23.10.2014 Views

10 Yeast Growth and the Cell Cycle

10 Yeast Growth and the Cell Cycle

10 Yeast Growth and the Cell Cycle

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>10</strong> <strong>Yeast</strong> <strong>Growth</strong> <strong>and</strong> <strong>the</strong> <strong>Cell</strong> <strong>Cycle</strong><br />

<strong>10</strong>.1 Vegetative Reproduction in <strong>Yeast</strong><br />

A concise but extremely informative description of <strong>the</strong> events during <strong>the</strong> cell cycle [Nurse, 2000] may<br />

be cited for clarity: “In eukaryotes <strong>the</strong> cell cycle is divided into two phases, interphase <strong>and</strong> mitosis.<br />

Interphase consists of G1, S <strong>and</strong> G2 phases. Mitosis can be sub-divided into prophase, metaphase,<br />

anaphase <strong>and</strong> telophase. Chromosomes condense during prophase, align during metaphase,<br />

separate during anaphase <strong>and</strong> decondense during telophase.”<br />

Mitotic division is initiated when cells attain a critical cell size <strong>and</strong> are stimulated by exogenous or<br />

endogenous signals during interphase. A prerequisite is that DNA syn<strong>the</strong>sis guarantees proper<br />

duplication of <strong>the</strong> genetic material <strong>and</strong> that all substrates critical to anabolic pathways are available. In<br />

mammalian cells, <strong>the</strong> mitotic phase can be subdivided into a number of steps that are characterized by<br />

particular states of orienting <strong>the</strong> spindle pole bodies, sorting <strong>and</strong> movement of <strong>the</strong> duplicated<br />

chromosomes, <strong>and</strong> finally, cell separation after cytokinesis (Figure <strong>10</strong>-1). While all processes pertinent<br />

to <strong>the</strong> cell cycle have been highly conserved among all eukaryotes, ascomycetous fungi, <strong>and</strong> in<br />

particular S.cerevisiae, exhibit several peculiarities with respect to cell division <strong>and</strong> cytokinesis<br />

[Bouquin et al., 2000; Bidlingmaier et al., 2001].<br />

Figure <strong>10</strong>-1: Phases in <strong>the</strong> mitotic cycle.<br />

<strong>10</strong>.1.1 <strong>Yeast</strong> Budding<br />

Budding is <strong>the</strong> most common mode of vegetative growth in yeasts <strong>and</strong> multilateral budding is a<br />

typical reproductive characteristic of ascomycetous yeasts, including S. cerevisiae. <strong>Yeast</strong> buds are<br />

initiated when mo<strong>the</strong>r cells attain a critical cell size at a time coinciding with <strong>the</strong> onset of DNA<br />

syn<strong>the</strong>sis. This is followed by localized weakening of <strong>the</strong> cell wall <strong>and</strong> this, toge<strong>the</strong>r with tension<br />

exerted by turgor pressure, allows extrusion of cytoplasm into an area bounded by new cell wall<br />

material. The regulation of particular cell wall syn<strong>the</strong>tic enzymes <strong>and</strong> transport of specific bud plasma<br />

membrane receptors are key steps in <strong>the</strong> emergence of a bud. Chitin forms a ring at <strong>the</strong> junction


etween <strong>the</strong> mo<strong>the</strong>r cell <strong>and</strong> <strong>the</strong> newly emerging bud to finally result in <strong>the</strong> generation of a daughter<br />

cell. After cell separation, this ring will be retained at <strong>the</strong> surface of <strong>the</strong> mo<strong>the</strong>r cell <strong>and</strong> form <strong>the</strong> socalled<br />

bud scar (Figures <strong>10</strong>-2 <strong>and</strong> <strong>10</strong>-3), <strong>and</strong> a birth scar at <strong>the</strong> surface of <strong>the</strong> daughter. The number of<br />

bud scars left on <strong>the</strong> surface of a yeast cell is a useful determinant of cellular age. Several recent<br />

papers have been devoted to budding in yeast [Roemer et al., 1996; Barral et al., 1999; Barrett et al.,<br />

2000; Manning et al., 1999; Sheu et al., 2000; Swaroop et al., 2000; Vogel et al., 2000, Ni & Snyder,<br />

2001].<br />

Figure <strong>10</strong>-2: Budding yeast cell.<br />

Figure <strong>10</strong>-3: <strong>Yeast</strong> bud <strong>and</strong> bud scar.<br />

During bud formation, only <strong>the</strong> bud but not <strong>the</strong> mo<strong>the</strong>r cell will grow. Once mitosis is complete <strong>and</strong> <strong>the</strong><br />

bud nucleus <strong>and</strong> o<strong>the</strong>r organelles have migrated into <strong>the</strong> bud, cytokinesis commences <strong>and</strong> a septum is<br />

formed in <strong>the</strong> isthmus between mo<strong>the</strong>r <strong>and</strong> daughter. A ring of proteins, called septins (Figure <strong>10</strong>-4),<br />

are involved in positioning cell division in that <strong>the</strong>y define <strong>the</strong> cleavage plain which bisects <strong>the</strong> spindle<br />

axis at cytokinesis. These septins encircle <strong>the</strong> neck between mo<strong>the</strong>r <strong>and</strong> daughter for <strong>the</strong> duration of<br />

<strong>the</strong> cell cycle.<br />

In S. cerevisiae, cell size at division is asymmetrical with buds being smaller than mo<strong>the</strong>r cells when<br />

<strong>the</strong>y separate. Also cell division cycle times are different, because daughter cells need time (in G1<br />

phase) to attain <strong>the</strong> critical cell size before <strong>the</strong>y are prepared to bud.


Table <strong>10</strong>-1: Examples of components important for budding<br />

Genes Gene product characteristics Mutant phenotypes<br />

General bud-site selection<br />

R<strong>and</strong>om budding in haploids <strong>and</strong> diploids<br />

BUD1/RSR1<br />

Ras-related protein<br />

BUD2<br />

GTPase-activating protein (GAP)<br />

BUD5<br />

GDP-GTP exchange factor (GEF)<br />

Axial bud-site selection<br />

Bipolar budding in haploids<br />

BUD3<br />

Novel<br />

BUD4<br />

GTP_binding domain<br />

AKL1<br />

a-factor protease<br />

BUD<strong>10</strong>/AXL2<br />

Type-I plasma membrane glycoprotein<br />

Polarity establishment<br />

Round, multinucleate cells unable to bud<br />

CDC24<br />

GEF for Cdc42p<br />

CDC42<br />

Rho/Rac GTPase<br />

BEM1<br />

SH3 domains<br />

Diploid bud-site selection<br />

R<strong>and</strong>om budding in diploids/mo<strong>the</strong>r cells<br />

ACT1<br />

Actin<br />

SPA2<br />

Coiled-coil domain<br />

RSV161, RSV167<br />

SH3 domain<br />

BNI1, BUD6, BUD7 ?<br />

BUD8 ?<br />

BUD9 ?<br />

Septin-ring<br />

Bipolar budding in haploids<br />

CDC3, CDC<strong>10</strong>, CDC11, CDC12 <strong>10</strong>-nm filament ring components<br />

GTP-binding domain<br />

Budding is not a r<strong>and</strong>omized, uncontrolled process; cellular geometry is explicitly important in<br />

localizing budding sites. Numerous studies have endeavoured to explain at <strong>the</strong> cellular <strong>and</strong> molecular<br />

level how polarized cell growth is regulated <strong>and</strong> how <strong>the</strong> site of emerging buds is chosen. Bud site<br />

selection depends on several physiological <strong>and</strong> genetic factors. For example, cell mating type is<br />

important, <strong>and</strong> a <strong>and</strong> α haploid cells are kwon to exhibit an axial budding pattern, where as<br />

a/α diploids exhibit a bipolar budding pattern. Axial budding means that mo<strong>the</strong>r <strong>and</strong> daughter cells<br />

form a new bud near <strong>the</strong> preceding bud scar <strong>and</strong> birth scar, respectively. Bipolar budding is when<br />

daughter cells bud firstly away from <strong>the</strong>ir mo<strong>the</strong>r, while mo<strong>the</strong>r cells ei<strong>the</strong>r bud away or toward<br />

daughter cells.<br />

The hierarchy of cell polarity is governed by <strong>the</strong> interplay of various genes that dictate <strong>the</strong> orientation<br />

of cytoskeletal elements. For example, <strong>the</strong> bud-site selection genes (BUD genes) are required for<br />

determining <strong>the</strong> orientation of actin fibres, <strong>and</strong> genes for bud formation (such as CDC24, CDC42,<br />

BEM1) direct cell surface growth to <strong>the</strong> developing bud. Budding is strictly connected to events in <strong>the</strong><br />

cell cycle (see below) in that cyclins <strong>and</strong> cyclin-dependent kinases play a decisive role in actin<br />

assembly <strong>and</strong> in localizing <strong>and</strong> timing of bud emergence.<br />

It may be mentioned briefly that fission yeasts, like Schizosaccharomyces pombe, divide exclusively<br />

by forming a cell septum analogous to <strong>the</strong> mammalian cell cleavage furrow, which constricts <strong>the</strong> cell<br />

into two equal-sized daughters.<br />

<strong>10</strong>.1.2 <strong>Yeast</strong> Septins<br />

The septin proteins assemble into filaments that lie underneath <strong>the</strong> plasma membrane. In<br />

Saccharomyces cerevisiae, where <strong>the</strong>y were first identified, septins are visible as electron-dense<br />

cortical rings at <strong>the</strong> mo<strong>the</strong>r bud neck. In multicellular organisms, <strong>the</strong>y are found at <strong>the</strong> cleavage furrow<br />

<strong>and</strong> o<strong>the</strong>r cortical locations. Consistent with <strong>the</strong>ir localization, septins have been shown to be required<br />

for cytokinesis in yeast, Drosophila melanogaster, <strong>and</strong> mammalian cells.


Septins are highly conserved cytoskeletal elements found in fungi, mammals, <strong>and</strong> all eukaryotes<br />

examined thus far, with <strong>the</strong> exception of plants [Barral et al., 2000; Casamayor & Snyder, 2004].<br />

Recent evidence in yeast has demonstrated that septins participate in a variety of o<strong>the</strong>r cellular<br />

processes, including cell morphogenesis, bud site selection, chitin deposition, cell cycle regulation, cell<br />

compartmentalization, <strong>and</strong> spore wall formation. Since septins participate in many cellular processes,<br />

it is not surprising that a diverse set of proteins have been found associated with <strong>the</strong> yeast septin<br />

cytoskeleton.<br />

The septins have a highly conserved structure. They contain a central GTP-binding domain flanked by<br />

a basic region at <strong>the</strong> amino terminus, <strong>and</strong> most septins contain a coiled-coil domain at <strong>the</strong> carboxy<br />

terminus. In yeast, five septins, Cdc3, Cdc<strong>10</strong>, Cdc11, Cdc12, <strong>and</strong> Shs1, localize to <strong>the</strong> mo<strong>the</strong>r bud<br />

neck in vegetatively growing cells. Cdc3 <strong>and</strong> Cdc12 are essential for growth at all temperatures,<br />

whereas Cdc<strong>10</strong> <strong>and</strong> Cdc11 are required only at elevated temperatures. Shs1 is a nonessential septin.<br />

<strong>Cell</strong>s containing temperature- sensitive mutations in ei<strong>the</strong>r CDC3, CDC<strong>10</strong>, CDC11, or CDC12 delay at<br />

a G2 checkpoint <strong>and</strong> arrest at <strong>the</strong> restrictive temperature, forming extensive chains of highly elongated<br />

cells.<br />

Figure <strong>10</strong>-4: Functions of septins.<br />

Table <strong>10</strong>-2: Diversity of septin expression <strong>and</strong> function.<br />

Gene Function Localization Biochemistry<br />

CDC3<br />

CDC<strong>10</strong><br />

CDC11<br />

CDC12<br />

Bud neck, site of<br />

bud emergence,<br />

base of schmoo<br />

Essential for cytokinesis <strong>and</strong><br />

polar-bud growth control. Cdc3p<br />

<strong>and</strong> Cdc12p required for viability,<br />

but not Cdc<strong>10</strong>p <strong>and</strong> Cdc11pin<br />

some backgrounds. Reqired for<br />

proper regulation of <strong>the</strong> Gin4p <strong>and</strong><br />

Hsl1p kinases<br />

SEP7 Required in vivo for proper<br />

regulation of Gin4p kinase<br />

Bud neck, site of<br />

bud emergence<br />

SPR3 Sporulation efficiency Prospore wall No data<br />

SPR28 No obvious phentype Prospore wall No data<br />

Found in 370 kDa complex<br />

that can form filaments in<br />

vitro<br />

Can complex with Cdc3p,<br />

Cdc<strong>10</strong>, Cdc11p, Cdc12p


Table <strong>10</strong>-3: Budding yeast septin-protein interations.<br />

Protein Function Septindependent<br />

localization<br />

Gin4p<br />

Hsl1<br />

Kcc4<br />

Bni4p<br />

Chs3p<br />

Chs4p<br />

Yck1p<br />

Yck2p<br />

Bud3p<br />

Bud4p<br />

Spa2p<br />

Protein kinases that function in<br />

septin localization <strong>and</strong> cell cycle<br />

progression<br />

Reqired for normal chitin<br />

deposition<strong>and</strong> morphology<br />

Required for normal chitin<br />

syn<strong>the</strong>sis<br />

Casein kinase I homologs,<br />

required for septin localization,<br />

cytokinesis, morphogenesis <strong>and</strong><br />

endocytosis<br />

Bud site selection<br />

Yes<br />

Yes<br />

Yes<br />

Genetic/physical<br />

interactions<br />

Gin4p interacts<br />

physically <strong>and</strong><br />

genetically with septins<br />

Two-hybrid interaction<br />

with Cdc<strong>10</strong>p <strong>and</strong> Chs4p<br />

Syn<strong>the</strong>tic lethal<br />

interaction between<br />

Chs4p <strong>and</strong> Cdc12p<br />

Localization<br />

Bud neck, site of bud<br />

emergence<br />

Bud neck<br />

Bud neck, site of bud<br />

emergence<br />

No Unknown Bud neck, sites of<br />

polarized growth, plasma<br />

membrane<br />

Yes<br />

Yes<br />

??<br />

Unknown<br />

Unknown<br />

Syn<strong>the</strong>tic lethal<br />

interaction with Cdc<strong>10</strong>p<br />

Bud neck, site of bud<br />

emergence<br />

Bni1p Cytokinesis/morphogenesis ?? Syn<strong>the</strong>tic lethal<br />

Bud tip<br />

interaction with Cdc12p<br />

Myo1p Type II myosin. Plays a role in Yes Unknown Bud neck<br />

cytogenesis<br />

Arf1p Morphogenesis during mating ?? Two-hybrid interaction<br />

with Cdc12p<br />

Base of mating projections<br />

<strong>10</strong>.1.3 <strong>Yeast</strong> Spindle Pole Body Dynamics<br />

The spindle pole body (SPB) is <strong>the</strong> sole site of microtubule organization in <strong>the</strong> budding yeast<br />

Saccharomyces cerevisiae. SPBs are embedded in <strong>the</strong> nuclear envelope throughout <strong>the</strong> yeast life<br />

cycle <strong>and</strong> are <strong>the</strong>refore able to nucleate both nuclear <strong>and</strong> cytoplasmic microtubules. The small size of<br />

<strong>the</strong> yeast SPB, its location in a membrane, <strong>and</strong> <strong>the</strong> fact that nearly all genes involved in SPB function<br />

are essential have presented significant challenges in its analysis. Never<strong>the</strong>less, <strong>the</strong> SPB is perhaps<br />

<strong>the</strong> best-characterized microtubule organizing center (MTOC).<br />

Nucleation of microtubules by eukaryotic microtubule organizing centers is required for a variety of<br />

functions, including chromosome segregation during mitosis <strong>and</strong> meiosis, cytokinesis, fertilization,<br />

cellular morphogenesis, cell motility, <strong>and</strong> intracellular trafficking. Analysis of MTOCs from different<br />

organisms shows that <strong>the</strong> structure of <strong>the</strong>se organelles is widely varied even though <strong>the</strong>y all share <strong>the</strong><br />

function of microtubule nucleation. Despite <strong>the</strong>ir morphological diversity, many components <strong>and</strong><br />

regulators of MTOCs, as well as principles in <strong>the</strong>ir assembly, seem to be conserved [Segal et al.,<br />

2001; Jaspersen & Winey, 2004; Cheeseman & Desay, 2004].<br />

The SPB is a cylindrical organelle that appears to consist of three disks or plaques of darkly staining<br />

material (Figure <strong>10</strong>-5): an outer plaque that faces <strong>the</strong> cytoplasm <strong>and</strong> is associated with cytoplasmic<br />

microtubules, an inner plaque that faces <strong>the</strong> nucleoplasm <strong>and</strong> is associated with nuclear microtubules<br />

<strong>and</strong> a central plaque that spans <strong>the</strong> nuclear membrane. One side of <strong>the</strong> central plaque is associated<br />

with an electron-dense region of <strong>the</strong> nuclear envelope termed <strong>the</strong> half-bridge. This is <strong>the</strong> site of new<br />

SPB assembly because darkly staining material similar in structure to <strong>the</strong> SPB accumulates on its<br />

distal, cytoplasmic tip during G1 phase of <strong>the</strong> cell cycle.


