You are on page 1of 26

Cell Cycle

• In eukaryotes, the cell cycle is a complex process and consists of four discrete phases.
• Cell growth is usually a continuous process
• DNA synthesis occurs only during one phase of the cell cycle
• The replicated chromosomes are then distributed to daughter nuclei by a complex series of events before cell
division.
• Progression between different stages of the cell cycle is a highly regulated and controlled mechanism.
• Regulatory apparatus coordinates with different events of the cell cycle and also links the cell cycle with
extracellular signals that control cell proliferation.
Phases of the cell cycle
• A typical Eukaryotic Cell approximately every 24 hours (though the time varies depending upon the type of cell).
• As viewed in the microscope, the cell cycle is divided into two basic parts:
• mitosis and interphase.
• Mitosis (nuclear division) is the most dramatic stage of the cell cycle, corresponding to the separation of
daughter chromosomes and usually ending with cell division (cytokinesis).
• Mitosis and cytokinesis last only about an hour.
• approximately 95% of the cell cycle is spent in interphase (the period between mitoses).
• During interphase, the chromosomes are decondensed and distributed throughout the nucleus, so the nucleus
appears morphologically uniform.
• At molecular details, interphase is the time during which both cell growth and DNA replication occur in an
orderly manner in preparation for cell division.
• The cell grows at a steady rate throughout interphase, with most dividing cells doubling in size between one
mitosis and the next.
• In contrast, DNA is synthesized during only a portion of interphase. The timing of DNA synthesis thus divides
the cycle of eukaryotic cells into four discrete phases.
• The M phase is mitosis, followed by cytokinesis. This phase is followed by the G1 phase (gap1), which
corresponds to the interval between mitosis and initiation of DNA replication. During G1, the cell is
metabolically active and continuously grows but does not replicate its DNA.
• S phase (synthesis), during which DNA replication takes place. The completion of DNA synthesis is followed by
the G2 phase (gap 2), during which proteins are synthesized in preparation for mitosis.
• For a typical proliferating human cell with a total cycle
time of 24 hours, the G1 about 11 hours, S phase about 8
hours, G2 about 4 hours, and M about 1 hour.
• Budding yeasts can progress through all four stages of the
cell cycle in only about 90 minutes.
• Even much shorter cycle occur in early stages of
embryonic cell division (approx. 30 mins).
• Some mature cells cease cell division (neurons)
• many other cells divide only occasionally, as needed to
replace cells that have been lost because of injury or cell
death. Examples are skin fibroblasts, liver cells.
• These cells exit G1 to enter a quiescent stage of the cycle called G0.
• Some cells that are arrested in G0 can re-enter the cell cycle when called on to do so by appropriate
extracellular signals.
• Other types of cells including most differentiated cells in adult animals, permanently withdraw from the cell
cycle and are incapable of resuming proliferation.
Regulation of the cell
• The progression of cell cycle is regulated by extracellular and
intracellular signals
• Different cellular processes, such as cell growth, DNA replication, and
mitosis, all must be coordinated during cell cycle progression. This is
accomplished by a series of control points that regulate progression through
various phases of the cell cycle.
A major cell cycle regulatory point in many types of cells controls
progression from G1 to S. This regulatory point was first defined by studies
of budding yeast (Saccharomyces cerevisiae), where it is known as START
Once cells have passed START, they are committed to entering S phase
and undergoing one cell division cycle. However, passage through START
is a highly regulated event in the yeast cell cycle, controlled by external
signals such as the availability of nutrients, as well as by cell size.
• Proliferation of most cells is regulated primarily in G1
• Some cell cycles are instead controlled principally in G2.
• An example is the cell cycle of the fission yeast Schizosaccharomyces pombe.
• Here the cycle is regulated primarily by control of the transition from G2 to
M, which is the principal point at which cell size and nutrient availability are
monitored.
• In animals, the primary example of cell cycle control in G2 is provided by
oocytes. Vertebrate oocytes can remain arrested in G2 for long periods of time
(several decades in humans) until their progression to M phase is triggered by
hormonal stimulation.
• Extracellular signals can thus control cell proliferation by regulating
progression from the G2 to M as well as the G1 to S phases of the cell cycle.
• Coordination between different phases of the cell cycle depends on a series of
cell cycle checkpoints.
• These prevent entry into the next phase of the cell cycle until the events of the
preceding phase have been completed.
• Several cell cycle checkpoints, called DNA damage checkpoints, function to
ensure that damaged DNA is not replicated and passed on to daughter cells.
• These checkpoints sense damaged or incompletely replicated DNA and
coordinate further cell cycle progression with the completion of DNA
replication or repair.
a cell in an ovary which may undergo meiotic division to
Protein
form kinases and cell cycle regulation:
an ovum. meiosis
• Oocytes are arrested in the G2 phase of the cell cycle until
hormonal stimulation (progesterone) triggers their entry into
the M phase of meiosis.
• In 1971, two independent teams of researchers (Yoshio Masui
and Clement Markert, as well as Dennis Smith and Robert
