15.12.2012 Views

Transcriptional regulation of meiosis in budding yeast

Transcriptional regulation of meiosis in budding yeast

Transcriptional regulation of meiosis in budding yeast

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> <strong>meiosis</strong><br />

<strong>in</strong> budd<strong>in</strong>g <strong>yeast</strong><br />

Yona Kassir 1* , Noam Adir 2 , Elisabeth Boger-Nadjar 3 , Noga Guttmann Raviv 1 , Ifat Rub<strong>in</strong>-<br />

Bejerano 4 , Shira Sagee 1 , and Galit Shenhar 5 .<br />

1. Department <strong>of</strong> Biology, Technion Haifa Israel; 2. Department <strong>of</strong> Chemistry, Technion Haifa<br />

Israel; 3. Food Eng<strong>in</strong>eer<strong>in</strong>g and Biotechnology, Technion Haifa Israel; 4. Whitehead Institute for<br />

Biomedical Research, Cambride USA. 5. Department <strong>of</strong> Molecular Cell Biology at Weizmann.<br />

Institute, Rehovot, Israel.<br />

*correspond<strong>in</strong>g author. Tel: 972-4-8294214, Fax: 972-4-8225153, e-mail:<br />

ykassir@tx.technion.ac.il<br />

0


I. Introduction<br />

II. <strong>Transcriptional</strong> cascade governs <strong>in</strong>itiation <strong>of</strong> <strong>meiosis</strong>.<br />

III. <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> IME1.<br />

A. The MAT signal<br />

1. Genes whose products transmit the MAT signal to IME1 and <strong>meiosis</strong><br />

1.1. RME1.<br />

1.2. IME4.<br />

1.3. RES1.<br />

B. The nitrogen signal<br />

C. The glucose Signal<br />

1. The cAMP signal-transduction pathway.<br />

1.1. The IREu element.<br />

1.1.1. Msn2 and Msn4.<br />

1.1.2. Sok2.<br />

1.2. Rim15<br />

2. The Snf1 signal-transduction pathway<br />

3. Respiration and the transcription <strong>of</strong> IME1.<br />

D. Additional prote<strong>in</strong>s regulat<strong>in</strong>g the transcription <strong>of</strong> IME1<br />

1. Mck1<br />

2. Tup1/Ssn6<br />

3. Yhp1<br />

4. The Swi/Snf chromat<strong>in</strong> remodel<strong>in</strong>g complex<br />

E. Positive and negative feedback <strong>regulation</strong>.<br />

F. Summary.<br />

IV. <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> early <strong>meiosis</strong>-specific genes<br />

A. Silenc<strong>in</strong>g <strong>of</strong> early <strong>meiosis</strong>-specific genes <strong>in</strong> vegetative growth.<br />

1. The WTM genes.<br />

2. The UME2 gene.<br />

3. The UME3 and UME5 genes.<br />

4. The SIN3, UME6, and RPD3 genes.<br />

5. The ISW2 gene.<br />

B. Expression <strong>of</strong> early <strong>meiosis</strong>-specific genes under meiotic conditions.<br />

1. Ume6 is also positive regulator.<br />

2. The function <strong>of</strong> Ime1.<br />

1


3. Prote<strong>in</strong>s required for the association <strong>of</strong> Ime1 with Ume6.<br />

3.1. The function <strong>of</strong> Rim11 and its homologs Mck1 and Mrk1.<br />

3.2. The function <strong>of</strong> Rim15.<br />

4. The function <strong>of</strong> Gcn5.<br />

5. The transcription <strong>of</strong> early <strong>meiosis</strong>-specific genes is also regulated by positive<br />

elements.<br />

6. Additional positive regulators.<br />

7. The role <strong>of</strong> premeiotic DNA replication and/or recomb<strong>in</strong>ation <strong>in</strong> controll<strong>in</strong>g early<br />

<strong>meiosis</strong>-specific genes expression.<br />

8. The choice between silenc<strong>in</strong>g and expression <strong>of</strong> early <strong>meiosis</strong>-specific genes.<br />

V. <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> middle <strong>meiosis</strong>-specific genes<br />

A. Positive regulators <strong>of</strong> middle <strong>meiosis</strong>-specific genes<br />

1. Ndt80<br />

2. Ime2<br />

3. S<strong>in</strong>3 and Rpd3<br />

B. Negative regulators <strong>of</strong> middle <strong>meiosis</strong>-specific genes<br />

C. The recomb<strong>in</strong>ation checkpo<strong>in</strong>t<br />

VI. <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> late <strong>meiosis</strong>-specific genes<br />

A. Early late genes.<br />

B. Mid late genes<br />

C. Late genes<br />

VII. A feedback loop controll<strong>in</strong>g <strong>meiosis</strong>.<br />

VIII. Conclud<strong>in</strong>g remark. The choice between developmental pathways<br />

2


Abstract<br />

Initiation <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae is regulated by mat<strong>in</strong>g type and nutritional<br />

conditions that restrict <strong>meiosis</strong> to diploid cells grown under starvation conditions. Specifically,<br />

<strong>meiosis</strong> occurs <strong>in</strong> MATa/MATα cells shifted to nitrogen depletion media <strong>in</strong> the absence <strong>of</strong> glucose<br />

and the presence <strong>of</strong> a non-fermentable carbon source. These conditions lead to the expression and<br />

activation <strong>of</strong> Ime1, the master regulator <strong>of</strong> <strong>meiosis</strong>. IME1 encodes a transcriptional activator<br />

recruited to promoters <strong>of</strong> early <strong>meiosis</strong>-specific genes by association with the DNA b<strong>in</strong>d<strong>in</strong>g<br />

prote<strong>in</strong>, Ume6. Under vegetative growth conditions these genes are silent due to recruitment <strong>of</strong><br />

the S<strong>in</strong>3/Rpd3 histone deacetylase and Isw2 chromat<strong>in</strong> remodel<strong>in</strong>g complexes by Ume6.<br />

Transcription <strong>of</strong> these meiotic genes occurs follow<strong>in</strong>g histone acetylation by Gcn5. Expression <strong>of</strong><br />

the early genes promote entry <strong>in</strong>to the meiotic cycle, s<strong>in</strong>ce they <strong>in</strong>clude genes required for<br />

premeiotic DNA synthesis, synapsis <strong>of</strong> homologous chromosomes, and meiotic recomb<strong>in</strong>ation.<br />

Two <strong>of</strong> the early <strong>meiosis</strong> specific genes, a transcriptional activator, Ndt80, and a CDK2 homolog,<br />

Ime2, are required for the transcription <strong>of</strong> middle <strong>meiosis</strong>-specific genes that are <strong>in</strong>volved with<br />

nuclear division and spore formation. Spore maturation depends on late genes whose expression<br />

is <strong>in</strong>directly dependent on Ime1, Ime2 and Ndt80. F<strong>in</strong>ally, phosphorylation <strong>of</strong> Ime1 by Ime2,<br />

leads to its degradation, and consequently to shutt<strong>in</strong>g down <strong>of</strong> the meiotic transcriptional cascade.<br />

This Review is focus<strong>in</strong>g on the <strong>regulation</strong> <strong>of</strong> gene expression govern<strong>in</strong>g <strong>in</strong>itiation and progression<br />

through <strong>meiosis</strong>.<br />

Key Word: <strong>meiosis</strong>, Saccharomyces cerevisiae, transcriptional repression and silenc<strong>in</strong>g,<br />

transcriptional activation, glucose, nitrogen, signal transduction pathways.<br />