Careful analysis of SPB size <strong>and</strong> structure indicates that <strong>the</strong> SPB is a dynamic organelle. In haploid<br />

cells, <strong>the</strong> SPB grows in diameter from 80 nm in G1 to 1<strong>10</strong> nm in mitosis. The molecular mass of a<br />

diploid SPB, including microtubules <strong>and</strong> microtubule associated proteins, is estimated to be 1–1.5<br />

GDa. However, only 17 components of <strong>the</strong> mitotic SPB have been identified to date (Table <strong>10</strong>-4).<br />

Table <strong>10</strong>-4: <strong>Yeast</strong> spindle pole body components.<br />

Protein SPB location Role in SPB function<br />

Tub4 γ-tubulin complex MT nucleation<br />

Spc98 γ-tubulin complex MT nucleation<br />

Spc97 γ-tubulin complex MT nucleation<br />

Spc72 OP, HB γ-tubulin binding protein<br />

Nud1 OP, satellite MEN signalling<br />

Cnm67 IL1, OP, satellite Spacer, anchors OP to CP<br />

Spc42 IL2, OP, satellite Structural SPB core<br />

Spc29 CP, satellite Structural SPB core<br />

Cmd1 CP Structural Spc1<strong>10</strong> binding protein<br />

Spc1<strong>10</strong> CP to IP Spacer, γ-tubulin binding protein<br />

Ndc1 SPB periphery Membrane protein, SPB insertion<br />

Mps2 SPB periphery Membrane protein, SPB insertion<br />

Bbp1 SPB periphery SPB core, HB linker to membrane<br />

Kar1 HB Membrane protein, SPB duplication<br />

Mps3 HB Membrane protein, SPB duplication<br />

Cdc31 HB SPB duplication<br />

Sfi1 HB SPB duplication<br />

Mpc54 MP Replace Spc72 in meosis I<br />

Spo21 MP Replace Spc72 in meosis II<br />

CP= Central plaque; HB=half-bridge; IL1=Inner layer 1; IL2= Inner layer 2; IP=Inner plaque;<br />

MEN= Microtubule envelope; cMT= Cytoplasmic MT (microtubule); nMT= Nuclear MT; OP=Outer plaque.<br />

Figure <strong>10</strong>-5: Location of protein components of <strong>the</strong> spindle pole body.


Regulators of SPB duplication <strong>and</strong> function associate with <strong>the</strong> SPB during all or part of <strong>the</strong> cell cycle.<br />

Mps1, a conserved protein kinase required for multiple steps in SPB duplication <strong>and</strong> also for <strong>the</strong><br />

spindle checkpoint, localizes to SPBs <strong>and</strong> to kinetochores<br />

Figure <strong>10</strong>-6: SPB duplication pathway.<br />

SPB duplication (Figure <strong>10</strong>-6) can be divided into three steps: (1) half-bridge elongation <strong>and</strong><br />

deposition of satellite material, (2) expansion of <strong>the</strong> satellite into a duplication plaque <strong>and</strong> retraction of<br />

<strong>the</strong> half-bridge, <strong>and</strong> (3) insertion of <strong>the</strong> duplication plaque into <strong>the</strong> nuclear envelope <strong>and</strong> assembly of<br />

<strong>the</strong> inner plaque. Following completion of SPB duplication, <strong>the</strong> bridge connecting <strong>the</strong> side-by-side<br />

SPBs is severed, <strong>and</strong> SPBs move to opposite sides of <strong>the</strong> nuclear envelope (4). The requirements for<br />

various gene products in each step are shown in <strong>the</strong> figure. SPC72, NUD1, <strong>and</strong> CNM67 are probably<br />

required for step 2. SPBs are not syn<strong>the</strong>sized de novo. Consequently, every time a cell divides it must<br />

duplicate its SPB, as well as its genome, to ensure that both <strong>the</strong> mo<strong>the</strong>r <strong>and</strong> daughter cell contain one<br />

copy of all 16 chromosomes <strong>and</strong> one SPB. SPB duplication occurs in G1 phase of <strong>the</strong> cell cycle;<br />

however, defects in SPB duplication are not detected until mitosis when cells fail to form a functional<br />

bipolar spindle. Generally, SPB defects cannot be reversed at this point, so cells will eventually<br />

attempt chromosome segregation with a monopolar spindle, which results in progeny with aberrant<br />

DNA content <strong>and</strong>/or SPB number. Therefore, accurate SPB duplication during G1 is essential to<br />

maintain genomic stability.<br />

<strong>10</strong>.2 The <strong>Yeast</strong> <strong>Cell</strong> <strong>Cycle</strong><br />

<strong>10</strong>.2.1 General<br />

The cell cycle can be defined as <strong>the</strong> period between division of a mo<strong>the</strong>r cell <strong>and</strong> subsequent division<br />

of its daughter progeny. The regulatory mechanisms that order <strong>and</strong> coordinate <strong>the</strong> progress of <strong>the</strong> cell<br />

cycle have been intensely studied [overviews: Mala & Nurse, 1998; Futcher, 2000; Lauren et al.,<br />

2001]. Numerous proteins that have been characterized through mutations are collectively designated<br />

as cell division cycle (Cdc) proteins.<br />

The eukaryotic cell cycle involves both continuous events (cell growth) <strong>and</strong> periodic events (DNA<br />

syn<strong>the</strong>sis <strong>and</strong> mitosis). Commencement <strong>and</strong> progression of <strong>the</strong>se events in yeast can formally been<br />

distinguished into pathways for DNA syn<strong>the</strong>sis <strong>and</strong> nuclear division, spindle formation, bud emergence<br />

<strong>and</strong> nuclear migration, <strong>and</strong> cytokinesis. However, from a molecular viewpoint <strong>the</strong>se processes are<br />

intimately coupled (Figure <strong>10</strong>-6).


Figure <strong>10</strong>-7: <strong>Cell</strong> cycle phases <strong>and</strong> physiological processes. (Modified from Hartwell, 2002).<br />

<strong>10</strong>.2.1.1 Some Historical Notes<br />

Early attention towards <strong>the</strong> ‘biology of <strong>the</strong> cell cycle’ arose from a book written by J.M. Mitchison which<br />

appeared in 1971 [Mitchison, 1971]. It was Lee Hartwell, who made <strong>the</strong> decisive step by introducing<br />

<strong>the</strong> budding yeast, S. cerevisiae, as an experimental system into this field <strong>and</strong> by characterizing a<br />

number of genes involved in cell division <strong>and</strong> cell cycle control (cf. Figure <strong>10</strong>-7), dubbed CDC genes<br />

[e.g., Hartwell et al., 1970; Hartwell, 1971; 1973; 1974]. He <strong>and</strong> his collaborators arrived at this issue<br />

after research on protein syn<strong>the</strong>sis <strong>and</strong> ribosome syn<strong>the</strong>sis in yeast [e.g., Hartwell, 1967; McLaughlin<br />

& Hartwell, 1969; Hartwell et al., 1970] as well as on studies of mating <strong>and</strong> mating pheromones in<br />

yeast (see below) carried out between 1967 <strong>and</strong> 1970. Research on <strong>the</strong> yeast cell cycle included work<br />

on synchronization of haploid yeast cells as a prelude to conjugation [Hartwell, 1973], on segregation<br />

[Wood & Hartwell, 1982] <strong>and</strong> <strong>the</strong> spindle checkpoint [Hartwell & Weinert, 1989; Paulovich et al., 1997].<br />

In 1980, Kim Nasmyth, whose fields comprised <strong>the</strong> mating phenomena in S. cerevisiae as well as <strong>the</strong><br />

cell cycle, succeeded in isolating cell cycle genes by molecular cloning [Nasmyth & Reed, 1980].<br />

Soon after Hartwell’s approach, <strong>the</strong> group of Paul Nurse chose to introduce <strong>the</strong> fission yeast, S.<br />

pombe, to this field as a fur<strong>the</strong>r simple model organism [Nurse et al., 1976] <strong>and</strong> to follow Hartwell’s<br />

approach by isolating Cdc mutants in fission yeast. The first mutants collected were mainly defective in<br />

<strong>the</strong> events of mitosis <strong>and</strong> cell division <strong>and</strong> subsequent screens carried out toge<strong>the</strong>r with Kim Nasmyth<br />

identified more mutants defective in S-phase [Nurse et al., 1976]. One important early finding was <strong>the</strong><br />

presence of a yeast homolog in S. pombe, cdc2, which later was shown to functionally correspond to<br />

<strong>the</strong> yeast CDC28 gene [Beach et al., 1982]. Fur<strong>the</strong>r, <strong>the</strong>re was <strong>the</strong> S. pombe wee1 gene that acted in<br />

G2 <strong>and</strong> controlled <strong>the</strong> cell cycle timing of mitosis [Nurse & Thuriaux, 1980]. Surprisingly, fur<strong>the</strong>r<br />

experiments disclosed that cdc2 was unusual in being required twice during <strong>the</strong> cell cycle, first in G1<br />

for onset of S-phase <strong>and</strong> <strong>the</strong>n again in G2 for onset of mitosis [Nurse & Bissett, 1981]. Obviously,<br />

S.pombe cdc2 had a central role controlling <strong>the</strong> fission yeast cell cycle. In G1 it was required to<br />

execute onset of S-phase, <strong>and</strong> in G2 it acted as a major rate limiting step determining <strong>the</strong> onset of<br />

mitosis. Next, <strong>the</strong> cdc2 gene product was identified as a kinase [Simanis & Nurse, 1986] <strong>and</strong> shown to


undergo tyrosine phosphorylation at <strong>the</strong> G2/M transition [Gould & Nurse, 1989]. A fur<strong>the</strong>r important<br />

player in control of <strong>the</strong> cell cycle turned out to be <strong>the</strong> Cdc13p cyclin, <strong>the</strong> level of which varied during<br />

<strong>the</strong> cell cycle, <strong>and</strong> which was required for Cdc2 protein kinase activation [Moreno et al., 1989].<br />

The proposition that cell cycle control was conserved in yeast <strong>and</strong> humans, <strong>and</strong> probably in all<br />

eukaryotes, was substantiated by isolating <strong>and</strong> characterizing equivalents for cdc2 [Lee & Nurse,<br />

1987]. The speculation was that human CDC2 might act at two points in <strong>the</strong> cell cycle, at <strong>the</strong> ‘G1<br />

restriction point’ known to operate in mammalian cells, <strong>and</strong> at <strong>the</strong> G2/M transition where it served as<br />

‘maturation promoting factor’ (MPF) known to control M-phase in metazoan eggs <strong>and</strong> oocytes [Nurse,<br />

1990]. These two functions accentuated <strong>the</strong> importance of cyclin-dependent kinases (CDKs) in<br />

regulating <strong>the</strong> orderly progression through S-phase <strong>and</strong> mitosis during <strong>the</strong> cell cycle. The onset of S-<br />

phase requires two sequential steps: <strong>the</strong> first one is only operative if CDK activity is absent (i.e. in<br />

early G1) whilst <strong>the</strong> second requires <strong>the</strong> presence of CDK activity, which later appears at <strong>the</strong> G1/S<br />

boundary, thus allowing progression through step two <strong>and</strong> bringing about <strong>the</strong> initiation of S-phase<br />

[Wuarin & Nurse, 1996]. During G2 <strong>the</strong> continued presence of CDK activity prevents step one from<br />

occurring again <strong>and</strong> this blocks onset of a fur<strong>the</strong>r S-phase. At <strong>the</strong> G2/M boundary, a fur<strong>the</strong>r increase in<br />

CDK activity triggers mitosis. Exit from mitosis <strong>and</strong> <strong>the</strong> ending of <strong>the</strong> cell cycle <strong>the</strong>refore requires<br />

destruction of CDK activity, <strong>and</strong> because <strong>the</strong> subsequent G1 cells lack CDK activity <strong>the</strong>y are able to<br />

carry out step one for S-phase <strong>and</strong> <strong>the</strong> whole series of events can be repeated [Stern & Nurse, 1996].<br />

On <strong>the</strong> long run, <strong>the</strong> findings from <strong>the</strong> two yeast models obviously had to be complemented by<br />

observations made in o<strong>the</strong>r organisms. An underst<strong>and</strong>ing what actually “drives” <strong>the</strong> cell cycle, came<br />

from most important discoveries in aquatic organisms: In 1980, Wu & Gerhart [1980] purified <strong>the</strong><br />

maturation promoting factor (MPF) from Xenopus eggs. Tim Hunt, who started research on <strong>the</strong> cell<br />

cycle <strong>the</strong> same year, was soon able to show <strong>the</strong> existence of <strong>the</strong> cyclins in sea urchins [Evans et al.,<br />

1982; Evans et al., 1983].<br />

Though many research groups contributed to this field, Hartwell, Hunt, <strong>and</strong> Nurse were honoured for<br />

<strong>the</strong>ir pioneering work <strong>and</strong> outst<strong>and</strong>ing discoveries in cell cycle research by awarding <strong>the</strong>m <strong>the</strong> Nobel<br />

Prize in 2001 [Hartwell, 2002; Nurse, 2001; Hunt, 2001]. The general principles come clear from Nurse<br />

[2000]: “There are several control points during <strong>the</strong> cell cycle: in late G1, called Start in yeast or <strong>the</strong><br />

Restriction point in mammals, in late G2 <strong>and</strong> just prior to anaphase. To pass each point, cells have to<br />

fulfil several prerequisites. Before passing Start, cells can undergo two developmental programmes,<br />

i.e. entry into <strong>the</strong> mitotic cell cycle or sexual development. Adequate nutritional conditions <strong>and</strong> a critical<br />

cell size are required to traverse Start. During G2, cells have to check whe<strong>the</strong>r DNA replication is<br />

completed <strong>and</strong> ensure that DNA is not damaged. Before chromosome separation cells also examine<br />

whe<strong>the</strong>r chromosomes are aligned <strong>and</strong> spindles are formed properly. These cell-cycle checkpoints are<br />

<strong>the</strong> mechanisms that govern <strong>the</strong> order of <strong>the</strong> cell-cycle events, because if <strong>the</strong> order of <strong>the</strong> events is<br />

incorrect <strong>the</strong>n a full complement of genetic information is not transmitted at cell division, which may<br />

lead to cancer in higher eukaryotes. Much is now known about regulation of <strong>the</strong> eukaryotic cell cycle<br />

but two areas have yet to be fully understood. The first area is concerned with <strong>the</strong> molecular<br />

mechanisms acting during <strong>the</strong> DNA-replication <strong>and</strong> DNA-damage checkpoints, which block mitosis if<br />

DNA replication is incomplete or DNA is damaged. The second concerns <strong>the</strong> controls which operate<br />

during <strong>the</strong> meiotic cell cycle.”


