translation is completed in the cytosol before the polypeptide is imported into the organelle.
While the mechanisms of translocation vary, each of these polypeptides has a ZIP code that ensures its delivery to
the correct cellular location.
Prokaryotes also employ signal sequences to target proteins for secretion.
Concept 17.5 RNA plays multiple roles in the cell: a review
The cellular machinery of protein synthesis and ER targeting is dominated by various kinds of RNA.
In addition to mRNA, these include tRNA; rRNA; and in eukaryotes, snRNA and SRP RNA.
A type of RNA called small nucleolar RNA (snoRNA) aids in processing pre-rRNA transcripts in the nucleolus, a
process necessary for ribosome formation.
Recent research has also revealed the presence of small, single-stranded and double-stranded RNA molecules that
play important roles in regulating which genes get expressed.
These types of RNA include small interfering RNA (siRNA) and microRNA (miRNA).
The diverse functions of RNA are based, in part, on its ability to form hydrogen bonds with other nucleic acid
molecules (DNA or RNA).
It can also assume a specific three-dimensional shape by forming hydrogen bonds between bases in different parts of
its polynucleotide chain.
DNA may be the genetic material of all living cells today, but RNA is much more versatile.
The diverse functions of RNA range from structural to informational to catalytic.
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.
For example, sickle-cell disease is caused by a mutation of a single base pair in the gene that codes for one of the
polypeptides of hemoglobin.
A change in a single nucleotide from T to A in the DNA template leads to an abnormal protein.
A point mutation that results in the replacement of a pair of complementary nucleotides with another nucleotide pair
is called a base-pair substitution.
Some base-pair substitutions have little or no impact on protein function.
In silent mutations, altered nucleotides still code for the same amino acids because of redundancy in the genetic
code.
Other changes lead to switches from one amino acid to another with similar properties.
Still other mutations may occur in a region where the exact amino acid sequence is not essential for function.
Other base-pair substitutions cause a readily detectable change in a protein.
These are usually detrimental but can occasionally lead to an improved protein or one with novel capabilities.
Changes in amino acids at crucial sites, especially active sites, are likely to impact function.
Missense mutations are those that still code for an amino acid but a different one.
Nonsense mutations change an amino acid codon into a stop codon, nearly always leading to a nonfunctional
protein.
Insertions and deletions are additions or losses of nucleotide pairs in a gene.
These have a disastrous effect on the resulting protein more often than substitutions do.
Unless insertion or deletion mutations occur in multiples of three, they cause a frameshift mutation.
All the nucleotides downstream of the deletion or insertion will be improperly grouped into codons.
The result will be extensive missense, ending sooner or later in nonsensepremature termination.
Mutations can occur in a number of ways.
Errors can occur during DNA replication, DNA repair, or DNA recombination.
These can lead to base-pair substitutions, insertions, or deletions, as well as mutations affecting longer stretches of
DNA.
These are called spontaneous mutations.
Rough estimates suggest that about 1 nucleotide in every 1010 is altered and inherited by daughter cells.
Mutagens are chemical or physical agents that interact with DNA to cause mutations.
Physical agents include high-energy radiation like X-rays and ultraviolet light.
Chemical mutagens fall into several categories.
Some chemicals are base analogues that may be substituted into DNA, but they pair incorrectly during DNA
replication.
Other mutagens interfere with DNA replication by inserting into DNA and distorting the double helix.
Still others cause chemical changes in bases that change their pairing properties.
Researchers have developed various methods to test the mutagenic activity of different chemicals.
These tests are often used as a preliminary screen of chemicals to identify those that may cause cancer.
This makes sense because most carcinogens are mutagenic and most mutagens are carcinogenic.
What is a gene? We revisit the question.
The Mendelian concept of a gene views it as a discrete unit of inheritance that affects phenotype.
Morgan and his colleagues assigned genes to specific loci on chromosomes.
We can also view a gene as a specific nucleotide sequence along a region of a DNA molecule.
We can define a gene functionally as a DNA sequence that codes for a specific polypeptide chain.
All these definitions are useful in certain contexts.
Even the one geneone polypeptide definition must be refined and applied selectively.
Most eukaryotic genes contain large introns that have no corresponding segments in polypeptides.
Promoters and other regulatory regions of DNA are not transcribed either, but they must be present for transcription
to occur.
Our molecular definition must also include the various types of RNA that are not translated into polypeptides, such
as rRNA, tRNA, and other RNAs.
This is our definition of a gene: A gene is a region of DNA whose final product is either a polypeptide or an RNA
molecule.

Lecture Outline for Campbell/Reece Biology, 7th Edition, Pearson Education, Inc. 17-1

S-ar putea să vă placă și