List <strong>of</strong> abbreviations: ad – activation doma<strong>in</strong>; bd- DNA b<strong>in</strong>d<strong>in</strong>g doma<strong>in</strong>; CDK – cycl<strong>in</strong><br />

dependent k<strong>in</strong>ase; EMG – early <strong>meiosis</strong>-specific genes; id – <strong>in</strong>teraction doma<strong>in</strong>; LMG – late<br />

<strong>meiosis</strong>-specific genes; MMG – middle <strong>meiosis</strong>-specific genes; MSE - middle sporulation<br />

element; MAPK – Mitogen activated k<strong>in</strong>ase; NIS – Nuclear import sequence; NLS - Nuclear<br />

localization sequence; PKA – cAMP-dependent prote<strong>in</strong> k<strong>in</strong>ase H; SA – synthetic growth media<br />

with acetate as the sole carbon source; SCB - Swi4/6 cell cycle box; SD – synthetic growth media<br />

with glucose as the sole carbon source; SPM – sporulation media; STRE-stress response element;<br />

UAS - Upstream activation sequence; UCS - Upstream controll<strong>in</strong>g sequence; URS - Upstream<br />

repression sequence.<br />

3


I. Introduction<br />

The budd<strong>in</strong>g <strong>yeast</strong> Saccharomyces cerevisiae is a simple unicellular eukaryote exhibit<strong>in</strong>g several<br />

optional developmental pathways. Fig. 1 schematically illustrates the developmental options <strong>of</strong><br />

MATa/MATα diploid cells. In the presence <strong>of</strong> both carbon and nitrogen sources both haploid and<br />

diploid cells adopt the <strong>yeast</strong> form morphology; upon nitrogen limitation, and <strong>in</strong> the presence <strong>of</strong><br />

high levels <strong>of</strong> glucose, a dimorphic transition to a filamentous growth takes places [For recent<br />

reviews see (Gancedo, 2001; Lengeler et al., 2000; Pan et al., 2000)]. Haploid as well as diploid<br />

cells manifest<strong>in</strong>g these forms propagate by the mitotic cell cycle. Upon nitrogen depletion the<br />

developmental decision made by the cells depends on the presence or absence <strong>of</strong> a fermentable<br />

carbon source such as glucose, as well as on the <strong>in</strong>formation at the MAT locus. In the presence <strong>of</strong><br />

functional MATa1 and MATα2 alleles, regardless <strong>of</strong> ploidy, i.e. MATa/MATα diploids,<br />

MATa/MATα disomic cells, and haploid cells express<strong>in</strong>g both MAT alleles, cells enter the meiotic<br />

cycle (Kassir and Simchen, 1976; Kassir and Simchen, 1985; R<strong>in</strong>e and Herskowitz, 1987; Roman<br />

and Sands, 1953; Roth and Fogel, 1971). In the presence <strong>of</strong> only a s<strong>in</strong>gle functional allele, i.e.<br />

MATa and MATα haploids or MATa/MATa and MATα/MATα diploids, depletion <strong>of</strong> nitrogen<br />

leads to a cell cycle arrest at G1 (Kassir and Simchen, 1976; Roman and Sands, 1953; Strathern et<br />

al., 1981). The last developmental option S. cerevisiae cells possess is that <strong>of</strong> mat<strong>in</strong>g, cells<br />

carry<strong>in</strong>g only one <strong>of</strong> the two MAT alleles can respond to the pheromone secreted by cells<br />

express<strong>in</strong>g the other MAT allele by temporal arrest <strong>in</strong> G1, mate, and then resume cell growth<br />

[reviews on the mat<strong>in</strong>g process as well as on how the MAT alleles control cell type are (Fields,<br />

1990; Herskowitz, 1995; Sprague, 1991)].<br />

In this review we focus on how the meiotic signals, namely, the presence <strong>of</strong> the MATa and<br />

MATα alleles, the presence <strong>of</strong> a non-fermentable carbon source, the absence <strong>of</strong> glucose and<br />

nitrogen, control <strong>in</strong>itiation and progression through the meiotic cycle. We focus on transcriptional<br />

<strong>regulation</strong> rather than the function <strong>of</strong> the prote<strong>in</strong>s <strong>in</strong>volved with the specific meiotic events [for<br />

reviews on these aspects <strong>of</strong> <strong>meiosis</strong> see (Kupiec et al., 1997; Roeder, 1997)].<br />

II. <strong>Transcriptional</strong> cascade governs <strong>in</strong>itiation <strong>of</strong> <strong>meiosis</strong>.<br />

Entry and progression through the meiotic cycle depends on the expression and activity <strong>of</strong> many<br />

genes that can be roughly divided <strong>in</strong>to 3 groups: <strong>meiosis</strong>-specific, cell-division cycle (CDC) and<br />

radiation sensitive (RAD) genes. The <strong>meiosis</strong>-specific genes are expressed only under meiotic<br />

conditions, and their function is required only <strong>in</strong> <strong>meiosis</strong>. The cell-division cycle genes that are<br />

4


<strong>in</strong>volved with the nuclear cell cycle are required for <strong>meiosis</strong> (Simchen, 1974), and are expressed<br />

<strong>in</strong> both the mitotic and meiotic cycles. The RAD genes are <strong>in</strong>volved with DNA repair as well as <strong>in</strong><br />

checkpo<strong>in</strong>t surveillance, and <strong>in</strong> <strong>meiosis</strong> are required for meiotic recomb<strong>in</strong>ation and surveillance<br />

mechanisms.<br />

Initiation and progression through the meiotic cycle is regulated by a transcriptional cascade<br />

consist<strong>in</strong>g <strong>of</strong> a temporal and programmed expression <strong>of</strong> 6 classes <strong>of</strong> <strong>meiosis</strong>-specific genes (early<br />

I, early II, early middle, middle, mid-late, and late) (Chu et al., 1998; Primig et al., 2000). Fig. 2<br />

shows a schematic draw<strong>in</strong>g <strong>of</strong> this transcriptional cascade, the tim<strong>in</strong>g <strong>of</strong> meiotic events, and the<br />

<strong>in</strong>put <strong>of</strong> the meiotic signals. IME1 is the master regulator gene absolutely required for entry <strong>in</strong>to<br />

the meiotic cycle and the transcription <strong>of</strong> all <strong>meiosis</strong>-specific genes (Kassir et al., 1988; Smith<br />

and Mitchell, 1989). Briefly, <strong>in</strong> the mitotic cell cycle IME1 is silent, as its transcription is<br />

regulated by the three meiotic signals (Kassir et al., 1988). The translation and activity <strong>of</strong> Ime1 is<br />

also regulated by nutrients (Rub<strong>in</strong>-Bejerano et al., 1996; Sherman et al., 1993). IME1 encodes a<br />

transcriptional activator that is directly required for the transcription <strong>of</strong> early <strong>meiosis</strong>-specific<br />

genes (EMG) (Mandel et al., 1994; Smith et al., 1993). The transcription <strong>of</strong> the middle genes<br />

(MMG) depends on the transcription factor, Ndt80, and on the ser<strong>in</strong>e/threon<strong>in</strong>e prote<strong>in</strong> k<strong>in</strong>ase,<br />

Ime2, two early <strong>meiosis</strong>-specific genes whose transcription depends on Ime1 (Chu and<br />