<strong>10</strong>.2.1.2 Periodic Events in <strong>the</strong> <strong>Cell</strong> <strong>Cycle</strong><br />

The periodic events can be divided into four phases (cf. Figure <strong>10</strong>-7): DNA syn<strong>the</strong>sis (S phase); a<br />

post-syn<strong>the</strong>tic gap (G2 phase); mitosis (M phase); <strong>and</strong> a pre-syn<strong>the</strong>tic gap (G1 phase). For division,<br />

yeast cells must reach a critical size. The key point in control of <strong>the</strong> cell cycle is START, <strong>the</strong> transition<br />

that initiates processes like DNA syn<strong>the</strong>sis in S phase, budding <strong>and</strong> spindle pole body duplication.<br />

Once cells have passed START, <strong>the</strong>y are irreversibly committed to replicating <strong>the</strong>ir DNA <strong>and</strong><br />

progressing through <strong>the</strong> cell cycle. START thus coordinates <strong>the</strong> cell cycle with cell growth. Nutrient<br />

starvation as well as induction of mating blocks passage through START. There are additional<br />

checkpoints that arrest cells during <strong>the</strong> cell cycle to avoid DNA damage or cell death due to events<br />

occurring out of order. These control points are situated at <strong>the</strong> G1-S <strong>and</strong> G2-M boundaries <strong>and</strong> can be<br />

considered as internal regulatory systems that arrest <strong>the</strong> cell cycle if prerequisites for progression are<br />

not met.<br />

After having passed <strong>the</strong> cell-size dependent START checkpoint, <strong>the</strong> level of cyclins (Cln,Clb)<br />

dramatically increase. Cyclins are periodically expressed <strong>and</strong> different cyclins (at least 11 in yeast)<br />

are known to be involved in <strong>the</strong> control of G1 (G1 cyclins), G2 (B-type cyclins) <strong>and</strong> DNA syn<strong>the</strong>sis (S<br />

phase cyclins). G1 cyclins are transcriptionally regulated. Cln3p, a particular G1 cyclin, is a putative<br />

sensor of cell size, which acts by modulating <strong>the</strong> levels of o<strong>the</strong>r cyclins. In addition to cyclin<br />

accumulation, <strong>the</strong> activity of a cylin-dependent kinase (CDK) which is an effector of START, is<br />

induced; this is <strong>the</strong> gene product of CDC28. Cdc28p (also termed Cdk1p) couples with G1 cyclins that<br />

activate its kinase potential. Homologues of this 34 kDa protein have been characterized in o<strong>the</strong>r<br />

eukaryotes. Cdk1 as well as being essential for S phase, is also important in controlling entry into<br />

mitosis. Complex formation of Cdk1 has been established with Cln1-3 (at G1), with Clb5,6 (at S), Clb<br />

3,4 (at S/G2) <strong>and</strong> Clb1,2 (at M). Alternation of cell cycle phases appears to be due to mechanisms that<br />

one cyclin family succeeds ano<strong>the</strong>r. The level of cyclins are controlled by syn<strong>the</strong>sis <strong>and</strong> programmed<br />

proteolysis by <strong>the</strong> proteasome. In this regard it has been shown that G2 cyclins are necessary for<br />

degradation of G1 cyclins, <strong>and</strong> that G2 cyclin syn<strong>the</strong>sis is coupled to removal of G1 cyclins (Figure <strong>10</strong>-<br />

8).<br />

Important players in this game are inhibitor proteins, known as CKIs, which block CDK activity in G1.<br />

They represent a key mechanism by which <strong>the</strong> onset of DNA replication is regulated. One such<br />

inhibitor is Sic1p: upon its destruction by programmed proteolysis, cyclin-Cdk1 activity is induced. This<br />

degradation <strong>the</strong>n triggers <strong>the</strong> G1-S transition [Lauren et al., 2001].


Figure <strong>10</strong>-8: Regulation of <strong>the</strong> yeast cell cycle.<br />

In addition to this pathway, cell cycle progression is controlled by <strong>the</strong> availability of nutrients (Figure<br />

<strong>10</strong>-9). Nutrient levels (for example, glucose or nitrogenous compounds) regulate <strong>the</strong> intracellular<br />

concentration of cAMP via a small G protein, Ras. The so-called Ras/cAMP pathway is well<br />

documented (see chapter 13). Decreasing levels lead to G1 arrest, while increasing levels induce <strong>the</strong><br />

cAMP-dependent protein kinase (PKA), which <strong>the</strong>n phosphorylates <strong>and</strong> <strong>the</strong>reby activates specific<br />

transcription factors involved in START.<br />

Figure <strong>10</strong>-9: G1 regulation in yeast.


G2-M control is characterized by <strong>the</strong> association of Cdk1 with B-type cyclins. Complexing of <strong>the</strong> kinase<br />

with <strong>the</strong> cyclins activates <strong>the</strong> kinase, leading to induction of M phase. At <strong>the</strong> end of M phase, <strong>the</strong><br />

mitotic cyclins are removed by programmed proteolysis.<br />

<strong>10</strong>.2.2 DNA Replication<br />

Chromosome duplication is central to cell division <strong>and</strong> is tightly controlled during <strong>the</strong> cell cycle. In<br />

eukaryotic cells, chromosome duplication is accomplished by initiating replication forks at many origins<br />

of replication on each chromosome; activation of <strong>the</strong> different replication origins is coordinated<br />

during S phase [Donaldson et al., 1999; Stillman, 2001; Raghuraman et al, 2001; Iyer et al., 2001;<br />

Heun et al., 2001; Wyrick et al., 2001]. ARS elements as substantial elements in yeast replication<br />

origins have been discussed above (chapter 2).<br />

Figure <strong>10</strong>-<strong>10</strong>: Licensing by <strong>the</strong> origin recognition complex.<br />

A two-step model of replication initiation, in which origins are licensed for firing during G1 but only<br />

activated under cellular conditions that preclude <strong>the</strong>ir licensing, has been proposed. In yeast, <strong>the</strong> sixprotein<br />

origin recognition complex (ORC) remains bound to DNA throughout <strong>the</strong> cell cycle <strong>and</strong><br />

forms <strong>the</strong> core of <strong>the</strong> origin complex to which o<strong>the</strong>r protein components are recruited. Interestingly,<br />

ORC is also involved in silencing of gene expression. The first step in pre-replicative complex<br />

formation (Figure <strong>10</strong>-<strong>10</strong>) appears to be <strong>the</strong> recruitment of Cdc6p by ORC, after which Cdc6p is in turn<br />

required for loading of Mcm/P1 protein complexes onto DNA, resulting in <strong>the</strong> origin being licensed for<br />

replication in <strong>the</strong> subsequent S phase. Cdc6p is an ATPase, <strong>and</strong> ATP hydrolysis seems essential for<br />

<strong>the</strong> loading of Mcm/P1. The role of <strong>the</strong> Mcm/P1 gene products is not sufficiently clarified yet. This<br />

protein family consists of six members, all of which are involved in <strong>the</strong> licensing process but <strong>the</strong>ir<br />

biochemical function is unclear. It may well be that <strong>the</strong>y supply helicase function required in both<br />

initiation <strong>and</strong> elongation phases of replication. Following Cdk activation in late G1, Cdc45p becomes<br />

associated with <strong>the</strong> origin; removal of Cdc6p may stimulate this event.


Figure <strong>10</strong>-11: Regulation of replication by cyclins.<br />

Interestingly, DNA replication must be restricted to only one round in each cell cycle. Work in yeast<br />

has shown that Cdk activity during G2 is required to prevent re-replication of DNA; <strong>the</strong> mechanism of<br />

this inhibition is not known, only <strong>the</strong> fact that <strong>the</strong> Cdks are involved in promoting <strong>the</strong> degradation of<br />

Cdc6p.<br />

Progression through <strong>the</strong> cell cycle is highly coordinated. Replication origin firing during S phase is not<br />

r<strong>and</strong>om but ra<strong>the</strong>r is under strict temporal <strong>and</strong> spatial control. Replication forks cluster in discrete<br />

'replication factories' within <strong>the</strong> nucleus <strong>and</strong> components required for elongation associate with nuclear<br />

structural components such as <strong>the</strong> lamina. Definitely, early <strong>and</strong> late origins have to be distinguished.<br />

Factors that share responsibility for promoting S phase are two B type cyclins, Clb5p <strong>and</strong> Clb6p, in<br />

conjunction with a single cyclin-dependent kinase, Cdc28p. As it apperas, Clb5p is executing <strong>the</strong> origin<br />

firing programme in both early <strong>and</strong> late origins, while Clb6-Cdc28 can only fire early replication origins.<br />

Fur<strong>the</strong>r, <strong>the</strong> origin-firing programme is subject to checkpoint controls (Figure <strong>10</strong>-11). One of <strong>the</strong><br />

essential players is Rad53p, which is involved in monitoring successful execution of <strong>the</strong> programme of<br />

DNA replication during S phase, <strong>and</strong> co-ordinating a controlled arrest if problems are encountered.<br />

Rad53p also seems to be required for maintaining <strong>the</strong> level of nucleotides in <strong>the</strong> normal S phase.<br />

The six subunits of <strong>the</strong> ORC complex, which was shown to be also involved in transcriptional<br />

silencing, were isolated <strong>and</strong> characterized [Diffley & Cocker, 1992; Micklem et al., 1993; Foss et al.,<br />

1993; Rao & Stillman, 1995; Loo et al., 1995; Bell et al.,1995; Li et al., 1998; Du & Stillman, 2002].<br />

Prereplicative (or preinitiation) complexes are assembled during M <strong>and</strong> G1 phase by binding of <strong>the</strong>


ORC complex to <strong>the</strong> multiple ARS sequences, whereby this binding persists throughout <strong>the</strong> cell cycle.<br />

In late M phase, most importantly, <strong>the</strong> ATP-dependent protein Cdc6 is recruited by <strong>the</strong> ORC, which in<br />

turn promotes loading of <strong>the</strong> minichromosome-maintenance (MCM) complex on to chromatin [Liang et<br />

al., 1995; Williams et al., 1997; Zou et al., 1997; Weinreich et al., 1999; Weinreich et al., 2001;<br />

Raghuraman et al., 2001; Stillman, 2001; Stillman, 2005]. Activation of two protein kinase complexes,<br />

Cdc28-B cyclins <strong>and</strong> Cdc7-Dbf4, <strong>the</strong>n serves as <strong>the</strong> final signal activating replication fork movement,<br />

whereupon <strong>the</strong> DNA-replication machinery, including DNA polymerases <strong>and</strong> proliferating-cell nuclear<br />

antigen (PCNA), initiates DNA syn<strong>the</strong>sis (cf. Figure 2-13).<br />

<strong>10</strong>.2.3 Spindle Dynamics<br />

In S. cerevisiae, <strong>the</strong> mitotic spindle must orient along <strong>the</strong> cell polarity axis, defined by <strong>the</strong> site of bud<br />

emergence, to ensure correct nuclear division between <strong>the</strong> mo<strong>the</strong>r <strong>and</strong> daughter cells [Segal <strong>and</strong><br />

Bloom, 2001]. Establishment of spindle polarity dictates this process <strong>and</strong> relies on <strong>the</strong> concerted<br />

control of spindle pole function <strong>and</strong> a precise programme of cues originating from <strong>the</strong> cell cortex that<br />

directs <strong>the</strong> cytoplasmic microtubule attachments during spindle morphogenesis. This cues cross talk<br />

with <strong>the</strong> machinery responsible for bud site selection, indicating that orientation of <strong>the</strong> spindle is<br />

mechanistically coupled to <strong>the</strong> definition of a polarity axis <strong>and</strong> <strong>the</strong> division plane.<br />

Spindle morphogenesis in yeast is initiated by <strong>the</strong> execution of START at <strong>the</strong> G1-S transition of <strong>the</strong><br />

cell cycle. Progression through START triggers bud emergence, DNA replication <strong>and</strong> <strong>the</strong> duplication of<br />

<strong>the</strong> microtubule-organizing centre (MTOC) - <strong>the</strong> spindle pole body (SPB) (Figures <strong>10</strong>-5 <strong>and</strong> <strong>10</strong>-12).<br />

The single stages of mitosis <strong>and</strong> intracellular movements can be distinguished by time-lapse phasecontrast<br />

microscopy. In addition to <strong>the</strong> polymerization <strong>and</strong> depolymerization of tubulin (<strong>the</strong> major<br />

microtubular protein), cytolasmic dynein is a mechanochemical enzyme or motor protein which drives<br />

microtubules motility in yeast. Actin filaments, ei<strong>the</strong>r as cytoskeletal cables or as cortical membrane<br />

patches, undergo dynamic changes during <strong>the</strong> cell cycle. The microtubules emanate from <strong>the</strong> SPBs<br />

toward <strong>the</strong> new bud <strong>and</strong> orientate <strong>the</strong> nucleus <strong>and</strong> intranuclear spindle at mitosis. The nuclear<br />

membrane remains intact throughout mitosis with <strong>the</strong> mitotic spindle forming intranuclearly between<br />

two SPBs embedded in <strong>the</strong> nuclear envelope. Once <strong>the</strong> genome replicates, <strong>the</strong> spindle aligns parallel<br />

to <strong>the</strong> mo<strong>the</strong>r bud axis <strong>and</strong> elongates eventually to provide each cell with one nucleus.<br />

The program for <strong>the</strong> establishment of spindle polarity, primed by cellular factors partioning<br />

asymmetrically between <strong>the</strong> bud <strong>and</strong> <strong>the</strong> mo<strong>the</strong>r cortex, coupling of this process to bud site selection<br />

<strong>and</strong> polarized growth has been elucidated in some detail [Segal & Bloom, 2001]. Several cortical<br />

components implicated in spindle orientation such as Bni1p, a target of <strong>the</strong> polarizing machinery<br />

essential in bud site selection <strong>and</strong> spindle orientation, <strong>and</strong> <strong>the</strong> actin interactor Aip3p/Bud6p are initially<br />

localized to <strong>the</strong> bud tip. O<strong>the</strong>r cortical elements (e.g. Num1p) are restricted initially to <strong>the</strong> mo<strong>the</strong>r cell<br />

during spindle assembly.