Ecker) found that oocytes arrested in G2 could be induced to
enter M phase by microinjection of cytoplasm from oocytes
that had been hormonally stimulated.
• It thus appeared that a cytoplasmic factor present in hormone-
treated oocytes was sufficient to trigger the transition from G2
to M in oocytes that had not been exposed to hormone.
• Because the entry of oocytes into meiosis is frequently
referred to as oocyte maturation, this cytoplasmic factor was
called maturation promoting factor (MPF). Further studies
showed, however, that the activity of MPF is not restricted to
the entry of oocytes into meiosis.
• MPF is also present in somatic cells, where it induces entry
into M phase of the mitotic cycle. Rather than being specific
to oocytes, MPF thus appeared to act as a general regulator of
the transition from G2 to M.
• Lee Hartwell and his colleagues in the early 1970s studied the budding yeast S. cerevisiae.
• They identified temperature-sensitive mutants that were defective in cell cycle progression.
• The key characteristic of these mutants (called cdc for cell division cycle mutants) was that they underwent
growth arrest at specific points in the cell cycle.
• For example, a particularly important mutant designated cdc28 caused the cell cycle to arrest at START,
indicating that the Cdc28 protein is required for passage through this critical regulatory point in G1.
• A similar collection of cell cycle mutants was isolated in the fission yeast S. pombe by Paul Nurse and his
collaborators.
• These mutants included cdc2, which arrests the S. pombe cell cycle both in G1 and at the G2 to M transition
(the major regulatory point in fission yeast).
• Comparative analysis showed that S. cerevisiae cdc28 and S. pombe cdc2 are functionally homologous genes,
which are required for passage through START as well as for entry into mitosis in both species of yeasts.
• Further studies demonstrated that cdc2 and cdc28 encode a protein kinase-the first indication of the prominent
role of protein phosphorylation in regulating the cell cycle.
• In addition, related genes were identified in other eukaryotes, including humans.
• The protein kinase encoded by the yeast cdc2 and cdc28 genes has since been shown to be a conserved cell
cycle regulator in all eukaryotes, which is known as Cdk1.
• studies of protein synthesis in early sea urchin embryos demonstrated other proteins involved.
• Following fertilization, these embryos go through a series of rapid cell divisions.
• Intriguingly, studies with protein synthesis inhibitors had revealed that entry into M phase of these
embryonic cell cycles requires new protein synthesis.
• In 1983, Tim Hunt and his colleagues identified two proteins that display a periodic pattern of
accumulation and degradation in sea urchin and clam embryos.
• These proteins accumulate throughout interphase and are then rapidly degraded toward the end of each
mitosis. Hunt called these proteins cyclins (the two proteins were designated cyclin A and cyclin B) and
suggested that they might function to induce mitosis, with their periodic accumulation and destruction
controlling entry and exit from M phase.
• Direct support for such a role of cyclins was provided in 1986, when Joan Ruderman and her colleagues
showed that microinjection of cyclin A into frog oocytes is sufficient to trigger the G2 to M transition.
In 1988 MPF was purified from frog eggs in the laboratory of
James Maller.
Molecular characterization of MPF in several laboratories then
showed that this conserved regulator of the cell cycle is composed
of two key subunits: Cdk1 and cyclin B.
Cyclin B is a regulatory subunit required for catalytic activity of
the Cdk1 protein kinase, consistent with the notion that MPF
activity is controlled by the periodic accumulation and degradation
of cyclin B during cell cycle progression.
Families of cyclins and cyclin-dependent kinases
• CDK1 controls entry at START as well as entry into mitosis with the help of different Cyclins.
• The G2 to M transition is driven by Cdk1 in association with the mitotic B-type cyclins (Clb1, Clb2, Clb3,
and Clb4).
• Passage through START, on the other hand, is controlled by Cdk1 in association with a distinct class of
cyclins called G1 cyclins (or Cln’s).
• Cdk1 then associates with different B-type cyclins (Clb5 and Clb6), which are required for progression
through S phase.
• These associations of Cdk1 with distinct B-type and G1 cyclins direct Cdk1 to phosphorylate different
substrate proteins, as required for progression through specific phases of the cell cycle.
• The cell cycles of higher eukaryotes are controlled not only by multiple cyclins but also by multiple Cdk1-
related protein kinases, which are known as Cdk’s for cyclin-dependent kinases.
• These multiple members of the Cdk family associate with specific cyclins to drive progression through the
different stages of the cell cycle.
• For example, progression from G1 to S is regulated principally by Cdk4, Cdk6, and Cdk2, in association
with cyclins D and E. Complexes of Cdk4 and Cdk6 with the D-type cyclins (D1, D2, and D3) play a
critical role in progression through the restriction point in G1.
• The E-type cyclins (E1 and E2) are then expressed, and Cdk2/ cyclin E complexes are required for the G1
to S transition and initiation of DNA synthesis. Complexes of Cdk2 with A-type cyclins (A1 and A2) are
activated later in S phase and function through G2. The initiation of mitosis and passage through M phase
are then driven by Cdk1 in complexes with A- and B-type cyclins (B1, B2, and B3)
Mitosis
Mitosis is a type of cell division that results in two daughter cells each having the same number and kind of
chromosomes as the parent nucleus, typical of ordinary tissue growth.