Herskowitz, 1998; Hepworth et al., 1998; Mitchell et al., 1990; Yoshida et al., 1990). In addition,<br />

the function <strong>of</strong> Ime2 is directly regulated by nutrients (Donzeau and Bandlow, 1999; Mitchell et<br />

al., 1990; Yoshida et al., 1990). <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> mid-late and late genes (LMG) is<br />

less clear, however, their regulated expression depends on the upstream regulators, Ime1, Ime2,<br />

and Ndt80, as well as on nitrogen depletion (Friesen et al., 1997; Kihara et al., 1991).<br />

There is a good correlation between time <strong>of</strong> transcription and meiotic function (Chu et al.,<br />

1998; Primig et al., 2000). The transcription <strong>of</strong> EMG is <strong>in</strong>duced prior to premeiotic DNA<br />

replication, and it <strong>in</strong>cludes genes required for pair<strong>in</strong>g <strong>of</strong> homologous chromosomes (i.e. HOP1),<br />

and meiotic recomb<strong>in</strong>ation (i.e. DMC1). Furthermore, the transcription <strong>of</strong> CDC genes required for<br />

premeiotic DNA replication is <strong>in</strong>duced at this time (i.e. POL1) (Johnston et al., 1986). Middle<br />

genes are <strong>in</strong>duced follow<strong>in</strong>g completion <strong>of</strong> DNA replication and prior to the first nuclear division.<br />

These genes are <strong>in</strong>volved with nuclear division (i.e. CLB1,3,4, CDC26). The late genes are<br />

expressed follow<strong>in</strong>g completion <strong>of</strong> nuclear divisions, and are <strong>in</strong>volved with spore formation and<br />

its maturation (i.e. DIT1, SPS100) (Chu et al., 1998). However, some genes don’t follow this rule.<br />

For example, CLB5,6 are middle genes (Chu et al., 1998) that are required for premeiotic DNA<br />

replication (Dirick et al., 1998; Stuart and Wittenberg, 1998). In the mitotic cell cycle Clb5 is also<br />

required for sp<strong>in</strong>dle orientation, a process occurr<strong>in</strong>g concomitantly with DNA replication (Segal<br />

5


et al., 2000). On the other hand, <strong>in</strong> the meiotic cycle, separation <strong>of</strong> the duplicated Sp<strong>in</strong>dle-Pole<br />

Bodies and sp<strong>in</strong>dle formation occur follow<strong>in</strong>g completion <strong>of</strong> DNA replication (Goetsch and<br />

Byers, 1982). It seems, therefore, that the transcription <strong>of</strong> CLB5/6 correlates with their second<br />

putative meiotic function.<br />

III. <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> IME1.<br />

The different signal pathways that lead to <strong>meiosis</strong> converge at IME1, promot<strong>in</strong>g its transcription<br />

(Kassir et al., 1988). In vegetative growth media with glucose as the sole carbon source (SD -<br />

synthetic dextrose) IME1 is not transcribed. A low basal level <strong>of</strong> IME1 mRNA is present <strong>in</strong><br />

growth media when acetate is the sole carbon source (SA - synthetic acetate) (Kassir et al., 1988).<br />

Upon nitrogen depletion and the presence <strong>of</strong> acetate (SPM - sporulation media), the level <strong>of</strong> IME1<br />

mRNA is transiently <strong>in</strong>creased <strong>in</strong> MATa/MATα diploids, but not <strong>in</strong> cells carry<strong>in</strong>g only a s<strong>in</strong>gle<br />

active MAT allele (Kassir et al., 1988). In addition, IME1 is subject to both positive and negative<br />

feedback <strong>regulation</strong> (Shefer-Vaida et al., 1995). An extremely large region, about 2.1 kb long,<br />

consist<strong>in</strong>g <strong>of</strong> dist<strong>in</strong>ct positive and negative elements mediates the regulated transcription <strong>of</strong> IME1<br />

(Granot et al., 1989; Sagee et al., 1998; Sherman et al., 1993; Smith et al., 1990). By deletion<br />

analysis and <strong>in</strong>sertion <strong>of</strong> specific elements <strong>of</strong> IME1 <strong>in</strong> heterologous reporter genes, the meiotic<br />

signals affect<strong>in</strong>g discrete elements were identified (Fig. 3) (Covitz and Mitchell, 1993; Sagee et<br />

al., 1998; Sherman et al., 1993; Smith et al., 1990). Below is a detailed review on how the MAT,<br />

Glucose, and Nitrogen signals are transmitted to IME1.<br />

A. The MAT signal<br />

The MAT signal is transmitted through two dist<strong>in</strong>ct elements, UCS3 and UCS4, which do not<br />

share any sequence homology (Covitz and Mitchell, 1993; Sagee et al., 1998). Deletion <strong>of</strong> any <strong>of</strong><br />

these elements leads to partial derepression, whereas deletion <strong>of</strong> both elements results <strong>in</strong> full<br />

expression <strong>of</strong> IME1 <strong>in</strong> MATa/MATa cells (Sagee et al., 1998). In addition, UCS3 is apparently<br />

required for complete relief <strong>of</strong> repression <strong>of</strong> UCS4. Nested deletion <strong>of</strong> UCS3 leads to a 2-fold<br />

reduction <strong>in</strong> the expression <strong>of</strong> ime1-lacZ <strong>in</strong> MATa/MATα cells <strong>in</strong>cubated under meiotic<br />

conditions, whereas <strong>in</strong> concomitant deletion <strong>of</strong> UCS4 and the entire upstream region, the<br />

expression <strong>of</strong> ime1-lacZ is complete (Fig. 4, compare l<strong>in</strong>e B to l<strong>in</strong>es A and C). The prote<strong>in</strong>s<br />

through which MAT <strong>regulation</strong> is transmitted to UCS3 are not known. As discussed below, Rme1<br />

transmits the MAT signal to UCS4 (Covitz and Mitchell, 1993) (see section IIIA1.1).<br />

6


Full expression <strong>of</strong> Ime1 <strong>in</strong> S288C stra<strong>in</strong>s deleted for RME1 or UCS3 and UCS4 does not lead<br />

to complete sporulation (Kassir et al., 1988; Sagee et al., 1998), suggest<strong>in</strong>g that the Mata1 Matα2<br />

prote<strong>in</strong>s might be required for a downstream meiotic event. On the other hand, <strong>in</strong> SK1<br />

background, this is not the case, and the MATa1 and MATα2 alleles are required only for the<br />

transcription <strong>of</strong> IME1 (Covitz and Mitchell, 1993). There are many reported cases <strong>of</strong> phenotypic<br />

differences between SK1 and other stra<strong>in</strong>s, i.e. S288C or W303. Genomic DNA hybridization<br />

reveals the presence <strong>of</strong> many deletions and polymorphisms between these stra<strong>in</strong>s (Primig et al.,<br />

2000). Thus, the discrepancy between results is attributed to the absence <strong>of</strong> some functions <strong>in</strong> the<br />

different stra<strong>in</strong>s.<br />

1. Genes whose products transmit the MAT signal to IME1 and <strong>meiosis</strong>. Three<br />

genes are known to transmit the MAT signal to IME1 and <strong>meiosis</strong>: RME1, IME4, and RES1.<br />

1.1. RME1. RME1 encodes a Z<strong>in</strong>c-f<strong>in</strong>ger DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> (Covitz et al., 1991;<br />

Shimizu et al., 1997a) that is a negative regulator for the transcription <strong>of</strong> IME1 and <strong>meiosis</strong> <strong>in</strong><br />

cells carry<strong>in</strong>g only one <strong>of</strong> the MAT alleles (Kassir et al., 1988; Kassir and Simchen, 1976;<br />