Figure <strong>10</strong>-12: Spindle dynamics.<br />

Figure <strong>10</strong>-13: Fluorescence imaging of microtubules.<br />

Factors mediating <strong>the</strong> process of microtubule attachment with <strong>the</strong> bud cell cortex are Bim1p <strong>and</strong><br />

Kar9p. Bim1p can directly bind to microtubules <strong>and</strong> is required for <strong>the</strong> high dynamic instability of<br />

microtubules that is characteristic of cells before spindle assembly. Kar9p has been implicated in <strong>the</strong><br />

orientation of functional microtubule attachments into <strong>the</strong> bud during vegetative growth. It is delivered<br />

to <strong>the</strong> bud by a Myo2-dependent mechanism presumably tracking on actin cables. Interaction of <strong>the</strong><br />

two factors, Bim1p <strong>and</strong> Kar9p, appears to provide a functional linkage between <strong>the</strong> actin <strong>and</strong><br />

microtubule cytoskeletons. In addition, Bud3p, a protein for axial budding of haploid cells, accumulates<br />

at <strong>the</strong> bud neck <strong>and</strong> is required for <strong>the</strong> efficient association of Bud6p to <strong>the</strong> neck region. Fur<strong>the</strong>r, a<br />

variety of motor proteins are necessary in spindle morphogenesis: dynein <strong>and</strong> <strong>the</strong> kinesin-like proteins<br />

Kip2p <strong>and</strong> Kip3p, as well as Kar3p are involved in regulating microtubule dynamics, mediating nuclear<br />

migration to <strong>the</strong> bud neck <strong>and</strong> facilitating spindle translocation (Figure <strong>10</strong>-13).<br />

<strong>10</strong>.2.4 Telomere Replication<br />

Several factors have been found to be implicated in stabilising telomeres <strong>and</strong> regulating telomere<br />

length; one of <strong>the</strong> earliest identified was Rap1, binding to both silencer <strong>and</strong> activator elements [Shore<br />

& Nasmyth, 1987; Shore et al., 1987; Kurtz & Shore, 1991]. Many discoveries in this field go back to<br />

<strong>the</strong> work of D. Shore <strong>and</strong> colleagues, who also found two proteins, Rif1 <strong>and</strong> Rif2, [Hardy et al., 1992;<br />

Wotton & Shore, 1997] <strong>and</strong> a SIR complex (Sir2,3 <strong>and</strong> 4) [Moretti et al., 1994] interacting with Rap1.


Remarkably <strong>the</strong>se factors are involved in telomere length regulation [Lustig et al., 1990; Hardy et al.,<br />

1992; Wotton & Shore, 1997; Shore, 2001, 2004]. Telomere length regulation is an issue for long<br />

discussed as a decisive phenomenon in cellular senescence <strong>and</strong> ageing [Shore, 1997, 1998; Smeal &<br />

Guarente, 1997; Blackburn et al., 2006] pertinent to all eukaryotic organisms.<br />

Because of <strong>the</strong>ir 'open-end' structure, telomeres have to be replicated by a specialized telomerase<br />

system [Cohn & Blackburn, 1995]. <strong>Yeast</strong> telomerase consists of <strong>the</strong> gene products of three EST genes<br />

(EST1, EST2, EST3) [Taggart et al., 2002; Taggart & Zakian, 2003; Lundblad, 2003], whereby Est2p<br />

acts as <strong>the</strong> catalaytic subunit, <strong>and</strong> an RNA component (TLC1) which is employed as a matrix in <strong>the</strong><br />

syn<strong>the</strong>sis of telomeric DNA [Brigati et al., 1993]. Additionally, telomere replication is dependent on:<br />

(i) <strong>the</strong> TRF1 complex, consisting of Ku70 (Hdf1) <strong>and</strong> Ku80 (Hdf2) proteins <strong>and</strong> interacting with<br />

Cdc13p; which is also crucial for non-homologous DNA-double str<strong>and</strong> repair <strong>and</strong> protects telomeres<br />

against nucleases <strong>and</strong> recombinases [e.g., Stellwagen et al., 2003; Fisher et al., 2004],<br />

(ii) a number of RAD proteins (Rad50, Rad51, Rad52),<br />

(iii) Sgs1p, a helicase, preventing deletirious recombination between telomeric sequences,<br />

(iv) <strong>and</strong> a number of o<strong>the</strong>r proteins such as Pif1p detected in Zakian’s group [Zhou et al., 2000].<br />

The participation of helicase Pif1p in telomere replication [Boule & Zakian, 2006] as well as <strong>the</strong><br />

involvement of <strong>the</strong> telomere replication apparatus in healing DNA breaks [Bianchi et al., 2004] have<br />

been solved in <strong>the</strong> budding yeast model (Figure <strong>10</strong>-14). Pif1p also helps flap elongation during<br />

Okazaki fragment maturation. Rrm3p, a helicase belonging to <strong>the</strong> Pif family, appears to be involved in<br />

replication fork progression [Boule & Zakian, 2006].<br />

Figure <strong>10</strong>-14: Models explaining Pif1p action at telomeres <strong>and</strong> during Okazaki<br />

fragment processing, <strong>and</strong> Rrm3p action during replication fork progression.<br />

(Reproduced from Boule <strong>and</strong> Zakian, 2006).<br />

In telomerase-deficient cells, telomeres shorten progressively (in about 60 generations) <strong>and</strong> lead to a<br />

shortening of telomeres <strong>and</strong> increased senescence. Two types of survival pathways are known to be<br />

induced upon defects in <strong>the</strong> telomerase system, which consist of telomere elongation by break-


induced replication (BIR): type I survivors maintain short TG1-3 repeats but amplify <strong>the</strong> Y' repeats, while<br />

type II survivors amplify <strong>the</strong> TG1-3 repeats to several kb in length but do not amplify <strong>the</strong> Y' elements.<br />

During chromosome replication in yeast, telomeres connect to <strong>the</strong> spindle pole bodies (SPBs), <strong>and</strong><br />

<strong>the</strong>re are multiple pathways for telomere te<strong>the</strong>ring [Taddei & Gasser, 2004].<br />

Recent reviews dealing with <strong>the</strong> mechanism of telomere replication <strong>and</strong> <strong>the</strong> participation of chromatin<br />

remodelling factors can be found in [Teixeira & Gilson, 2005; Gilson & Geli, 2007].<br />

<strong>10</strong>.2.5 Sister Chromatid Cohesion <strong>and</strong> Separation<br />

Sister chromatid cohesion is essential for accurate chromosome segregation during <strong>the</strong> cell cycle<br />

[Nasmyth, 1999; Biggins & Murray, 1999; Robert et al., ; Nasmyth, 2002; Camobel & Cohen-Fix;<br />

2002; Uhlmann, 2004]. A number of structural proteins are required for sister chromatid cohesion <strong>and</strong><br />

<strong>the</strong>re seems be a link in some organisms between <strong>the</strong> processes of cohesion <strong>and</strong> condensation.<br />

Likewise, a number of proteins that induce <strong>and</strong> regulate <strong>the</strong> separation of sister chromatids have been<br />

identified.<br />

As long ago as 1879, Fleming noticed that “<strong>the</strong> impetus causing nuclear threads to split longitudinally<br />

acts simultaneously on all of <strong>the</strong>m” [Fleming, 1879]. Chromosome splitting is an irreversible process<br />

<strong>and</strong> must <strong>the</strong>refore be highly regulated. Damage to <strong>the</strong> genome cannot easily be repaired by<br />

recombination nor can aberrant chromosome alignments be corrected, once sister chromatids<br />

separate from one ano<strong>the</strong>r. Therefore, sister chromatids have to be kept toge<strong>the</strong>r during mitosis after<br />

chromosome replication <strong>and</strong> <strong>the</strong>y have to be disentangled in metaphase before cell division starts in<br />

anaphase (Figure <strong>10</strong>-15). Keeping toge<strong>the</strong>r sister chromatids is absolutely required until all control<br />

mechanisms have been exerted that guarantee intactness of all replicated chromosomes.<br />

However, chromosomes do not remain inactive at this process: cohesion between sister chromatids<br />

generates <strong>the</strong> tension by which cells align <strong>the</strong>m on <strong>the</strong> metaphase plate. Cohesion also prevents<br />

chromosomes falling apart because of double-str<strong>and</strong>ed breaks <strong>and</strong> facilitates <strong>the</strong>ir repair by<br />

recombination.<br />

To date, we have ample knowledge of <strong>the</strong> molecular subtleties of how sister chromatids are kept<br />

toge<strong>the</strong>r <strong>and</strong> separated at appropriate moments of cell division, largely stemming from yeast as a<br />

model system. Before <strong>the</strong>se details were revealed, Koshl<strong>and</strong> & Hartwell [1987] had used SV40<br />

minichromosomes to show that <strong>the</strong> sister molecules are properly segregated when a cell cycle block is<br />

removed, demonstrating that sister minichromosome molecules need not remain topologically<br />

interlocked until anaphase in order to be properly segregated, <strong>and</strong> topological interlocking of sister<br />

DNA molecules apparently is not <strong>the</strong> primary force holding sister chromatids toge<strong>the</strong>r. This system has<br />

been kept since to study how sister chromatids behave during segragation [Ivanov & Nasmyth, 2005].<br />

Cohesion is mediated by a multisubunit complex called cohesin, which binds to chromosomes at<br />

multiple sites along chromosomes, from telophase until <strong>the</strong> onset of anaphase in <strong>the</strong> next cell cycle<br />

[Hirano, 2000; Nasmyth, 2001]. Cohesin consists of four core subunits, namely Smc1, Smc3, Scc1,<br />

<strong>and</strong> Scc3. Smc1 <strong>and</strong> Smc3 proteins (SMC complex) are characterized by 50-nm-long anti-parallel<br />

coiled coils flanked by a globular hinge domain <strong>and</strong> an ABC-like ATPase head domain. Whereas<br />

Smc1 <strong>and</strong> Smc3 heterodimerize via <strong>the</strong>ir hinge domains, <strong>the</strong> kleisin subunit Scc1 connects <strong>the</strong>ir<br />

ATPase heads, <strong>and</strong> this results in <strong>the</strong> formation of a large ring (Figure <strong>10</strong>-15). Cleavage of Scc1 by a<br />

cystein protease called ‘separin’ or ‘separase’ (Esp1 in yeast) triggers poleward movement of sisters


at <strong>the</strong> metaphase-to-anaphase transition (see below). Scc1 in turn recruits <strong>and</strong> binds <strong>the</strong> fourth<br />

cohesion subunit, Scc3, which has two orthologues in mammals.<br />

Figure <strong>10</strong>-16: Cohesins: <strong>the</strong> ‘glue’ between sister chromatids<br />

An early postulate was that cohesin connects sister DNA molecules through <strong>the</strong> binding of its two<br />

heads to each sister DNA molecule, thus forming a ‘glue’ between <strong>the</strong> sister chromatids [Toth et al.,<br />

1999; Anderson et al., 2002]. However, <strong>the</strong> finding that <strong>the</strong> N- <strong>and</strong> C-termini of Scc1 bind, respectively,<br />

to <strong>the</strong> Smc3 <strong>and</strong> Smc1 heads of <strong>the</strong> Smc1/Smc3 heterodimer [Haering et al., 2002] suggested that<br />

cohesin forms a large proteinaceous ring in which DNA str<strong>and</strong>s could be trapped (Figure <strong>10</strong>-15). It<br />

was concluded that <strong>the</strong> connection between sisters may be a topological ra<strong>the</strong>r than a chemical one<br />

[Nasmyth, 2002]. This hypo<strong>the</strong>sis turned out to be correct [e.g. Haering et al., 2002; Gruber et al.,<br />

2006]; it was confirmed <strong>and</strong> refined by experiments investigating <strong>the</strong> topological interaction between<br />

cohesin rings <strong>and</strong> a circular minichromosome [Ivanov & Nasmyth, 2005].<br />

With a diameter of close to 50 nm, <strong>the</strong> ring is sufficiently large to hold two sister DNA str<strong>and</strong>s toge<strong>the</strong>r<br />

even when wrapped around histones. At anaphase onset, it is <strong>the</strong> Scc1 subunit that is cleaved by<br />

separase, which in turn disrupts <strong>the</strong> interaction between <strong>the</strong> Smc heads in cohesin, thus opening <strong>the</strong><br />

ring [Weitzer et al., 2003; Uhlmann, 2003]. Critical to <strong>the</strong> cleavage of Scc1 is that its C-terminal<br />

cleavage product is quickly destroyed by targeted proteolysis [Rao et al., 2001] in order of preventing<br />

<strong>the</strong> Smc heads from interacting. Recently, it has been demonstrated that kleisin not only connects he<br />

Smc1 <strong>and</strong> Smc3 ATPase heads but also regulates <strong>the</strong>ir ATPase activity [Arumugam et al., 2006].<br />

For a long time, an unsolved problem was how <strong>the</strong> two DNA str<strong>and</strong>s are ‘entering’ <strong>the</strong> cohesin ring<br />

during (or after) DNA replication [Uhlmann, 2003]. Could <strong>the</strong> DNA replication fork simply slide through<br />

<strong>the</strong> cohesin rings that were put around DNA before S-phase? This would leave two replication<br />

products trapped inside <strong>the</strong> same ring without fur<strong>the</strong>r transport, <strong>and</strong> at <strong>the</strong> same time provide an<br />

intrinsic solution to <strong>the</strong> crucial requirement to only establish sister chromatid cohesion between<br />

au<strong>the</strong>ntic replication products <strong>and</strong> never between any o<strong>the</strong>r two sequences of DNA. It is in fact known<br />

that a number of yeast proteins, for example Eco1 (Cft7) [Skibbens et al., 1999; Toth et al., 1999;<br />

2000], or alternative chromatin remodeling complexes (RFC) [Mayer et al., 2001] as well as RSC<br />

[Huang et al., 2004] are required for efficient cohesion establishment during S-phase. These might<br />

even be regulators of a more sophisticated machinery which puts cohesin around both sister<br />

chromatids.