Stages of mitosis

• Mitosis is conventionally divided into four primary stages—prophase, metaphase, anaphase, and telophase
• The beginning of prophase is marked by the appearance of condensed chromosomes, each of which consists of
two sister chromatids (the daughter DNA molecules produced in S phase).
• These newly replicated DNA molecules remain intertwined throughout S and G2, becoming untangled during
the process of chromatin condensation.
• The condensed sister chromatids are then held together at the centromere, which is a chromosomal region to
which proteins bind to form the kinetochore—the site of eventual attachment of the spindle microtubules.
• Also, cytoplasmic changes lead to the development of the mitotic spindle initiate during prophase.
• The centrosomes (which had duplicated during interphase) separate and move to opposite sides of the nucleus.
• There they serve as the two poles of the mitotic spindle, which begins to form during late prophase.
• In higher eukaryotes, the end of prophase corresponds to
the breakdown of the nuclear envelope. The cell then
enters prometaphase—a transition period between
prophase and metaphase. During prometaphase the
microtubules of the mitotic spindle attach to the
kinetochores of condensed chromosomes.
• The kinetochores of sister chromatids are oriented on
opposite sides of the chromosome, so they attach to
microtubules emanating from opposite poles of the
spindle. The chromosomes shuffle back and forth until
they eventually align on the metaphase plate in the center
of the spindle. At this stage, the cell has reached
metaphase
• The transition from metaphase to anaphase is triggered
by breakage of the link between sister chromatids, which
then separate and move to opposite poles of the spindle.
Mitosis ends with telophase, during which nuclei reform
and the chromosomes decondense. Cytokinesis usually
begins during late anaphase and is almost complete by
the end of telophase, resulting in the formation of two
interphase daughter cells.
Meiosis

Meiosis is a type of cell division that results in four daughter cells each with half the number of
chromosomes of the parent cell, as in the production of gametes and plant spores.
• In meiosis, homologous chromosomes form pairs; that is, they synapse (or undergo) synapsis.
• Each synapsed structure, initially called a bivalent, eventually gives rise to a tetrad consisting of four
chromatids.
• The presence of four chromatids demonstrates that both homologs (making up the bivalent) have, in fact,
duplicated.
• To achieve haploidy, two divisions are necessary.
• The first division occurs in meiosis I and is described as a reductional division (because the number of
centromeres, each representing one chromosome, is reduced by one-half).
• Components of each tetrad—representing the two homologs— separate, yielding two dyads.
• Each dyad is composed of two sister chromatids joined at a common centromere.
• The second division occurs during meiosis II and is described as an equational division (because the number
of centromeres remains equal).
• Here each dyad splits into two monads of one chromosome each. Thus, the two divisions potentially produce
four haploid cells.
Meiosis Reduces the Parental Chromosome Number
• A second type of eukaryotic cell division is specialized to produce cells that have half the number of
chromosomes as the parental cell.
• These cells go on to form egg and sperm cells involved in mating. This is accomplished by following DNA
replication with two rounds of chromosome segregation.
• Like the mitotic cell cycle, the meiotic cell cycle includes a G1, S, and an elongated G2 phase.
• During the meiotic S phase, each chromosome is replicated, and the daughter chromatids remain associated as
in the mitotic S phase.
• Cells that enter meiosis must be diploid and thus contain two copies of each chromosome before DNA
replication, one derived from each parent.
• After DNA replication, these related sister-chromatid pairs, called homologs, pair with each other and
recombine. Recombination between the homologs creates a physical linkage between the two homologs that
is required to connect the two related sister-chromatid pairs during chromosome segregation.
• The most significant difference between the mitotic and meiotic cell cycles occurs during chromosome
segregation. Unlike mitosis, during which a single round of chromosome segregation follows DNA
replication, chromosomes participating in meiosis go through two rounds of segregation known as meiosis I
and II.
The First Meiotic Division: Prophase I
• As in mitosis, chromatin present in interphase thickens and coils into visible chromosomes.
• Also, as in mitosis, each chromosome is a double structure, held together by the molecular complex
called cohesin.
• Unlike mitosis, members of each homologous pair of chromosomes pair up, undergoing synapsis.
• Crossing over occurs between chromatids of synapsed homologs.
• Because of the complexity of these genetic events, this stage of meiosis is divided into five substages:
leptonema, zygonema, pachynema, diplonema,* and diakinesis.
• Leptonema During the leptotene stage, the interphase chromatin material begins to condense, and
the chromosomes, though still extended, become visible. Along each chromosome are chromomeres,
localized condensations that resemble beads on a string. Evidence suggests that a process called
homology search, which precedes and is essential to the initial pairing of homologs, begins during
leptonema.
• Zygonema The chromosomes continue to shorten and thicken during the zygotene stage. During the
process of homology search, homologous chromosomes undergo initial alignment with one another.
This so-called rough pairing is complete by the end of zygonema.
• It is upon completion of zygonema that the paired homologs are referred to as bivalents.
• Although both members of each bivalent have already replicated their DNA, it is not yet visually apparent
that each member is a double structure.
• The number of bivalents in each species is equal to the haploid (n) number.