Mitchell and Herskowitz, 1986; R<strong>in</strong>e et al., 1981). Recessive mutations <strong>in</strong> RME1 and deletion <strong>of</strong><br />

RME1 lead to complete expression <strong>of</strong> IME1 and sporulation <strong>in</strong> mata1/MATα, MATa/MATa and<br />

MATα/MATα stra<strong>in</strong>s (Kassir et al., 1988; Kassir and Simchen, 1976; Mitchell and Herskowitz,<br />

1986). Haploid cells carry<strong>in</strong>g the rme1-1 mutation complete premeiotic DNA replication, but<br />

arrest as mono nucleate cells without loss <strong>of</strong> viability (Kassir and Simchen, 1976). This is most<br />

probably due to the pachytene checkpo<strong>in</strong>t mechanism that monitors lack <strong>of</strong> synapsis and/or<br />

recomb<strong>in</strong>ation [for review on this checkpo<strong>in</strong>t see (Roeder and Bailis, 2000)]. In MATa/MATα<br />

stra<strong>in</strong>s relief <strong>of</strong> repression is due to the substantial reduction (10 to 20-fold) <strong>in</strong> the level <strong>of</strong> RME1<br />

mRNA and prote<strong>in</strong> [(Mitchell and Herskowitz, 1986), and L. Johnston, personal communication].<br />

The Mata1/Matα2 complex b<strong>in</strong>ds to a specific element <strong>in</strong> the promoter <strong>of</strong> RME1 repress<strong>in</strong>g its<br />

transcription (Goutte and Johnson, 1988; Johnson and Herskowitz, 1985; Li et al., 1995; Stark<br />

and Johnson, 1994).<br />

Rme1 b<strong>in</strong>ds the IME1 5’ region to two sites localized with<strong>in</strong> UCS4, at -2040 to –2030 and –<br />

1959 to –1949 (Shimizu et al., 1998). The consensus sequence for b<strong>in</strong>d<strong>in</strong>g is GWACCTCAARA<br />

(Shimizu et al., 1998). The presence <strong>of</strong> these two b<strong>in</strong>d<strong>in</strong>g sites; def<strong>in</strong>ed as the RRE and the<br />

modulation element (Covitz and Mitchell, 1993), is required to repress the transcription <strong>of</strong> IME1.<br />

Deletion or mutations <strong>in</strong> a s<strong>in</strong>gle site cause only partial derepression (Shimizu et al., 1998).<br />

Moreover, deletion <strong>of</strong> both sites does not lead to complete relief <strong>of</strong> repression, (Shimizu et al.,<br />

1998), probably due to the repression activity <strong>of</strong> the UCS3 site (Sagee et al., 1998). Insertion <strong>of</strong><br />

7


the UCS4 element upstream <strong>of</strong> the CYC1 UAS results <strong>in</strong> 17-fold repression, depend<strong>in</strong>g on overexpression<br />

<strong>of</strong> Rme1, as well as on the presence <strong>of</strong> both the RRE and the modulation sites (Covitz<br />

and Mitchell, 1993). Unlike the <strong>in</strong>tact IME1 gene, repression dependent on over-expression <strong>of</strong><br />

Rme1 and dependence on MAT was not reported. It is not surpris<strong>in</strong>g, therefore, that <strong>in</strong> this<br />

heterologous system repression is accomplished by sequester<strong>in</strong>g the b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> a transcriptional<br />

activator (Shimizu et al., 1997b). The mode by which Rme1 represses transcription <strong>of</strong> IME1 is<br />

not known, however, it repression activity depends on S<strong>in</strong>4 and Rgr1 (Covitz et al., 1994), two<br />

components <strong>of</strong> the RNA polymerase SRB mediator complex (Myer and Young, 1998). These<br />

prote<strong>in</strong>s are not required for the stability <strong>of</strong> Rme1 (Covitz et al., 1994), or it’s b<strong>in</strong>d<strong>in</strong>g to DNA<br />

(Shimizu et al., 1998).<br />

Rme1 also functions as a transcriptional activator; it activates the transcription <strong>of</strong> CLN2 whose<br />

promoter <strong>in</strong>cludes Rme1’s b<strong>in</strong>d<strong>in</strong>g site (Toone et al., 1995). Accord<strong>in</strong>gly, Rme1 also activates the<br />

transcription <strong>of</strong> CYC1 and HIS3 when artificially tethered to their promoters (Covitz and<br />

Mitchell, 1993; Covitz et al., 1994). The transcription <strong>of</strong> RME1 is ma<strong>in</strong>ly <strong>in</strong>duced <strong>in</strong> G1 (Frenz et<br />

al., 2001), suggest<strong>in</strong>g that although this gene is non-essential, it might contribute to the high level<br />

expression <strong>of</strong> Cln2 at this stage <strong>in</strong> the cell cycle. <strong>Transcriptional</strong> activation is <strong>in</strong>dependent <strong>of</strong> Rgr1<br />

or S<strong>in</strong>4 (Blumental-Perry et al., 2002), although the function <strong>of</strong> Rme1 as an activator or repressor<br />

is mediated through the same doma<strong>in</strong> (Blumental-Perry et al., 2002). These results suggest that<br />

the function <strong>of</strong> Rme1 is determ<strong>in</strong>ed by <strong>in</strong>teract<strong>in</strong>g prote<strong>in</strong>s b<strong>in</strong>d<strong>in</strong>g to nearby DNA sites.<br />

The effect <strong>of</strong> MAT on the transcription <strong>of</strong> IME1 is evident only upon nitrogen depletion; In SA<br />

media the level <strong>of</strong> IME1 mRNA is low and identical <strong>in</strong> haploid and MATa/MATα diploid cells<br />

(Kassir et al., 1988), and only SPM and <strong>in</strong> MATa/MATα that IME1 is fully <strong>in</strong>duced. Moreover,<br />

s<strong>in</strong>ce nitrogen depletion <strong>in</strong>duces the transcription <strong>of</strong> RME1 (Frenz et al., 2001), it is possible that<br />

Rme1 represses transcription only <strong>in</strong> the absence <strong>of</strong> nitrogen. This hypothesis predicts that <strong>in</strong> the<br />

absence <strong>of</strong> Rme1, <strong>in</strong>itiation <strong>of</strong> <strong>meiosis</strong> will not depend on nitrogen depletion. However, cells<br />

where either RME1 or UCS3 along with UCS4 (the sites mediat<strong>in</strong>g the MAT signal) have been<br />

deleted, <strong>in</strong>duce <strong>meiosis</strong> only upon nitrogen depletion (Kassir and Simchen, 1976; Sagee et al.,<br />

1998). The effect <strong>of</strong> Rme1 is mediated through one <strong>of</strong> two mechanisms: by direct b<strong>in</strong>d<strong>in</strong>g and<br />

repression <strong>of</strong> IME1, and by activation <strong>of</strong> CLN2 transcription (Frenz et al., 2001; Toone et al.,<br />

1995). The <strong>in</strong>crease <strong>in</strong> activity <strong>of</strong> the Cln/Cdc28 complex could result <strong>in</strong> an <strong>in</strong>crease <strong>in</strong><br />

phosphorylation <strong>of</strong> Ime1, and its sequester<strong>in</strong>g from the nucleus (Colom<strong>in</strong>a et al., 1999).<br />