Recently this problem has come close to solution: Gruber et al. [2006] showed that cohesin’s hinges<br />

are not merely dimerization domains that are holding toge<strong>the</strong>r <strong>the</strong> cohesin ring by preventing kleisin’s<br />

dissociation from <strong>the</strong> SCM heads. Ra<strong>the</strong>r, entry of DNA into <strong>the</strong> cohesin ring requires opening <strong>and</strong><br />

transient dissociation of <strong>the</strong> Smc1 <strong>and</strong> Smc3 hinge domains [see also: Shintomi & Hirano, 2007].<br />

The binding of condensin (<strong>the</strong> second ‘glue’ complex) to chromatin has similarities with chromatin<br />

binding to cohesin [Lavoie et al., 2002]. Condensin, a 13 S complex, consists of two Smc proteins,<br />

Smc2p <strong>and</strong> Smc4p, <strong>and</strong> contains three o<strong>the</strong>r essential subunits, one of which is homologous to Scc1p;<br />

<strong>the</strong> newly found Smc5p-Smc6p complex preserves nuclear integrity [Torres-Rosell et al., 2005]. Just<br />

like cohesin, <strong>the</strong> topological structure of condensin is a ring. Condensin associates with chromatin<br />

independently of ATP, but ATP hydrolysis is needed for <strong>the</strong> binding reaction. A particular feat of<br />

condensin is that chromatin wraps around it, generating a torsion in <strong>the</strong> DNA [Wang et al., 2005]. Thus<br />

condensin is amenable of contributing to chromosome compaction; it also participates in DNA repair<br />

[Chen et al., 2004]. After partial removal of cohesin rings by <strong>the</strong> separase reaction, condensin replaces<br />

cohesin, both in mitosis <strong>and</strong> meiosis [Yu & Koshl<strong>and</strong>, 2005].<br />

What ensures ordered segregation <strong>and</strong> movement of chromatids to opposite poles of <strong>the</strong> cell?<br />

Segregation of chromosomes must be a tightly regulated process (cf. Figure <strong>10</strong>-16). The fraction of<br />

cohesin that persists on chromosomes until metaphase is responsible for holding sisters toge<strong>the</strong>r<br />

while <strong>the</strong>y bi-orient during prometaphase. It is an absolute requirement that <strong>the</strong> sister chromatids<br />

adopt an amphitelic orientation, i.e. that <strong>the</strong> spindle apparatus can be activated in a way that allows<br />

<strong>the</strong> traction of sister chromatids to opposite poles of <strong>the</strong> cell by <strong>the</strong> microtubules to occur. (Remember<br />

that in yeast microtubules connect kinetochores to spindle poles throughout <strong>the</strong> cell cycle).<br />

Sister chromatids are pulled to opposite 'halves' of <strong>the</strong> cell by microtubules that emanate from<br />

opposite spindle poles. These microtubules interdigitate <strong>and</strong> keep <strong>the</strong> two poles apart. Subsequently,<br />

a second set of microtubules attaches to chromosomes through specialized 'kinetochores' <strong>and</strong> pulls<br />

<strong>the</strong>m to <strong>the</strong> poles. In this way, sister chromatides separate <strong>and</strong> start to move into opposing poles.<br />

Figure <strong>10</strong>-16: Sister chromatid separation.


Two phases can be distinguished: (i) cohesin’s dissociation from <strong>and</strong> condensin’s association with<br />

chromosomes occurs in <strong>the</strong> complete absence of microtubules yet is capable of separating sister<br />

chromatids to ~0.5 µm. This first step in <strong>the</strong> individualisation process is triggered by <strong>the</strong> participation of<br />

PLK (‘Polo-like kinase’) which will phosphorylate cohesin (<strong>and</strong> possibly some o<strong>the</strong>r proteins). The<br />

second step, orienting sister chromatids on <strong>the</strong> mitotic spindle (Bi-orientation) <strong>and</strong> attaching<br />

microtubules to sister kinetochores, whereby chromosomes come under tension, is promoted by <strong>the</strong><br />

Aurora-like kinase Ipl1p <strong>and</strong> controlled by <strong>the</strong> spindle checkpoint (see below). (ii) The first clue to <strong>the</strong><br />

molecular basis for sister-chromatid separation in <strong>the</strong> second phase arose from <strong>the</strong> discovery of<br />

mitotic cyclins as proteins whose abundance fluctuates during embryonic-cleavage divisions [Evans et<br />

al., 1983] <strong>and</strong> that mitotic cyclins are regulatory subunits of a cyclin-dependent kinase (CDK1), whose<br />

activation in late G2 phase triggers mitosis. At commencement of anaphase, CDK activity is<br />

destructed by degrading of <strong>the</strong> cyclins [Glotzer et al., 1991], but is not required for <strong>the</strong> separation of<br />

sister chromatids [Surana et al., 1993; Dirick et al., 1995]. This suggested that proteolysis of fur<strong>the</strong>r<br />

proteins is necessary for sister chromatid separation [Holloway et al., 1993]. The apparatus<br />

responsible for targeting cyclin degradation turned out to be a highly conserved multi-subunit complex<br />

that possesses ubiquitin-protein ligase activity [Zachariae et al., 1998; Yu et al., 1998]. Initially named<br />

<strong>the</strong> cyclosome [Sudakin et al., 1995], <strong>the</strong> ligase moiety is now called <strong>the</strong> anaphase promoting<br />

complex (APC). It mediates <strong>the</strong> destruction of many proteins o<strong>the</strong>r than cyclins <strong>and</strong> was shown to be<br />

essential for <strong>the</strong> separation of sister chromatids [Irniger et al., 1995; King et al., 1995; Zachariae et. al.,<br />

1996; Michaelis et al., 1997; Guacci et al., 1997; Losada et al., 1998; Zachariae & Nasmyth, 1999].<br />

Ubiquitination mediated by APC/C requires rate-limiting activator proteins that bind to APC; in yeast<br />

<strong>the</strong>se are at least two proteins, namely Cdc20p <strong>and</strong> Cdh1p.These activators specify both substrate<br />

specificity <strong>and</strong> <strong>the</strong> timing of proteolysis. In <strong>the</strong> absence of APC/C function, yeast cells arrest in<br />

metaphase, <strong>and</strong> sister chromatids fail to segregate owing to <strong>the</strong> persistence of securin, <strong>the</strong> inhibitor of<br />

separase. Separase is activated by proteolysis of securin by APC/C.<br />

Once chromosomes have successfully bi-oriented, cells utilize an evolutionarily conserved machinery<br />

to initiate anaphase. Rapid disjunction of sister chromatids at anaphase requires cleavage of cohesin<br />

subunit Scc1 by a cysteine protease called separase (Esp1 in budding yeast), inactive through most of<br />

<strong>the</strong> cell cycle by its association with an inhibitor, securin [Yamamoto et al., 1996a; Ciosk et al., 1998;<br />

Uhlmann et al., 1999]. The metaphase-to-anaphase transition is triggered when securin (Pds1 in<br />

budding yeast [Cohen-Fix et al., 1996; Yamamoto et al., 1996b; Cohen-Fix & Koshl<strong>and</strong>, 1997]) is<br />

degraded by <strong>the</strong> proteasome following ubiquitination by <strong>the</strong> multicomponent E3 ubiquitin ligase known<br />

as <strong>the</strong> APC/C (cf. Figure 16). Securin accumulates within nuclei during late G1 phase, is maintained<br />

during G2 <strong>and</strong> early M phase, but degraded shortly before anaphase, so that separase becomes<br />

active. Separase resides in <strong>the</strong> cytoplasm until cells enter mitosis, whereupon it accumulates at <strong>the</strong><br />

mitotic spindle until late anaphase.<br />

APC/C function is regulated by (i) phosphorylation, <strong>and</strong> (ii) association of activator proteins such as<br />

Cdc 20 <strong>and</strong> its homologue Hct1 (homologue of Cdc20/Cdh1), which modulate <strong>the</strong> affinity of APC/C for<br />

different substrates [Peters, 2002]. The activation of <strong>the</strong> SCP inhibits <strong>the</strong> ubiquitin-dependent<br />

proteolysis of securin by <strong>the</strong> APC Cdc20 (APC/C activated by Cdc20) [Hwang et al., 1998].<br />

<strong>10</strong>.2.6 Spindle Checkpoint (SCP)


The activity of <strong>the</strong> APC is controlled by <strong>the</strong> spindle checkpoint, a mechanochemical surveillance<br />

mechanism shared by most eukaryotic cells, that prevents sister chromatid separation when spindles<br />

are damaged or chromosomes fail to form spindle attachments [Amon, 1999; Nasmyth, 2002; Lew,<br />

2003; Lew & Burke, 2003; Gillett et al., 2004; Tan et al., 2005]. The kinetochore of a single lagging<br />

chromosome emits a signal capable of blocking separation of all sister pairs. It also blocks any fur<strong>the</strong>r<br />

cell cycle progression.<br />

Figure <strong>10</strong>-17: The kinetochore. (Reproduced from Tan et al., 2005).<br />

In yeast, <strong>the</strong> kinetochore (Figure <strong>10</strong>-17) is composed of protein assemblies that can be broadly<br />

classified as inner, central or outer kinetochore complexes [Tan et al., 2005, Ciferri et al., 2007]. The<br />

outer kinetochore complex DAM1, composed of Duo1 <strong>and</strong> Mps1 (monopolar spindle 1-interacting<br />

complex), plays a crucial role in mediating <strong>the</strong> kinetochore–microtubular connection, <strong>and</strong> is regulated<br />

through phosphorylation by <strong>the</strong> Ipl1p/Aurora B kinase. The central complex contains <strong>the</strong> IPL1, CFT19,<br />

NDC80, <strong>and</strong> MWT1 complexes which are associated with both microtubules via <strong>the</strong> DAM1 complex<br />

<strong>and</strong> kinetochores via <strong>the</strong> inner complex, <strong>the</strong> most critical complex is CBF3 (centromere binding factor<br />

3). CBF3 consists of <strong>the</strong> essential proteins Ndc<strong>10</strong>, Cep3, Cft13, <strong>and</strong> Skp1, as well as a number of<br />

chromatin-specific proteins, which are required to build up a kinetochore at each centromere. The IPL1<br />

complex responds to <strong>the</strong> lack of tension in monotelic attachments, <strong>and</strong> acts to resolve <strong>the</strong>se<br />

inappropriate attachments, probably through its substrates. In <strong>the</strong> absence of tension, Ipl1 causes an<br />

increased turnover of kinetochore–microtubule connections, perhaps by influencing <strong>the</strong> Ndc80–DAM1<br />

interaction. Experiments in budding yeast have demonstrated that <strong>the</strong> tension resulting from <strong>the</strong><br />

physical connection between bi-oriented kinetochores, <strong>and</strong> <strong>the</strong> activity of Ipl1 (ra<strong>the</strong>r than any specific<br />

chromosomal architecture or kinetochore geometry) is sufficient for <strong>the</strong> proper alignment of sister<br />

chromatids [Dewar et al., 2004].<br />

The process of APC control is triggered by <strong>the</strong> recruitment of a complex to <strong>the</strong> SCP (Figure <strong>10</strong>-18),<br />

which contains <strong>the</strong> proteins Mad1p, Mad2p, Mad3p, <strong>and</strong> <strong>the</strong> protein kinases Bub1p <strong>and</strong> Bub3p, of<br />

which Bub1p is activated through Cdc2. Bub3p binds to <strong>the</strong> activator Cdc20p of <strong>the</strong> APC/C complex<br />

<strong>and</strong> <strong>the</strong>reby blocks ubiquitination of both securin <strong>and</strong> cyclin B. In addition, protein kinase Msp1p is<br />

also required for spindle pole duplication as well as <strong>the</strong> subunit Ncd<strong>10</strong>p of <strong>the</strong> centromere binding<br />

complex CBF3. The Mad complex has been shown to be highly conserved among o<strong>the</strong>r eukaryotes.<br />

The functions of APC <strong>and</strong> fur<strong>the</strong>r factors involved in regulating mitotic spindle disassembly have been


intensely described [Schuyler et al., 2003; Pereira & Schiebel, 2003; Buvelot et al., 2003; Tan et al.,<br />

2005].<br />

A more detailed view of <strong>the</strong> function of <strong>the</strong> APC is presented in Figure <strong>10</strong>-19.<br />

Figure <strong>10</strong>-18: The spindle checkpoint.<br />

Figure <strong>10</strong>-19: Different states of <strong>the</strong> spindle checkpoint. (Reproduced from Tan et al., 2005.)<br />

(A) Recruitment: Mad1 <strong>and</strong> Mad2 form a tight complex that localizes to NPC (nuclear pore complex). Upon SCP engagement,<br />

subcomplexes of checkpoint proteins are recruited to <strong>the</strong> kinetochore. Cdc2-dependent Bub1 phosphorylation is required to<br />

activate <strong>the</strong> SCP after spindle damage. The kinases Mps1 <strong>and</strong> Bub1 act in concert <strong>and</strong> trigger <strong>the</strong> rapid recruitment of <strong>the</strong><br />

Mad1–Mad2 complex that disengages from <strong>the</strong> NPC upon SCP activation. Recruitment of SCP complexes presumably occurs<br />

via <strong>the</strong>ir direct or indirect interaction with kinetochore components <strong>and</strong> o<strong>the</strong>r checkpoint proteins. The DAM1 complex has been<br />

shown to bind Mps1, Mad1, Mad3, Bub1 <strong>and</strong> Bub2. Bub1 binds Skp1. CENP-I is also essential for <strong>the</strong> kinetochore association<br />

of Mad1 <strong>and</strong> Mad2. Mad1 also interacts with <strong>and</strong> forms a complex with <strong>the</strong> Bub1–Bub3 heterodimer in such a way that Bub3<br />

binds both Mad1 <strong>and</strong> Mad2. The domain of Bub1 required for binding Bub3 is <strong>the</strong> same domain that is required for localization<br />

of Bub1 to kinetochores, suggesting that, upon association with kinetochores, Bub3 dissociates from Bub1. Bub3 also forms a<br />

complex with Mad3 throughout <strong>the</strong> cell cycle <strong>and</strong> may target Mad3 to kinetochores. In addition to <strong>the</strong> SCP signalling molecules<br />

depicted in <strong>the</strong> Figure, Cdc20 <strong>and</strong> <strong>the</strong> APC/C are also recruited to kinetochores upon SCP activation. Red arrows show<br />

phosphorylation events.<br />

(B) Reorganization <strong>and</strong> turnover: red arrows show phosphorylation events <strong>and</strong> green arrows show protein turning over. SCP<br />

activation results in <strong>the</strong> hyperphosphorylation of Mad1 by Mps1 <strong>and</strong> perhaps also Bub1, <strong>and</strong> causes <strong>the</strong> formation of <strong>the</strong> Mad1–<br />

Bub1–Bub3 complex. The Mad1–Mad2 complex recruited to <strong>the</strong> kinetochore functions as a template for <strong>the</strong> rapid catalytic<br />

conversion <strong>and</strong> release of soluble Mad2 in an APC Cdc20 -inhibitory form, labelled in purple. Mad3/BubR1 <strong>and</strong> Bub3 are also<br />

turned over at kinetochores <strong>and</strong> released, although it is not known if <strong>the</strong>y are modified in <strong>the</strong> process. The APC Cdc20 -inhibitory<br />

form of Mad2 is transferred to <strong>the</strong> Bub3–Cdc20–Mad3/BubR1 complex, which is a potent inhibitor of APC/C.<br />

(C) Inhibition of anaphase: <strong>the</strong> SCP deploys various strategies to inhibit anaphase onset. Targeting of substrates by APC Cdc20 is<br />

inhibited independently <strong>and</strong> collaboratively by <strong>the</strong> various Mad–Bub complexes. In in vitro assays, both BubR1 <strong>and</strong> Mad2 block<br />

<strong>the</strong> binding of Cdc20 to APC/C. Mitotic APC/C (with some APC/C subunits phosphorylated by Bub1 <strong>and</strong> Mps1) is more


susceptible to inhibition upon SCP activation as compared with APC/C from o<strong>the</strong>r stages of <strong>the</strong> cell cycle. It has also been<br />

shown that Cdk1–cyclin B-dependent inhibitory phosphorylation of Cdc20/Fizzy contributes to keeping APC/C inactive under<br />

conditions of SCP activation. Fur<strong>the</strong>rmore, SCP activation triggers Cdc20 degradation <strong>and</strong> keeps <strong>the</strong> level of Cdc20 low until all<br />

chromosomes have bi-oriented. Restoration of attachment/tension at kinetochores results in <strong>the</strong> dynein-dependent depletion of<br />

SCP proteins by transport away from kinetochores towards <strong>the</strong> poles along spindle MTs. Regulated translocation of Mad1 <strong>and</strong><br />

Mad2 away from kinetochores is believed to make <strong>the</strong> depletion less reversible. Bub1 persists longer than Mad1 <strong>and</strong> Mad2 at<br />

kinetochores, <strong>and</strong> <strong>the</strong> consequent physical separation of components of <strong>the</strong> SCP catalytic scaffold turns off <strong>the</strong> SCP signal,<br />

even though Mad2 continues to turn over at <strong>the</strong> spindle poles.<br />

<strong>10</strong>.3 Sexual Reproduction<br />

Though vegetative growth is <strong>the</strong> major way of yeast reproduction, sexual reproduction is an<br />

alternative when nutrient supplies fall short. The latter process involves <strong>the</strong> conjugation of cells of<br />

opposite mating type [Shimoda, 2004]. A heterokaryon with a diploid set of chromosomes is formed,<br />

which is capable of reproduction by budding. Under starvation conditions, meiosis is induced which<br />

leads to sporulation <strong>and</strong> finally to <strong>the</strong> propagation of four haploid spores that segregate 2:2 (cf. chapter<br />

1).<br />

<strong>10</strong>.3.1 Mating Types <strong>and</strong> Mating<br />

<strong>Cell</strong>s of opposite mating type (a <strong>and</strong> α) synchronize each o<strong>the</strong>rs' cell cycles at START in response to<br />

<strong>the</strong>ir mating factors. Once cells progress through START, <strong>the</strong>y are unable to mate until <strong>the</strong> next cell<br />

cycle. Conjugation occurs by surface contact of specialized projections ('schmoo' formation) on mating<br />

cells followed by plasma membrane fusion. The mechanism of <strong>the</strong> mating response will be discussed<br />

in more detail in chapter 13. After spore germination, haploid cells have <strong>the</strong> capability of undergoing<br />

mating type switch which maximizes <strong>the</strong> potential of diploidy. Wild-type strains are homothallic <strong>and</strong> a/α<br />

diploids represent <strong>the</strong> usual vegetative state of this yeast. Industrial (brewing) yeast strains are<br />

generally polyploid <strong>and</strong> do not undergo a sexual life cycle.<br />

<strong>10</strong>.3.2 Meiosis <strong>and</strong> Sporulation<br />

Figure <strong>10</strong>-20: Meiosis in yeast.