• Pachynema In the transition from the zygotene to the pachytene stage, the chromosomes continue to coil
and shorten, and further development of the synaptonemal complex occurs between the two members of each
bivalent.
• This leads to synapsis, a more intimate pairing. Compared to the rough-pairing characteristic of zygonema,
homologs are now separated by only 100 nm.
• During pachynema, each homolog is now evident as a double structure, providing visual evidence of the
earlier replication of the DNA of each chromosome.
• Thus, each bivalent contains four member chromatids.
• As in mitosis, replicates are called sister chromatids, whereas chromatids from maternal and paternal
members of a homologous pair are called nonsister chromatids.
• The four-membered structure, also referred to as a tetrad, contains two pairs of sister chromatids.
• Diplonema During the ensuing diplotene stage, it is even more apparent that each tetrad consists of two
pairs of sister chromatids.
• Within each tetrad, each pair of sister chromatids begins to separate.
• However, one or more areas remain in contact where chromatids are intertwined called as chiasma (pl.
chiasmata)
• It is thought to represent a point where non-sister chromatids have undergone genetic exchange through the
process referred to above as crossing over.
• Although the physical exchange between chromosome areas occurred during the previous pachytene stage,
the result of crossing over is visible only when the duplicated chromosomes begin to separate.
• Crossing over is an important source of genetic variability, and as indicated earlier, new combinations of
genetic material are formed during this process.
• Diakinesis The final stage of prophase I is diakinesis.
• The chromosomes pull farther apart, but non-sister chromatids remain loosely associated at the chiasmata.
• As separation proceeds, the chiasmata move toward the ends of the tetrad.
• This process of terminalization begins in late diplonema and is completed during diakinesis.
• During this final substage, the nucleolus and nuclear envelope break down, and the two centromeres of
each tetrad attach to the recently formed spindle fibers.
• By the completion of prophase I, the centromeres of each tetrad structure are present on the metaphase plate
of the cell.
• During the metaphase I, the homologs attach to opposite poles of the microtubule-based spindle, as both
kinetochores of each sister-chromatid pair are attached to the same pole of the microtubule spindle, it is called
monovalent attachment.
• As in mitosis, the paired homologs initially resist the tension of the spindle pulling them apart. In the case of
meiosis I, this resistance is mediated through the physical connections between the homologs, called chiasma or
crossovers, that are the result of recombination between the homologs.
• This resistance also requires sister-chromatid cohesion along the arms of the sister chromatids. When cohesion
along the arms is eliminated during anaphase I, the recombined homologs are released from each other and
segregate to opposite poles of the cell. Importantly, the cohesion between the sisters is maintained near the
centromere, keeping the sister chromatids paired.
• The second round of segregation during meiosis, meiosis II, is very similar to mitosis. The major difference is
that a round of DNA replication does not precede this segregation event. Instead, a spindle is formed in
association with each of the two newly separated sister-chromatid pairs. As in mitosis, during metaphase II,
these spindles attach in a bivalent manner to the kinetochores of each sister-chromatid pair.
• The cohesion that remains at the centromeres after meiosis I is critical to oppose the pull of the spindle.
• The second round of chromosome segregation occurs in anaphase II and is initiated by the elimination of
centromeric cohesion. At this point, there are four sets of chromosomes in the cell, each of which contains a
single copy of each chromosome. A nucleus forms around each set of chromosomes, and then the cytoplasm is
divided to form four haploid cells. These cells are now ready to mate to form new diploid cells.

You might also like