Consequently, the reduction <strong>in</strong> positive auto<strong>regulation</strong>, would lead to a decrease <strong>in</strong> the level <strong>of</strong><br />

IME1 mRNA.<br />

8


1.2. IME4. IME4 encodes a positive regulator absolutely required for the transcription <strong>of</strong><br />

IME1 (Shah and Clancy, 1992). The transcription <strong>of</strong> IME4 is regulated by the meiotic signals; it<br />

requires the presence <strong>of</strong> the MATa1 and MATα2 gene products and nitrogen depletion. Rme1<br />

does not transmit the MAT signal to IME4 (Shah and Clancy, 1992), suggest<strong>in</strong>g that Rme1 and<br />

Ime4 function <strong>in</strong> two dist<strong>in</strong>ct signal pathways. The region with<strong>in</strong> IME1 respond<strong>in</strong>g to Ime4, and<br />

the mode by which Ime4 activates the transcription <strong>of</strong> IME1 is not known. Over expression <strong>of</strong><br />

Ime1 partially suppresses ime4∆, suggest<strong>in</strong>g that Ime4 have an additional role <strong>in</strong> <strong>meiosis</strong> (Shah<br />

and Clancy, 1992). Ime4 associates with Mum2/SpoT-8 (Uetz et al., 2000) that is required for<br />

premeiotic DNA replication (Engebrecht et al., 1998; Tsuboi, 1983). It is possible that the second<br />

meiotic function <strong>of</strong> Ime4 is to regulate the function <strong>of</strong> this prote<strong>in</strong>.<br />

1.3. RES1. The dom<strong>in</strong>ant mutation Res1-1 bypasses the requirement for the presence <strong>of</strong><br />

both Mata1 and Matα2 for the expression <strong>of</strong> IME1 and <strong>meiosis</strong> (Kao et al., 1990). Res1-1 is not<br />

allelic to RME1, IME1 or IME4 (Kao et al., 1990; Shah and Clancy, 1992), neither is Res1 <strong>in</strong> the<br />

Ime4 or Rme1 signal pathway (Kao et al., 1990; Shah and Clancy, 1992). This gene has not been<br />

cloned, and its normal function is not known.<br />

B. The nitrogen signal<br />

The nitrogen signal is transmitted to IME1 through the UCS1 element (Fig. 3). Nested deletion <strong>of</strong><br />

this region leads to a 5 and 15-fold <strong>in</strong>crease <strong>in</strong> the expression <strong>of</strong> ime1-lacZ <strong>in</strong> vegetative growth<br />

media with either glucose (SD) or acetate (SA) as the sole carbon source, respectively [Fig. 4,<br />

compare l<strong>in</strong>es D and E, (Sagee et al., 1998)]. These results suggest that UCS1 functions as a<br />

negative element <strong>in</strong> the presence <strong>of</strong> glucose and nitrogen. By itself UCS1 has no UAS activity, it<br />

cannot promote expression <strong>of</strong> a his4-lacZ reporter gene lack<strong>in</strong>g its own UAS (Nadjar-Boger,<br />

2000). Insertion <strong>of</strong> UCS1 between the HIS4 UAS and TATA box <strong>in</strong> the HIS4UAS-HIS4TATA-lacZ<br />

reporter gene leads to a substantial reduction <strong>of</strong> expression <strong>in</strong> SD and SA [Fig. 5, (Sagee et al.,<br />

1998)]. Level <strong>of</strong> repression is 2 to 3-fold higher <strong>in</strong> SD <strong>in</strong> comparison to SA, confirm<strong>in</strong>g that the<br />

activity <strong>of</strong> UCS1 is partially regulated by glucose. Upon nitrogen depletion (SPM) a substantial<br />

<strong>in</strong>crease <strong>in</strong> expression is observed [Fig. 5 and (Sagee et al., 1998)], suggest<strong>in</strong>g that nitrogen is the<br />

major signal regulat<strong>in</strong>g the function <strong>of</strong> UCS1.<br />

The cAMP/PKA pathway transmits a nitrogen signal to IME1 and <strong>meiosis</strong> (Matsumoto et al.,<br />

1983; Matsuura et al., 1990). Mutations that cause low or no activity <strong>of</strong> PKA, such as cyr1, ras2,<br />

and cdc25 lead to the expression <strong>of</strong> IME1 and spore formation <strong>in</strong> the presence <strong>of</strong> nitrogen<br />

(Matsumoto et al., 1983; Matsuura et al., 1990; Shilo et al., 1978; Smith and Mitchell, 1989)<br />

[CYR1 encodes adenylate cyclase, RAS2 encodes a small G prote<strong>in</strong> that functions as a positive<br />

9


egulator <strong>of</strong> Cyr1, CDC25 encodes the Ras GDP/GTP exchange factor, which serves as a positive<br />

regulator <strong>of</strong> adenylate cyclase (Broach, 1991; Broek et al., 1987; Toda et al., 1985)]. On the other<br />

hand, mutations that cause constitutive PKA activity, such as RAS2-val19 (activated Ras) and<br />

bcy1 [the regulatory subunit <strong>of</strong> PKA (Broach, 1991)] are sporulation deficient, and are suppressed<br />

by over-expression <strong>of</strong> IME1 (Matsuura et al., 1990). The repression activity <strong>of</strong> UCS1 is reduced<br />

<strong>in</strong> vegetatively grown cdc25-5 cells <strong>in</strong>cubated at the non-permissive temperature (Fig. 5). Upon<br />

nitrogen depletion (SPM), this phenomenon is not observed; the same levels <strong>of</strong> expression are<br />

observed <strong>in</strong> cells <strong>in</strong>cubated at the permissive or restrictive temperature (Fig. 5). These results<br />

suggest that Cdc25 transmits the nitrogen signal to UCS1. Cdc25 is a positive regulator <strong>of</strong> both<br />

the, cAMP/PKA (Broek et al., 1987) and MAPK (Shenhar, 2001) signal transduction pathways.<br />

Therefore, further experiments are required to determ<strong>in</strong>e if this effect <strong>of</strong> Cdc25 is mediated<br />

through PKA, MAPK or a different signal pathway.<br />

MDS3 and its homologue, PMD1, are negative regulators <strong>of</strong> IME1 transcription <strong>in</strong> vegetative<br />

growth media with acetate as the sole carbon source. Deletion <strong>of</strong> these genes results <strong>in</strong> an<br />

<strong>in</strong>crease <strong>in</strong> the transcription <strong>of</strong> IME1 and the EMG IME2 and HOP1 (Benni and Neigeborn,<br />

1997). In addition, when stationary phase mds3∆ pmd1∆ diploid cells are shifted from SD to SA<br />

media, growth is ceased and about 10% <strong>of</strong> the cells enter and complete the meiotic cycle. These<br />

results suggest that Mds3/Pmd1 function upstream <strong>of</strong> Ime1, <strong>in</strong> prevent<strong>in</strong>g cell cycle arrest <strong>in</strong> the<br />

presence <strong>of</strong> nitrogen (Benni and Neigeborn, 1997). Growth arrest and sporulation are suppressed<br />