Meiosis is a variation on <strong>the</strong> <strong>the</strong>me of mitosis. It produces haploid gametes that contain recombinant<br />

chromosomes, parts of which are derived from one parent <strong>and</strong> parts of which are derived from <strong>the</strong><br />

o<strong>the</strong>r parent. As in <strong>the</strong> case of mitosis, meiosis starts with a round of DNA replication in diploid cells,<br />

which produces connected sister chromatids (Figure <strong>10</strong>-20). Recombination between homologous<br />

chromatids brings homologues toge<strong>the</strong>r.<br />

During <strong>the</strong> first meiotic division (meiosis I), sister kinetochores are treated as a single unit, <strong>and</strong><br />

homologous kinetochore pairs are attached to <strong>and</strong> are pulled towards opposite spindle poles.<br />

However, as long as at least one reciprocal exchange has occurred, cohesion between sister<br />

chromatid arms will oppose disjunction of homologues <strong>and</strong> <strong>the</strong>ir segregation to opposite poles of <strong>the</strong><br />

cell. Thus, loss of sister-chromatid cohesion along chromosome arms is essential for chromosome<br />

segregation during meiosis I. Meanwhile, however, cohesion between sister centromeres persists so<br />

that it can be later used to align sisters on <strong>the</strong> meiosis II metaphase plate. The difference in timing of<br />

sister chromatid cohesion loss is <strong>the</strong>refore a crucial aspect. Similar cohesion proteins as required for<br />

mitosis seem to work in meiosis.<br />

Meiosis produces haploid gametes (spores in yeast) from diploid cells in two stages that in many ways<br />

resemble mitosis [Murakami & Nurse, 2000]. During meiosis, two rounds of chromosome<br />

segregation occur after a single round of DNA replication, producing haploid progeny from diploid<br />

progenitors. Chromosome behaviour during meiosis I, however, is distinct from that in mitosis: (i)<br />

crossovers between maternal <strong>and</strong> paternal sister chromatids (detected cytologically as chiasmata)<br />

bind replicated maternal <strong>and</strong> paternal chromosomes toge<strong>the</strong>r; (ii) sister kinetochores attach to<br />

microtubules from <strong>the</strong> same pole (mono-polar orientation), causing maternal <strong>and</strong> paternal centromere<br />

pairs (<strong>and</strong> not sister chromatids) to be separated <strong>and</strong> segregated to <strong>the</strong> opposite sides of a cell; (iii)<br />

sister chromatid cohesion near centromeres must be protected throughout anaphase when cohesion<br />

along chromosome arms is beginning to be destroyed, a process which is regulated by Polo-like<br />

kinases such as Cdc5p. Moreover, Cdc5 is required both for <strong>the</strong> formation of chiasmata <strong>and</strong> for<br />

cosegregation of sister centromeres at meiosis I [Clyne et al., 2003]. The residual cohesion around<br />

centromeres plays an essential role at <strong>the</strong> second meiotic division, when spindle microtubules from<br />

opposite poles attach to sister chromatids. What protects centromeric cohesin from <strong>the</strong> prophase<br />

pathway? Recent studies have identified novel <strong>and</strong> conserved meiosis-specific kinetochore proteins,<br />

such as monopolin <strong>and</strong> shugoshin, which take this role of guardians [Watanabe, 2004; 2005]. The<br />

yeast kinetochore associated protein, Mam1 (Monopolin), is essential for monopolar attachment [Toth<br />

et al., 2000], while <strong>the</strong> meiosis-specific cohesin, Rec8, is essential for maintaining cohesion between<br />

sister centromeres but not for monopolar attachment. The yeast shugoshin, Sgo1p, prevents removal<br />

of meiotic cohesin complexes from centromeres at <strong>the</strong> first meiotic division, <strong>and</strong> it appears that<br />

shugoshin prevents phosphorylation of cohesin's Scc3-SA2 subunit at centromeres during mitosis.<br />

This ensures that cohesin persists at centromeres until activation of separase causes cleavage of its<br />

alpha kleisin subunit [Rabitsch et al., 2003; Katis et al., 2004a; 2004b; McGuinness et al., 2005]. In<br />

yeast, shugoshin also has a crucial role in sensing <strong>the</strong> loss of tension at kinetochores <strong>and</strong> in<br />

generating <strong>the</strong> spindle checkpoint signal [Watanabe, 2005].<br />

Sporulation, which in yeast is induced at (nitrogen) starvation, is regulated by a specialized MAP<br />

kinase signalling pathway (see chapter 13). In many aspects, starvation is similar to o<strong>the</strong>r stress<br />

responses in yeast. Spore formation requires <strong>the</strong> activation of a large number of genes, several of<br />

which are involved in <strong>the</strong> biosyn<strong>the</strong>sis of spore walls.


References<br />

Amon, A. (1999) The spindle checkpoint. Curr. Opin. Genet.Dev. 9, 69-75.<br />

Anderson, D.E., Losada, A., Erickson, H.P., Hirano, T. (2002) Condensin <strong>and</strong> cohesin display different arm<br />

conformations with characteristic hinge angles. J. <strong>Cell</strong> Biol. 156, 419-424.<br />

Arumugam, P., Nishino, T., Haering, C.H., Gruber, S., Nasmyth, K. (2006) Cohesin's ATPase activity is stimulated<br />

by <strong>the</strong> C-terminal Winged-Helix domain of its kleisin subunit. Curr Biol. 16, 1998-2008.<br />

Barral, Y., Mermall, V., Mooseker, M.S., Snyder, M. Compartmentalization of <strong>the</strong> cell cortex by septins is required<br />

for maintenance of cell polarity in yeast. Mol. <strong>Cell</strong>. 5 (2000) 841-51<br />

Barral, Y., Parra, M., Bidlingmaier, S., Snyder, M. (1999) Nim1-related kinases coordinate cell cycle progression<br />

with <strong>the</strong> organization of <strong>the</strong> peripheral cytoskeleton in yeast. Genes Dev. 13, 176–187.<br />

Barrett, J.G., Manning, B.D., Snyder, M. (2000) The Kar3p kinesin-related protein forms a novel heterodimeric<br />

structure with its associated protein Cik1p. Mol Biol <strong>Cell</strong>. 11, 2373-2385.<br />

Beach, D., Durkacz, B. et al. (1982) Functionally homologous cell cycle control genes in budding <strong>and</strong> fission<br />

yeast. Nature 300, 706–709.<br />

Bell, S.P., Mitchell, J., Leber, J., Kobayashi, R., Stillman, B. (1995) The multidomain structure of Orc1p reveals<br />

similarity to regulators of DNA replication <strong>and</strong> transcriptional silencing. <strong>Cell</strong> 83, 563-568.<br />

Bidlingmaier, S., Weiss, E.L., Seidel, C., Drubin, D.G., Snyder, M. (2001) The Cbk1p pathway is important for<br />

polarized cell growth <strong>and</strong> cell separation in Saccharomyces cerevisiae. Mol. <strong>Cell</strong>. Biol. 21, 2449-2462.<br />

Biggins, S., Murray, A.W. Sister chromatid cohesion in mitosis. Curr. Opin. Genet. Dev. 9 (1999) 230-236.<br />

Blackburn, E.H., Greider, C.W., Szostak, J.W. (2006) Telomeres <strong>and</strong> telomerase: <strong>the</strong> path from maize,<br />

Tetrahymena <strong>and</strong> yeast to human cancer <strong>and</strong> aging. Nat Med. 12, 1133-1138.<br />

Boule, J.P., Zakian, V.A. (2006) Roles of Pif1-like helicases in <strong>the</strong> maintenance of genomic stability. Nucleic Acids<br />

Res. 34, 4147–4153.<br />

Bouquin, N., Barral, Y., Courbeyrette, R., Blondel, M., Snyder, M., Mann, C. Regulation of cytokinesis by <strong>the</strong> Elm1<br />

protein kinase in Saccharomyces cerevisiae. J. <strong>Cell</strong> Sci. 113 (2000) 1435-1445.<br />

Brigati, C., Kurtz, S., Balderes, D., Vidali, G., Shore, D. (1993) An essential yeast gene encoding a TTAGGG<br />

repeat-binding protein. Mol <strong>Cell</strong> Biol. 13, 1306-1314.<br />

Buvelot, S., Tatsutani, S.Y., Vermaak, D., Biggins, S. (2003) The budding yeast Ipl1/Aurora protein kinase<br />

regulates mitotic spindle disassembly. J. <strong>Cell</strong> Biol. 160, 329-339.<br />

Camobel, J.L., Cohen-Fix,O. Chromosome cohesion: ring around <strong>the</strong> sisters? Trends Biochem. Sci. 27 (2002)<br />

492-495.<br />

Casamayor, A., Snyder, M. (2003) Molecular Dissection of a <strong>Yeast</strong> Septin: Distinct Domains Are Required for<br />

Septin Interaction, Localization, <strong>and</strong> Function. Mol. <strong>Cell</strong> Biol. 23, 2762–2777.<br />

Cheeseman, I.M., Desai, A. (2004) <strong>Cell</strong> division: AAAtacking <strong>the</strong> mitotic spindle. Curr Biol. 14, R70-72. Review.<br />

Ciosk, R., Zachariae, W., Michaelis, C., Shevchenko, A., Mann, M., Nasmyth, K. (1998) An ESP1/PDS1 complex<br />

regulates loss of sister chromatid cohesion at <strong>the</strong> metaphase to anaphase transition in yeast. <strong>Cell</strong> 93, <strong>10</strong>67–<br />

<strong>10</strong>76.<br />

Clyne, R.K., Katis, V.L., Jessop, L., Benjamin, K.R., Herskowitz, I., Lichten, M., Nasmyth, K. (2003) Polo-like<br />

kinase Cdc5 promotes chiasmata formation <strong>and</strong> cosegregation of sister centromeres at meiosis I. Nat <strong>Cell</strong><br />

Biol. 5, 480-485.<br />

Coelho, P.S., Kumar, A., Snyder, M. Genome-wide mutant collections: toolboxes for functional genomics. Curr.<br />

Opin. Microbiol. 3 (2000) 309-315.<br />

Cohen-Fix, O., Koshl<strong>and</strong>, D. (1997) The anaphase inhibitor of Saccharomyces cerevisiae Pds1p is a target of <strong>the</strong><br />

DNA damage checkpoint pathway. Proc Natl Acad Sci U S A 94, 14361-1436.<br />

Cohen-Fix, O., Peters, J.M., Kirschner, M.W., Koshl<strong>and</strong>, D. (1996) Anaphase initiation in Saccharomyces<br />

cerevisiae is controlled by <strong>the</strong> APC-dependent degradation of <strong>the</strong> anaphase inhibitor Pds1p. Genes Dev <strong>10</strong>,<br />

3081-3093.<br />

Cohn, M., Blackburn, E.H. (1995) Telomerase in yeast. Science 269, 396-400.<br />

Cross, R.A. Molecular motors: dyneins gearbox. Cur. Biol. 14 (2004) R355-356.<br />

Dewar, H., Tanaka, K., Nasmyth, K., Tanaka, T.U. (2004) Tension between two kinetochores suffices for <strong>the</strong>ir biorientation<br />

on <strong>the</strong> mitotic spindle. Nature 428, 93-97.<br />

Diffley, J.F., Cocker, J.H. (1992) Protein-DNA interactions at a yeast replication origin. Nature 357, 169-172.<br />

Dirick, L., Bohm, T., Nasmyth, K. (1995) Roles <strong>and</strong> regulation of Cln-Cdc28 kinases at <strong>the</strong> start of <strong>the</strong> cell cycle of<br />

Saccharomyces cerevisiae. EMBO J.14, 4803-4813.<br />

Donaldson, A.D., Blow, J.J. The regulation of replication origin activation. Curr. Opin. Genet. Dev. 9 (1999) 62-<br />

68.<br />

Du, Y.C., Stillman, B. (2002) Yph1p, an ORC-interacting protein: potential links between cell proliferation control,<br />

DNA replication, <strong>and</strong> ribosome biogenesis. <strong>Cell</strong> <strong>10</strong>9, 835-848.<br />

Evans, T., Hunt, T., Youngblom, J. (1982) On <strong>the</strong> role of maternal mRNA in sea urchins: studies of a protein<br />

which appears to be destroyed at a particular point in each cell division cycle. Biol. Bull. 163, 372.<br />

Evans, T., Rosenthal, E.T., Youngbloom, J., Distel, D., Hunt, T. (1983). Cyclin: a protein specified by maternal<br />

mRNA in sea urchin eggs that is destroyed at each cleavage division. <strong>Cell</strong> 33, 389–396.<br />

Fisher, T.S., Taggart, A.K.P. , Zakian,V.A. (2004) <strong>Cell</strong> cycle-dependent regulation of yeast telomerase by Ku. Nat.<br />

Struct. Mol. Biol., 11, 1198–1205.<br />

Fleming, W. (1879) Archiv f. mikroskop. Anatomie 18, 151-259. Koshl<strong>and</strong>, D., Hartwell, L.H.(1987) The structure<br />

of sister minichromosome DNA before anaphase in Saccharomyces cerevisiae. Science 238, 1713-1736.