<strong>in</strong> cells carry<strong>in</strong>g the RAS2-val19 mutation, suggest<strong>in</strong>g that these genes might be upstream<br />

activators <strong>of</strong> the Ras pathway (Benni and Neigeborn, 1997). The region <strong>in</strong> IME1 that is regulated<br />

by these genes is not known. It will be <strong>in</strong>terest<strong>in</strong>g to determ<strong>in</strong>e if their effect is mediated through<br />

the UCS1 element.<br />

As discussed above, the Mata1 and Matα2 gene products are required to <strong>in</strong>duce expression <strong>in</strong><br />

the absence <strong>of</strong> nitrogen, suggest<strong>in</strong>g that UCS1 might also mediate the MAT signal. However, the<br />

repression activity <strong>of</strong> UCS1 and its relief (<strong>in</strong> a UASHIS4-UCS1-his4-lacZ reporter) is identical <strong>in</strong><br />

isogenic haploid and diploid cells (Boger-Nadjar, 2000), <strong>in</strong>dicat<strong>in</strong>g that the activity <strong>of</strong> UCS1 is<br />

not regulated by MAT. However, it is still possible that with<strong>in</strong> the context <strong>of</strong> IME1 promoter,<br />

UCS1 might <strong>in</strong>teract with the UCS3 and/or UCS4 elements that mediate MAT <strong>regulation</strong>.<br />

C. The glucose Signal<br />

The UAS activity <strong>of</strong> IME1 is conf<strong>in</strong>ed to UCS2, a region that conta<strong>in</strong>s alternate positive and<br />

negative elements [Fig. 3, (Sagee et al., 1998)]. Nested deletions <strong>of</strong> either one <strong>of</strong> its three positive<br />

elements, UASru, IREu, and UASrm lead to a reduction <strong>in</strong> expression <strong>of</strong> ime1-lacZ <strong>in</strong> SPM (Fig.<br />

10


4, l<strong>in</strong>es F,G,H), confirm<strong>in</strong>g their function as UAS elements (Sagee et al., 1998; Shenhar and<br />

Kassir, 2001). In addition, <strong>in</strong> the presence <strong>of</strong> glucose UASru and IREu function as repression<br />

elements [Fig. 4, l<strong>in</strong>es F,G and (Shenhar and Kassir, 2001)]. Deletion <strong>of</strong> UASru <strong>in</strong>creases<br />

expression <strong>in</strong> SD, but <strong>in</strong> SA or SPM expression is absent (Fig. 4, l<strong>in</strong>e F), suggest<strong>in</strong>g that UASru<br />

functions as a URS element <strong>in</strong> the presence <strong>of</strong> glucose and as a UAS element <strong>in</strong> the absence <strong>of</strong><br />

glucose and/or the presence <strong>of</strong> acetate. Nested deletion <strong>of</strong> IREu leads to the same levels <strong>of</strong><br />

expression <strong>of</strong> ime1-lacZ <strong>in</strong> both SD and SA media [Fig. 4, l<strong>in</strong>e G, (Sagee et al., 1998)],<br />

suggest<strong>in</strong>g that its URS activity is regulated by either the carbon or nitrogen source. These<br />

UASru, IREu and UASrm element function as a carbon source regulated UAS when <strong>in</strong>serted<br />

upstream <strong>of</strong> a his4-lacZ reporter gene (Fig. 6). Decreased levels <strong>of</strong> expression are observed <strong>in</strong> SD,<br />

and <strong>in</strong>creased expression is observed <strong>in</strong> SA, while <strong>in</strong> SPM there is a slight <strong>in</strong>crease <strong>in</strong> activity, but<br />

only <strong>in</strong> the presence <strong>of</strong> UASru [Fig. 6 and (Sagee et al., 1998)]. These results imply that IREu and<br />

UASrm are regulated only by the glucose signal, and that UASru is ma<strong>in</strong>ly regulated by glucose<br />

but nitrogen has also a partial effect on its activity.<br />

1. The cAMP/PKA signal-transduction pathway. Genetic analysis suggests that the<br />

cAMP/PKA signal pathway transmits a nitrogen signal [see section IIIB and (Gancedo, 2001; Pan<br />

et al., 2000)]. However, biochemical and genetic analysis demonstrate that this pathway is one <strong>of</strong><br />

the major signal transduction pathways transmitt<strong>in</strong>g a glucose signal to <strong>yeast</strong> cells. This is an<br />

essential pathway that transiently <strong>in</strong>creases the level <strong>of</strong> cAMP <strong>in</strong> response to glucose addition [for<br />

reviews see (Broach, 1991; Thevele<strong>in</strong> and de W<strong>in</strong>de, 1999)]. Addition <strong>of</strong> 10mM cAMP to diploid<br />

cells <strong>in</strong>cubated <strong>in</strong> sporulation conditions leads to a small but significant reduction <strong>in</strong> the<br />

expression <strong>of</strong> IME1 and the level <strong>of</strong> sporulation (Fig. 7). The expression <strong>of</strong> either UASrm-his4lacZ<br />

or UASru-his4-lacZ is slightly <strong>in</strong>creased by addition <strong>of</strong> cAMP. However, about 1.6-fold<br />

reduction <strong>in</strong> the expression <strong>of</strong> IREu-his4-lacZ is observed, similar to the effect cAMP has on<br />

sporulation and IME1 expression, suggest<strong>in</strong>g that the activity <strong>of</strong> IREu is negatively regulated by<br />

cAMP. The signal transduction pathways that transmit the glucose signal to UASru and UASrm<br />

are not known.<br />

1.1. The IREu element. The sequence <strong>of</strong> IREu (-1153 to –1121) reveals that it <strong>in</strong>cludes<br />

two known positive elements, a stress response element, STRE, and the Swi4/Swi6 cell cycle<br />

box, SCB (Fig. 8). An almost identical element, IREd, is present at –788 to –756 <strong>of</strong> IME1 (Fig.<br />

3). IREd differs from IREu <strong>in</strong> two residues localized to the STRE and SCB elements (Fig. 8). An<br />

ime1-lacZ chimeras whose 5’ end are term<strong>in</strong>ated downstream or upstream <strong>of</strong> either IREu or IREd<br />

show that IREu functions as a positive element, whereas IREd as a negative element (Sagee et al.,<br />

11


1998). Furthermore, unlike IREu, IREd promotes only low UAS activity when <strong>in</strong>serted upstream<br />

<strong>of</strong> his4-lacZ (Fig. 6), suggest<strong>in</strong>g that the STRE and/or the SCB elements are the transcriptional<br />

activation sequences with<strong>in</strong> IREu. The cAMP/PKA pathway negatively regulates IREu,<br />

promot<strong>in</strong>g its URS activity and prevent<strong>in</strong>g its UAS activity <strong>in</strong> the presence <strong>of</strong> glucose (Sagee et<br />

al., 1998; Shenhar and Kassir, 2001). Deletion <strong>of</strong> BCY1- the regulatory subunit <strong>of</strong> PKA (Toda et<br />

al., 1987a) leads to no expression <strong>of</strong> IREu-his4-lacZ, whereas a temperature sensitive mutations<br />

<strong>in</strong> CDC25 leads to a substantial <strong>in</strong>crease <strong>in</strong> the activity <strong>of</strong> IREu <strong>in</strong> the presence <strong>of</strong> either glucose<br />

or acetate as the sole carbon source (Sagee et al., 1998). Transcription factors that regulate the<br />

activity <strong>of</strong> IREu are the two-homologous DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s, Msn2 and Msn4, Ime1 and<br />

Sok2. Msn2 and Msn4 are absolutely required for the activity <strong>of</strong> IREu, Ime1 is required for the<br />

complete UAS activity <strong>of</strong> IREu, and Sok2 is a negative regulator for IREu activity (Sagee et al.,<br />

1998; Shenhar and Kassir, 2001).<br />

1.1.1. Msn2 and Msn4. These C2H2 Z<strong>in</strong>c-f<strong>in</strong>ger prote<strong>in</strong>s b<strong>in</strong>d the STRE site present <strong>in</strong><br />

many stress-<strong>in</strong>duced genes (Mart<strong>in</strong>ez-Pastor et al., 1996; Schmitt and McEntee, 1996), <strong>in</strong>clud<strong>in</strong>g<br />

the IREu and IREd elements <strong>in</strong> IME1 (Sagee et al., 1998). Competition experiments reveal that<br />

IREu is a better competitor than IREd (Sagee et al., 1998), suggest<strong>in</strong>g that Msn2 and Msn4 b<strong>in</strong>d<br />

to the STRE element <strong>in</strong> IREu. In vitro transcribed and translated Msn2 can b<strong>in</strong>d the STRE<br />

sequence and the IREu element (Mart<strong>in</strong>ez-Pastor et al., 1996; Shenhar, 2001), suggest<strong>in</strong>g that<br />

b<strong>in</strong>d<strong>in</strong>g is <strong>in</strong>dependent <strong>of</strong> post-translational modifications, and/or the presence <strong>of</strong> additional<br />

prote<strong>in</strong>s. However, two l<strong>in</strong>es <strong>of</strong> evidence suggest that Msn2 forms a heterodimer with Sok2 (see<br />

section IIIC1.1.2 below): i. Msn2 physically associates with Sok2, and ii. Deletion <strong>of</strong> the<br />

postulated Sok2 b<strong>in</strong>d<strong>in</strong>g site with<strong>in</strong> IREu (the SCB element) abolishes the activity <strong>of</strong> IREu<br />

(Shenhar and Kassir, 2001). These results suggest that <strong>in</strong> vivo, Msn2 forms a heterodimer with<br />

Sok2, and that Sok2 facilitates its b<strong>in</strong>d<strong>in</strong>g to the DNA. In the absence <strong>of</strong> Sok2, an imposter<br />

prote<strong>in</strong> can promote the b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> Msn2/4 to STRE (Fig. 8) (Shenhar and Kassir, 2001). Msn2/4<br />

are apparent targets <strong>of</strong> the cAMP/PKA pathway. This is concluded from the follow<strong>in</strong>g<br />

observations: i. Deletion <strong>of</strong> both MSN2 and MSN4 suppresses the lethality <strong>of</strong> a stra<strong>in</strong> deleted for<br />

the three TPK genes [TPK1-3 are the three homologous genes encod<strong>in</strong>g the catalytic activity <strong>of</strong><br />

PKA (Smith et al., 1998; Toda et al., 1987b)], ii. Msn2 <strong>in</strong>cludes sites required for its nuclear<br />

import (NLS) as well as for its export (NIS), whose functions are regulated by glucose through<br />

the cAMP/PKA pathway (Gorner et al., 2002). Glucose starvation and <strong>in</strong>activation <strong>of</strong> the<br />

cAMP/PKA pathway through mutations, leads to nuclear localization, whereas addition <strong>of</strong><br />

glucose or cAMP leads to cytoplasmic localization (Gorner et al., 1998; Gorner et al., 2002).<br />

Interest<strong>in</strong>gly, the NIS element is also regulated by the TOR signal<strong>in</strong>g pathway. When this signal<br />

12


pathway is activated, the b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> Msn2/4 to Bmh2, the 14-3-3 adaptor, sequesters it from the<br />

nuclei (Beck and Hall, 1999). 3. The NLS site <strong>of</strong> Msn2 conta<strong>in</strong>s four PKA consensus sites. These<br />

residues are phosphorylated <strong>in</strong> vivo <strong>in</strong> response to glucose addition, depend<strong>in</strong>g on the presence <strong>of</strong><br />

the three TPK genes. Ser<strong>in</strong>e to alan<strong>in</strong>e mutations <strong>of</strong> all these sites lead to constitutive nuclear<br />

localization <strong>of</strong> Msn2, whereas ser<strong>in</strong>e to aspartic acid, mimick<strong>in</strong>g a phosphorylated ser<strong>in</strong>e residue,<br />

leads to its cytoplasmic localization (Gorner et al., 2002).<br />

1.1.2. Sok2. SOK2 encodes a DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> that is highly homologous to the<br />

DNA-b<strong>in</strong>d<strong>in</strong>g doma<strong>in</strong> <strong>of</strong> Swi4 and Phd1 prote<strong>in</strong>s that are known to b<strong>in</strong>d the SCB consensus<br />

sequence. This homology suggests that Sok2 might also b<strong>in</strong>d the same sequence. Direct b<strong>in</strong>d<strong>in</strong>g<br />

<strong>of</strong> Sok2 to IREu was not reported, however, the follow<strong>in</strong>g genetic evidence suggests that Sok2<br />

b<strong>in</strong>ds IREu under all growth conditions. The IREu-his4-lacZ reporter shows a 10-fold <strong>in</strong>crease <strong>in</strong><br />

expression <strong>in</strong> the triple mutant sok2∆ msn2∆ msn4∆ <strong>in</strong> comparison to the double mutant msn2∆<br />

msn4∆. On the other hand, IREu-his4-lacZ is not expressed <strong>in</strong> the sok2T598A msn2∆ msn4∆<br />

triple mutant. The sok2T598A allele, similar to the SOK2 deletion allele is defective <strong>in</strong><br />

repression, but unlike the latter, the defective Sok2 prote<strong>in</strong> is present <strong>in</strong> the cells (Shenhar and<br />

Kassir, 2001). These results imply that Sok2 b<strong>in</strong>ds IREu under all growth conditions, and that <strong>in</strong><br />

its physical absence an imposter prote<strong>in</strong> can b<strong>in</strong>d and activate transcription. Sok2 functions as a<br />

transcriptional repressor for several PKA target genes such as GAC1, SSA3, SWI5 and IME1 (Pan<br />

and Heitman, 2000; Shenhar and Kassir, 2001; Ward et al., 1995). Furthermore, it functions as a<br />

negative regulator <strong>of</strong> pseudohyphal growth (Ward et al., 1995) and <strong>meiosis</strong> (Shenhar and Kassir,<br />

2001), and as a positive regulator <strong>in</strong> the mitotic cell cycle (Ward et al., 1995). Glucose regulates<br />

both the transcription <strong>of</strong> SOK2 and the repression activity <strong>of</strong> Sok2 prote<strong>in</strong>. The latter is concluded<br />

from the observation that a Sok2-Gal4(bd), expressed from the ADH1 promoter, represses the<br />

transcription <strong>of</strong> UASGAL1-UASHIS4-his4-lacZ, but only <strong>in</strong> the presence <strong>of</strong> glucose (SD media)<br />

(Shenhar and Kassir, 2001). The signal pathway regulat<strong>in</strong>g the transcription <strong>of</strong> SOK2 is not<br />

known (Shenhar and Kassir, 2001). However, the signal pathway transmitt<strong>in</strong>g the glucose signal<br />

regulat<strong>in</strong>g Sok2 repression activity is the cAMP/PKA pathway. This is evident from the<br />

follow<strong>in</strong>g observations: i. Over-expression <strong>of</strong> SOK2 suppresses the temperature sensitive growth<br />

defect <strong>of</strong> a tpk1∆ tpk2-ts tpk3∆ mutant (Ward et al., 1995); ii. Repression <strong>of</strong> IREu-his4-lacZ and<br />