Foss, M., McNally, F.J., Laurenson, P., Rine, J. (1993) Origin recognition complex (ORC) in transcriptional<br />

silencing <strong>and</strong> DNA replication in S. cerevisiae. Science 262, 1838-1844.<br />

Futcher, B. Microarrays <strong>and</strong> cell cycle transcription in yeast. Cur. Opin. <strong>Cell</strong> Biol. 12 (2000) 7<strong>10</strong>–715.<br />

Gillett, E. S., Espelin, C. W., Sorger, P. K. (2004) Spindle checkpoint proteins <strong>and</strong> chromosome-microtubule<br />

attachment in budding yeast. J. <strong>Cell</strong> Biol. 164, 535–546.<br />

Gilson, E., Geli, V. (2007) How telomeres are replicated. Nat Rev Mol <strong>Cell</strong> Biol 8, 825-838.<br />

Glotzer, M, Murray, A.W., Kirschner, M.W. (1991). Cyclin is degraded by <strong>the</strong> ubiquitin pathway. Nature 349: 132–<br />

138.<br />

Gould, K. L., Nurse, P. (1989) Tyrosine phosphorylation of <strong>the</strong> fission yeast cdc2+ protein kinase regulates entry<br />

into mitosis. Nature 342, 39–45.<br />

Gruber, S., Arumugam, P., Katou, Y., Kuglitsch, D., Helmhart, W., Shirahige, K., Nasmyth, K. (2006) Evidence<br />

that loading of cohesin onto chromosomes involves opening of its SMC hinge. <strong>Cell</strong> 127, 523-537.<br />

Guacci, V., Koshl<strong>and</strong>, D. Strunnikov, A. (1997) A direct link between sister chromatid cohesion <strong>and</strong> chromosome<br />

condensation revealed through <strong>the</strong> analysis of MCD1 in S. cerevisiae. <strong>Cell</strong> 91, 47-57.<br />

Haering, C.H., Lowe, J., Hochwagen, A., Nasmyth, K. (2002) Molecular architecture of SMC proteins <strong>and</strong> <strong>the</strong><br />

yeast cohesin complex. Mol. <strong>Cell</strong> 9, 773-788.<br />

Hardy, C.F., Sussel, L., Shore, D. (1992) A RAP1-interacting protein involved in transcriptional silencing <strong>and</strong><br />

telomere length regulation. Genes Dev. 6, 801-814.<br />

Hartwell, L. H. (1971b) Genetic control of <strong>the</strong> cell division cycle in yeast. IV. Genes controlling bud emergence<br />

<strong>and</strong> cytokinesis. Exp. <strong>Cell</strong> Res. 69, 265–276.<br />

Hartwell, L. H., Hutchison, H. T., Holl<strong>and</strong>, T. M., McLaughlin, C. S. (1970a) The effect of cycloheximide upon<br />

polyribosome stability in two yeast mutants defective respectively in <strong>the</strong> initiation of polypeptide chains <strong>and</strong><br />

in messenger RNA syn<strong>the</strong>sis. Mol. Gen. Genet. <strong>10</strong>6, 347-361.<br />

Hartwell, L. H., Mortimer, R.K., Culotti, J., Culotti, M. (1974) Genetic control of <strong>the</strong> cell division cycle in yeast. V.<br />

Genetic analysis of cdc mutants. Genetics 74, 267–286.<br />

Hartwell, L., Weinert, T. (1989) Checkpoints: controls that ensure <strong>the</strong> order of cell cycle events. Science 246,<br />

629–634.<br />

Hartwell, L.H. (1967) Macromolecule syn<strong>the</strong>sis in temperature-sensitive mutants of yeast. J Bacteriol. 93, 1662-<br />

1670.<br />

Hartwell, L.H. (1971a) Genetic control of <strong>the</strong> cell division cycle in yeast. II. Genes controlling DNA replication <strong>and</strong><br />

its initiation. J Mol Biol. 59, 183-194.<br />

Hartwell, L.H. (1973) Synthronization of haploid yeast cell cycles, a prelude to conjugation. Exp. <strong>Cell</strong> Res.<br />

76,111–117.<br />

Hartwell, L.H. (1974) Saccharomyces cerevisiae cell cycle. Bacteriol Rev. 38, 164-198. Review.<br />

Hartwell, L.H. (2002) Nobel Lecture. <strong>Yeast</strong> <strong>and</strong> cancer. Biosci Rep. 22, 373-394.<br />

Heun, P., Laroche, T., Shimada, K., Furrer, P., Gasser, S.M. Chromosome dynamics in <strong>the</strong> yeast interphase<br />

nucleus. Science 294 (2001) 2181-2186.<br />

Hirano, T (2000) Chromosome cohesion, condensation, <strong>and</strong> separation. Ann. Rev. Biochem. 69, 115 -144.<br />

Holloway, S.L., Glotzer, M., King, R.W., Murray, A.W. (1993) Anaphase is initiated by proteolysis ra<strong>the</strong>r than by<br />

<strong>the</strong> inactivation of maturation-promoting factor. <strong>Cell</strong> 73, 1393-1402.<br />

Hunt, T. (2001) Protein Syn<strong>the</strong>sis, Proteolysis <strong>and</strong> <strong>Cell</strong> <strong>Cycle</strong> Transitions. Nobel Lecture, 2001.<br />

Hwang, L. H., Lau, L. F., Smith, D. L., Mistrot, C. A., Hardwick, K. G., Hwang, E. S., Amon, A., Murray, A. W.<br />

(1998) Budding yeast Cdc20: a target of <strong>the</strong> spindle checkpoint. Science 279, <strong>10</strong>41–<strong>10</strong>44.<br />

Irniger, S., Piatti, S., Michaelis, C., Nasmyth, K. (1995) Genes involved in sister chromatid separation are needed<br />

for B-type cyclin proteolysis in budding yeast. <strong>Cell</strong> 81, 269-278.<br />

Ivanov, D., Nasmyth, K. (2005) A topological interaction between cohesin rings <strong>and</strong> a circular minichromosome.<br />

<strong>Cell</strong> 122, 849-860.<br />

Iyer, V.R., Horak, C.E., Scafe, C.S., Botstein, D., Snyder, M., Brown, P.O. Genomic binding sites of <strong>the</strong> yeast cellcycle<br />

transcription factors SBF <strong>and</strong> MBF. Nature 409 (2001) 533-538.<br />

Jaspersen, S.L., Winey, M. (2004) The budding yeast spindle pole body: Structure, Duplication, <strong>and</strong> Function.<br />

Annu. Rev. <strong>Cell</strong> Dev. Biol. 20, 1–28.<br />

Katis V.L., Matos, J., Mori, S., Shirahige, K., Zachariae, W., Nasmyth, K. (2004a) Spo13 facilitates monopolin<br />

recruitment to kinetochores <strong>and</strong> regulates maintenance of centromeric cohesion during yeast meiosis. Curr<br />

Biol. 14, 2183-2196.<br />

Katis, V.L., Galova, M., Rabitsch, K.P., Gregan, J., Nasmyth, K. (2004b) Maintenance of cohesin at centromeres<br />

after meiosis I in budding yeast requires a kinetochore-associated protein related to MEI-S332. Curr Biol. 14,<br />

560-572.<br />

King, R.W. Peters, J.M., Tugendreich, S., Rolfe, M., Hieter, P., Kirschner, M.W. (1995) A 20S complex containing<br />

CDC27 <strong>and</strong> CDC16 catalyzes <strong>the</strong> mitosis-specific conjugation of ubiquitin to cyclin B. <strong>Cell</strong> 81, 269-278.<br />

Kurtz, S., Shore, D. (1991) RAP1 protein activates <strong>and</strong> silences transcription of mating-type genes in yeast.<br />

Genes Dev. 5, 616-628.<br />

Lauren M., DeSalle, L.M., Pagano, M. Regulation of <strong>the</strong> G1 to S transition by <strong>the</strong> ubiquitin pathway. FEBS Lett.<br />

490 (2001) 179-189.<br />

Lavoie, B.D., Hogan, E., Koshl<strong>and</strong>, D. (2002) In vivo dissection of <strong>the</strong> chromosome condensation machinery:<br />

reversibility of condensation distinguishes contributions of condensin <strong>and</strong> cohesin. J <strong>Cell</strong> Biol. 156, 805-815.<br />

Lee, M. G., Nurse, P. (1987) Complementation used to clone a human homologue of <strong>the</strong> fission yeast cell cycle<br />

control gene cdc2. Nature 327, 31–35.<br />

Lew, D. J., Burke, D. J. (2003) The spindle assembly <strong>and</strong> spindle position checkpoints. Annu. Rev. Genet. 37,<br />

251–282.


Lew, D.J. (2003) The morphogenesis checkpoint: how yeast cells watch <strong>the</strong>ir figures. Curr. Opin. <strong>Cell</strong> Biol. 15,<br />

648–653.<br />

Li, R., Yu, D.S., Tanaka, M., Zheng, L., Berger, S.L., Stillman, B. (1998) Activation of chromosomal DNA<br />

replication in Saccharomyces cerevisiae by acidic transcriptional activation domains. Mol <strong>Cell</strong> Biol. 18, 1296-<br />

1302.<br />

Liang, C., Weinreich, M., Stillman, B. (1995) ORC <strong>and</strong> Cdc6p interact <strong>and</strong> determine <strong>the</strong> frequency of initiation of<br />

DNA replication in <strong>the</strong> genome. <strong>Cell</strong> 81, 667-676.<br />

Loo, S., Fox, C.A., Rine, J., Kobayashi, R., Stillman, B., Bell, S. (1995) The origin recognition complex in<br />

silencing, cell cycle progression, <strong>and</strong> DNA replication. Mol Biol <strong>Cell</strong> 6, 741-756.<br />

Losada, A., Hirano, M., Hirano, T. (1998) Identification of Xenopus SMC protein complexes required for sister<br />

chromatid cohesion. Genes Dev. 12, 1986-1997.<br />

Lundblad, V. (2003) Telomere replication: an Est fest. Curr. Biol., 13, R439–R441.<br />

Lustig, A.J., Kurtz, S., Shore, D. (1990) Involvement of <strong>the</strong> silencer <strong>and</strong> UAS binding protein RAP1 in regulation of<br />

telomere length. Science 250, 549-553.<br />

Mala, J., Nurse, P. Trends <strong>Cell</strong> Biol. 8 (1998) 163-169.<br />

Manning, B.D., Barrett, J.G., Wallace, J.A., Granok, H., Snyder, M. Differential regulation of <strong>the</strong> Kar3p kinesinrelated<br />

protein by two associated proteins, Cik1p <strong>and</strong> Vik1p. J. <strong>Cell</strong> Biol. 22 (1999) 1219-1233.<br />

Manning, B.D., Barrett, J.G., Wallace, J.A., Granok, H., Snyder, M. (1999) Differential regulation of <strong>the</strong> Kar3p<br />

kinesin-related protein by two associated proteins, Cik1p <strong>and</strong> Vik1p. J. <strong>Cell</strong> Biol. 22 , 1219-1233.<br />

Manning, B.D., Snyder, M. (2000) Drivers <strong>and</strong> passengers wanted! <strong>the</strong> role of kinesin-associated proteins. Trends<br />

<strong>Cell</strong>.Biol. <strong>10</strong>, 281-289.<br />

Manning, B.D., Snyder, M. Drivers <strong>and</strong> passengers wanted! <strong>the</strong> role of kinesin-associated proteins. Trends<br />

<strong>Cell</strong>.Biol. <strong>10</strong> (2000) 281-289.<br />

McGuinness, B.E., Hirota, T., Kudo, N.R., Peters, J.M., Nasmyth, K. (2005) Shugoshin prevents dissociation of<br />

cohesin from centromeres during mitosis in vertebrate cells. PLoS Biol. 3, e86.<br />

McLaughlin, C.S., Hartwell, L.H. (1969) Mutants of yeast defective in <strong>the</strong> initiation of protein biosyn<strong>the</strong>sis. Cold<br />

Spring Harb Symp Quant Biol. 34, 317-319.<br />

Michaelis, C., Ciosk, R., Nasmyth, K. (1997) Cohesins: chromosomal proteins that prevent premature separation<br />

of sister chromatids. <strong>Cell</strong> 91, 35-45.<br />

Micklem, G., Rowley, A., Harwood, J., Nasmyth, K., Diffley, J.F. (1993) <strong>Yeast</strong> origin recognition complex is<br />

involved in DNA replication <strong>and</strong> transcriptional silencing. Nature 366, 87-89.<br />

Mitchison, J.M. (1971) The biology of <strong>the</strong> cell cycle. Cambridge: Cambridge University Press.<br />

Moreno, S., Hayles, J., Nurse, P. (1989) Regulation of p34cdc2 protein kinase during mitosis. <strong>Cell</strong> 58, 361–372.<br />

Moretti, P., Freeman, K., Coodly, L., Shore, D. (1994) Evidence that a complex of SIR proteins interacts with <strong>the</strong><br />

silencer <strong>and</strong> telomere-binding protein RAP1. Genes Dev. 8, 2257-2269.<br />

Murakami, H., Nurse, P. (2000) DNA replication <strong>and</strong> damage checkpoints <strong>and</strong> meiotic cell cycle controls in <strong>the</strong><br />

fission <strong>and</strong> budding yeasts. Biochem. J. 349, 1-12. Review.<br />

Nasmyth, K. (2001) Disseminating <strong>the</strong> genome: joining, resolving, <strong>and</strong> separating sister chromatids during mitosis<br />

<strong>and</strong> meiosis. Ann. Rev. Genet. 35, 673 – 745.<br />

Nasmyth, K. Segregating sister genomes: <strong>the</strong> molecular biology of chromosome separation. Science 297 (2002)<br />

559-565.<br />

Nasmyth, K. Separating sister chromatids. Trends Biochem. Sci. 24 (1999) 98-<strong>10</strong>4.<br />

Nasmyth, K.A., Reed, S,I. (1980) Isolation of genes by complementation in yeast: molecular cloning of a cell-cycle<br />

gene. Proc Natl Acad Sci U S A 77, 2119-2123.<br />

Ni, L., Snyder, M. (2001) A genomic study of <strong>the</strong> bipolar bud site selection pattern in Saccharomyces cerevisiae.<br />

Mol. Biol. <strong>Cell</strong> 12, 2147-2170.<br />

Nurse, P. (1990) Universal control mechanism regulating onset of M-phase. Nature 344, 503–508.<br />

Nurse, P. (2000) A long twentieth century of <strong>the</strong> cell cycle <strong>and</strong> beyond. <strong>Cell</strong> <strong>10</strong>0, 71–78.<br />

Nurse, P., Bissett, Y. (1981) Gene required in G1 for commitment to cell cycle <strong>and</strong> in G2 for control of mitosis in<br />

fission yeast. Nature 292, 558–560.<br />

Nurse, P., Thuriaux, P. (1980) Regulatory genes controlling mitosis in <strong>the</strong> fission yeast Schizosaccharomyces<br />

pombe. Genetics 96, 627–637.<br />

Nurse, P., Thuriaux, P., Nasmyth, K. (1976) Genetic control of <strong>the</strong> cell division cycle in <strong>the</strong> fission yeast S. pombe.<br />

Mol. Gen. Genet. 146, 167-178.<br />

Nurse, P.M. (2001) Cyclin dependent kinases <strong>and</strong> cell cycle control. Nobel Lecture, December 9, 2001<br />

Ooi, S.L., Shoemaker, D.D., Boeke, J.D. A DNA microarray-based genetic screen for non-homologous endjoining<br />

mutants in Saccharomyces cerevisiae. Science 294 (2001) 2552-2556.<br />

Paulovich, A.G., Toczyski, D.P., Hartwell, L.H. (1997) When checkpoints fail. <strong>Cell</strong> 88, 315-321.<br />

Pereira, G., Schiebel, E. (2003) Separase regulates INCENPAurora B anaphase spindle function through Cdc14.<br />

Science 302, 2120-2124.<br />

Peters, J. M. (2002) The anaphase-promoting complex: proteolysis in mitosis <strong>and</strong> beyond. Mol. <strong>Cell</strong> 9, 931–943.<br />

Rabitsch, K.P., Petronczki, M., Javerzat, J.P., Genier, S., Chwalla, B., Schleiffer, A., Tanaka, T.U., Nasmyth, K.<br />

(2003) Kinetochore recruitment of two nucleolar proteins is required for homolog segregation in meiosis I.<br />

Dev <strong>Cell</strong>. 4, 535-548.<br />

Raghuraman, M.K., Winzeler, E.A., Collingwood, D., Hunt, S., Wodicka, L., Conway, A., Lockhart, D.J., Davis,<br />

R.W., Brewer, B.J., Fangman, W.L. (2001) Replication dynamics of <strong>the</strong> yeast genome. Science 294, 115-<br />

121.<br />

Rao, H., Stillman, B. (1995) The origin recognition complex interacts with a bipartite DNA binding site within yeast<br />

replicators. Proc Natl Acad Sci U S A. 92, 2224-2228.


Rao, H., Uhlmann, F., Nasmyth, K,, Varshavsky, A. (2001) Degradation of a cohesin subunit by <strong>the</strong> N-end rule<br />

pathway is essential for chromosome stability. Nature 4<strong>10</strong>, 955-959.<br />

Reid, R.J., Rothstein, R. (2004) Stay close to your sister. Mol <strong>Cell</strong>. 14, 418-420. Review<br />

Ricchetti, M., Fairhead, C., Dujon, B. Mitochondrial DNA repairs double-str<strong>and</strong> breaks in yeast chromosomes.<br />

Nature 402 (1999) 96-<strong>10</strong>0.<br />

Roemer, T., Vallier, L.G., Snyder, M. (1996) Selection of polarized growth sites in yeast. Trends <strong>Cell</strong> Biol. 6, 434-<br />

441.<br />

Schuyler, S.C., Liu, J.Y., Pellman, D. (2003) The molecular function of Ase1p: evidence for a MAP-dependent<br />

midzone-specific spindle matrix. Microtubule-associated proteins. J. <strong>Cell</strong> Biol. 160, 517-528.<br />

Segal, M., Bloom, K. (2001) Control of spindle polarity <strong>and</strong> orientation in S. cerevisiae. Trends <strong>Cell</strong> Biol. 11, 160-<br />

166.<br />

Sheu, Y.J., Barral, Y., Snyder, M.(2000) Polarized growth controls cell shape <strong>and</strong> bipolar bud site selection in<br />

Saccharomyces cerevisiae. Mol. <strong>Cell</strong>. Biol. 20, 5235-5247.<br />

Shimoda, C. (2004) Forespore membrane assembly in yeast: coordinating SPBs <strong>and</strong> membrane trafficking. J.<br />

<strong>Cell</strong> Sci. 17, 389-395.<br />

Shintomi, K., Hirano, T. (2007) How are cohesin rings opened <strong>and</strong> closed? Trends Biochem Sci. 32, 154-157.<br />

Shore, D. (1997) Telomerase <strong>and</strong> telomere-binding proteins: controlling <strong>the</strong> endgame. Trends Biochem Sci. 22,<br />

233-235. Review.<br />

Shore, D. (1998) <strong>Cell</strong>ular senescence: lessons from yeast for human aging? Curr Biol. 8, R192-195. Review.<br />

Shore, D. (2001) Telomeric chromatin: replicating <strong>and</strong> wrapping up chromosome ends. Curr Opin Genet Dev. 11,<br />

189-198. Review.<br />

Shore, D., Nasmyth, K. (1987) Purification <strong>and</strong> cloning of a DNA binding protein from yeast that binds to both<br />

silencer <strong>and</strong> activator elements. <strong>Cell</strong> 51, 721-732.<br />

Shore, D., Stillman, D.J., Br<strong>and</strong>, A.H., Nasmyth, K.A. (1987) Identification of silencer binding proteins from yeast:<br />

possible roles in SIR control <strong>and</strong> DNA replication. EMBO J. 6, 461-467.<br />

Skibbens, R.V., Corson, L.B., Koshl<strong>and</strong>, D., Hieter, P. (1999) Ctf7p is essential for sister chromatid cohesion <strong>and</strong><br />

links mitotic chromosome structure to <strong>the</strong> DNA replication machinery. Genes Dev. 13, 307-319.<br />

Smeal, T., Guarente, L. (1997) Mechanisms of cellular senescence. Curr Opin Genet Dev. 7, 281-287. Review.<br />

Stellwagen, A.E., Haimberger, Z.W., Veatch, J.R., Gottschling, D.E. (2003) Ku interacts with telomerase RNA to<br />

promote telomere addition at native <strong>and</strong> broken chromosome ends. Genes Dev., 17, 2384–2395.<br />

Stern, B., Nurse, P. (1996) A quantitative model for <strong>the</strong> cdc2 control of S phase <strong>and</strong> mitosis in fission yeast.<br />

Trends Genet 12, 345–350.<br />

Stillman, B. (2001) DNA replication. Genomic views of genome duplication. Science 294, 2301-2304.<br />

Stillman, B. (2005) Origin recognition <strong>and</strong> <strong>the</strong> chromosome cycle. FEBS Lett. 579, 877-884. Review.<br />

Stillman, B. DNA replication. Genomic views of genome duplication. Science 294 (2001) 2301-2304.<br />

Sudakin, V., Ganoth, D., Dahan, A., Heller, H., Hershko, J., Luca, F.C., Ruderman, J.V., Hershko, A. (1995) The<br />

cyclosome, a large complex containing cyclin-selective ubiquitin ligase activity, targets cyclins for destruction<br />

at <strong>the</strong> end of mitosis. Mol. Biol. <strong>Cell</strong> 6, 185-198.<br />

Surana, U., Amon, A., Dowzer, C., McGrew, J., Byers, B., Nasmyth, K. (1993) Destruction of <strong>the</strong> CDC28/CLB<br />

mitotic kinase is not required for <strong>the</strong> metaphase to anaphase transition in budding yeast. EMBO J. 12, 1969-<br />

1978.<br />

Swaroop, M., Wang, Y., Miller, P., Duan, H., Jatkoe, T., Madore, S.J., Sun, Y. <strong>Yeast</strong> homolog of human<br />

SAG/ROC2/Rbx2/Hrt2 is essential for cell growth, but not for germination: chip profiling implicates its role in<br />

cell cycle regulation. Oncogene 19 (2000) 2855-2866.<br />

Swaroop, M., Wang, Y., Miller, P., Duan, H., Jatkoe, T., Madore, S.J., Sun, Y. (2000) <strong>Yeast</strong> homolog of human<br />

SAG/ROC2/Rbx2/Hrt2 is essential for cell growth, but not for germination: chip profiling implicates its role in<br />

cell cycle regulation. Oncogene 19, 2855-2866.<br />

Taddei, A., Gasser, S.M. (2004) Multiple pathways for telomere te<strong>the</strong>ring: functional implications of subnuclear<br />

position for heterochromatin formation. Biochim. Biophys. Acta 1677, 120–128. Review.<br />

Taggart, A.K., Zakian, V.A. (2003) Telomerase: What are <strong>the</strong> EST proteins doing? Curr. Opin. <strong>Cell</strong> Biol., 15, 275–<br />

280.<br />

Taggart, A.K.P., Teng, S.C., Zakian, V.A. (2002) Est1p as a cell cycle-regulated activator of telomere-bound<br />

telomerase. Science, 297, <strong>10</strong>23–<strong>10</strong>26.<br />

Tan, A.L.C., Rida, P.G.C., Surana, U. (2005) Essential tension <strong>and</strong> constructive destruction: <strong>the</strong> spindle<br />

checkpoint <strong>and</strong> its regulatory links with mitotic exit. Biochem. J. 386, 1–13. Review.<br />

Teixeira, M.T.,Gilson, E. (2005) Telomere maintenance, function <strong>and</strong> evolution: <strong>the</strong> yeast paradigm. Chromo Res<br />

13, 535-548.<br />

Torres-Rosell, J., Machin, F., Aragon, L. (2005) Smc5-Smc6 complex preserves nucleolar integrity in S.<br />

cerevisiae. <strong>Cell</strong> <strong>Cycle</strong> 4, 868-872.<br />

Toth, A., Ciosk, R., Uhlmann, F., Galova, M., Schleiffer, A., Nasmyth, K. (1999) <strong>Yeast</strong> cohesin complex requires a<br />

conserved protein, Eco1p(Ctf7), to establish cohesion between sister chromatids during DNA replication.<br />

Genes Dev. 13, 320-333.<br />

Tóth, A., Rabitsch, K.P., Gálová, M., Schleiffer, A., Buonomo, S.B., Nasmyth, K. (2000) Functional genomics<br />

identifies monopolin: a kinetochore protein required for segregation of homologs during meiosis i. <strong>Cell</strong> <strong>10</strong>3,<br />

1155-1168.<br />

Uhlmann, F. The mechanism of sister chromatid cohesion. Exp. <strong>Cell</strong> Res. 296 (2004) 80-85.<br />

Uhlmann, F., Lottspeich, F., Nasmyth, K. (1999) Sister-chromatid separation at anaphase onset is promoted by<br />

cleavage of <strong>the</strong> cohesin subunit Scc1. Nature 400, 37–42.


Vogel, J., Drapkin, B., Oomen, J., Beach, D., Bloom, K., Snyder, M. Phosphorylation of gamma-tubulin regulates<br />

microtubule organization in budding yeast. Dev. <strong>Cell</strong>. 1 (2001) 621-631.<br />

Vogel, J., Snyder, M. (2000) Gamma-tubulin of budding yeast. Curr. Top. Dev. Biol. 49, 75-<strong>10</strong>4.<br />

Wang, B.D., Eyre, D., Basrai, M., Lichten, M., Strunnikov, A. (2005) Condensin binding at distinct <strong>and</strong> specific<br />

chromosomal sites in <strong>the</strong> Saccharomyces cerevisiae genome. Mol <strong>Cell</strong> Biol. 25, 7216-7225.<br />

Watanabe, Y. (2004) Modifying sister chromatid cohesion for meiosis. J <strong>Cell</strong> Sci. 117, 4017-4023.<br />

Watanabe, Y. (2005) Shugoshin: guardian spirit at <strong>the</strong> centromere. Curr Opin <strong>Cell</strong> Biol.17, 590-595.<br />

Weinreich, M., Liang, C., Chen, H.H., Stillman, B. (2001) Binding of cyclin-dependent kinases to ORC <strong>and</strong> Cdc6p<br />

regulates <strong>the</strong> chromosome replication cycle. Proc Natl Acad Sci U S A. 98, 11211-11217.<br />

Weinreich, M., Liang, C., Stillman, B. (1999) The Cdc6p nucleotide-binding motif is required for loading mcm<br />

proteins onto chromatin. Proc Natl Acad Sci U S A. 96, 441-446.<br />

Weitzer, S., Lehane, C., Uhlmann, F. (2003) A model for ATP hydrolysis-dependent binding of cohesin to DNA.<br />

Curr Biol. 13, 1930-1940.<br />

Williams, R.S., Shohet, R.V., Stillman, B. (1997) A human protein related to yeast Cdc6p. Proc Natl Acad Sci U S<br />

A. 94, 142-147.<br />

Wood, J.S., Hartwell, L.H. (1982) A dependent pathway of gene functions leading to chromosome segregation in<br />

Saccharomyces cerevisiae. J <strong>Cell</strong> Biol. 94, 718-726.<br />

Wotton, D., Shore, D. (1997) A novel Rap1p-interacting factor, Rif2p, cooperates with Rif1p to regulate telomere<br />

length in Saccharomyces cerevisiae. Genes Dev. 11, 748-760.<br />

Wu, M., Gerhart, J.C. (1980). Partial purification <strong>and</strong> characterization of <strong>the</strong> maturation-promoting factor from<br />

eggs of Xenopus laevis. Dev. Biol. 1980 79: 465–477.<br />

Wuarin, J., Nurse, P. (1996) Regulating S phase: CDKs, Licensing <strong>and</strong> proteolysis. <strong>Cell</strong> 85: 785–787.<br />

Wyrick, J.J., Aparicio, J.Gg., Chen, T., Barnett, J.D., Jennings, E.G., Young, R.A., Bell, S.P., Aparicio, O.M.<br />

Genomewide distribution of ORC <strong>and</strong> MCM proteins in S. cerevisiae: highresolution mapping of replication<br />

origins. Science 294 (2001) 2357-2360.<br />

Yamamoto, A., Guacci, V., Koshl<strong>and</strong>, D. (1996a) Pds1p is required for faithful execution of anaphase in <strong>the</strong> yeast,<br />

Saccharomyces cerevisiae. J <strong>Cell</strong> Biol 133, 85-97.<br />

Yamamoto, A., Guacci, V., Koshl<strong>and</strong>, D. (1996b) Pds1p, an inhibitor of anaphase in budding yeast, plays a critical<br />

role in <strong>the</strong> APC <strong>and</strong> checkpoint pathway(s). J <strong>Cell</strong> Biol 133, 99-1<strong>10</strong>.<br />

Yu, H.G., Koshl<strong>and</strong>, D. (2005) Chromosome morphogenesis: condensin-dependent cohesin removal during<br />

meiosis. <strong>Cell</strong> 123, 397-407.<br />

Zachariae, W., Nasmyth, K. (1999) Whose end is destruction: cell division <strong>and</strong> <strong>the</strong> anaphase-promoting complex.<br />

Genes Dev. 13, 2039-2058.<br />

Zachariae, W., Schwab, M., Nasmyth, K., Seufert, W. (1998) Control of cyclin ubiquitination by CDK-regulated<br />

binding of Hct1 to <strong>the</strong> anaphase promoting complex. Science 282, 1721-1724.<br />

Zachariae, W., Shin, T.H., Galova, M., Obermaier, B., Nasmyth, K. (1996) Identification of subunits of <strong>the</strong><br />

anaphase-promoting complex of Saccharomyces cerevisiae. Science 274, 1201-1204.<br />

Zhou, J., Monson, E.K., Teng, S.C., Schulz, V.P., Zakian, V.A. (2000) Pif1p helicase, a catalytic inhibitor of<br />

telomerase in yeast. Science. 289, 771-774.<br />

Zhu, H., Klemic, J.F., Chang, S., Bertone, P., Casamayor, A., Klemic, K.G., Smith, D., Gerstein, M., Reed, M.A.,<br />

Snyder, M. Analysis of yeast protein kinases using protein chips. Nat. Genet. 26 (2000) 283-289.<br />

Zou, L., Mitchell, J., Stillman, B. (1997) CDC45, a novel yeast gene that functions with <strong>the</strong> origin recognition<br />

complex <strong>and</strong> Mcm proteins in initiation of DNA replication. Mol <strong>Cell</strong> Biol. 17, 553-563.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!