UASgal1-UAShis4-his4-lacZ by either Sok2 or Gal4(bd)-Sok2, respectively, is relieved <strong>in</strong> cdc25-<br />

5 cells <strong>in</strong>cubated at the non-permissive temperature, as well as by a threon<strong>in</strong>e to alan<strong>in</strong>e mutation<br />

<strong>in</strong> a PKA consensus site <strong>in</strong> Sok2 (Shenhar and Kassir, 2001). Furthermore, Sok2 is a<br />

phosphoprote<strong>in</strong> whose phosphorylation depends on this same threon<strong>in</strong>e (T598) residue (Shenhar<br />

and Kassir, 2001).<br />

13


In the absence <strong>of</strong> glucose Sok2 does not repress transcription. Relief <strong>of</strong> repression <strong>in</strong> the<br />

absence <strong>of</strong> glucose depends on its N-term<strong>in</strong>al doma<strong>in</strong>, am<strong>in</strong>o acids 1-247 as well as on Ime1 [see<br />

section IIIE and (Shenhar and Kassir, 2001)]. The mode by which Sok2 represses transcription<br />

and how Ime1 relieves repression is not known.<br />

1.2. Rim15. Rim15 encodes a ser<strong>in</strong>e/threon<strong>in</strong>e prote<strong>in</strong> k<strong>in</strong>ase whose activity <strong>in</strong> glucose<br />

growth media is <strong>in</strong>hibited due to its phosphorylation by PKA (Re<strong>in</strong>ders et al., 1998). In addition,<br />

the transcription <strong>of</strong> RIM15, and consequently the steady state level <strong>of</strong> Rim15, is <strong>in</strong>creased <strong>in</strong><br />

acetate growth media (Vidan and Mitchell, 1997). Diploid cells deleted for RIM15 or carry<strong>in</strong>g a<br />

k<strong>in</strong>ase-dead po<strong>in</strong>t mutation are sporulation deficient (Vidan and Mitchell, 1997). Deletion <strong>of</strong><br />

RIM15 causes a drastic reduction <strong>in</strong> the transcription <strong>of</strong> IME1. The mode <strong>of</strong> action <strong>of</strong> Rim15, and<br />

the IME1 element regulated by it are not known.<br />

2. The Snf1 signal-transduction pathway. SNF1 encodes a prote<strong>in</strong> k<strong>in</strong>ase whose activity as a<br />

k<strong>in</strong>ase is <strong>in</strong>hibited <strong>in</strong> the presence <strong>of</strong> high levels <strong>of</strong> glucose (Jiang and Carlson, 1996). Snf1<br />

function is required for the expression <strong>of</strong> glucose-repressed genes (Lesage et al., 1996). Snf1<br />

phosphorylates the DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>, Mig1 (Treitel et al., 1998) that recruits the Tup1/Ssn6<br />

repression complex to the glucose-repressed genes to whom it b<strong>in</strong>ds (Nehl<strong>in</strong> et al., 1991). Snf1 is<br />

required for the high level <strong>in</strong>duction <strong>in</strong> the transcription <strong>of</strong> IME1 <strong>in</strong> sporulation conditions, and<br />

diploid cells deleted for SNF1 arrest prior to premeiotic DNA replication, meiotic recomb<strong>in</strong>ation,<br />

and spore formation (Honigberg and Lee, 1998). Over expression <strong>of</strong> either Msn2 or Msn4<br />

suppresses a temperature sensitive allele <strong>of</strong> SNF1 (Estruch and Carlson, 1993), suggest<strong>in</strong>g that<br />

Snf1 effect on IME1 might be mediated through Msn2/4. It is not known, though, whether<br />

Msn2/4 are direct targets <strong>of</strong> Snf1. Genetic evidence suggests that Snf1 is <strong>in</strong> a separate pathway<br />

than the cAMP/PKA pathway (Thompson-Jaeger et al., 1991), and both regulate Msn2/4. The<br />

sequence <strong>of</strong> IME1 shows no homology to the Mig1 b<strong>in</strong>d<strong>in</strong>g site, and the region <strong>in</strong> IME1 regulated<br />

by Snf1 is not known.<br />

In the meiotic cycle Snf1 is also required follow<strong>in</strong>g the transcription <strong>of</strong> IME1. When IME1 is<br />

placed on a multi-copy plasmid it is highly expressed <strong>in</strong> both wild type and snf1∆ diploid cells,<br />

nevertheless, <strong>in</strong> the SNF1 deleted stra<strong>in</strong> over express<strong>in</strong>g Ime1, the EMG IME2 is only partially<br />

expressed, 10-15% <strong>of</strong> the cells complete premeiotic DNA replication and meiotic recomb<strong>in</strong>ation,<br />

but spores are not formed (Honigberg and Lee, 1998). It was suggested that Snf1 is required <strong>in</strong><br />

the meiotic cycle at three po<strong>in</strong>ts, for the transcription <strong>of</strong> Ime1, for the transcription <strong>of</strong> Ime2, and<br />

for spore formation (Honigberg and Lee, 1998).<br />

14


3. Respiration and the transcription <strong>of</strong> IME1. The presence <strong>of</strong> glucose prevents the activity<br />

<strong>of</strong> the mitochondria whose function is required for sporulation (Berger and Yaffe, 2000). The<br />

transcription <strong>of</strong> IME1 is absent <strong>in</strong> diploid cells without mitochondria – petites, or <strong>in</strong> cells treated<br />

with oligomyc<strong>in</strong> - an <strong>in</strong>hibitor <strong>of</strong> the mitochondrial ATPase (Tre<strong>in</strong><strong>in</strong> and Simchen, 1993). It is not<br />

known whether mitochondrial function is also required at a stage preced<strong>in</strong>g the transcription <strong>of</strong><br />

IME1. The site with<strong>in</strong> IME1 that responds to mitochondria function is not known, nor is it known<br />

how the respiration signal is transmitted. The connection between mitochondria function and<br />

IME1 transcription is also evident from the identification <strong>of</strong> the Rim1 (Rim101) Rim8, Rim9,<br />

Rim13 and Rim20 prote<strong>in</strong>s as positive regulators <strong>of</strong> IME1 (Li and Mitchell, 1997; Su and<br />

Mitchell, 1993a). RIM1 encodes a C2H2 z<strong>in</strong>c f<strong>in</strong>ger prote<strong>in</strong> that is required for high-level<br />

expression <strong>of</strong> IME1 as well as for efficient sporulation (Su and Mitchell, 1993b). Activation <strong>of</strong><br />

Rim1 requires its cleavage by the products <strong>of</strong> Rim8, Rim9 and Rim13 (Li and Mitchell, 1997), as<br />

well as the activity <strong>of</strong> Rim20 that might be required for present<strong>in</strong>g Rim1 to Rim13 (Xu and<br />

Mitchell, 2001). Rim1 is localized to mitochondria, and is required for mitochondria DNA<br />

replication, and consequently for the ma<strong>in</strong>tenance <strong>of</strong> the mitochondria (Berger and Yaffe, 2000;<br />

Van Dyck et al., 1992). Thus, the effect <strong>of</strong> these RIM1,8,9,13,20 genes on the transcription <strong>of</strong><br />

IME1 may be <strong>in</strong>direct, through their effect on the ma<strong>in</strong>tenance <strong>of</strong> the mitochondria.<br />

D. Additional prote<strong>in</strong>s regulat<strong>in</strong>g the transcription <strong>of</strong> IME1.<br />

1. Mck1. MCK1 encodes a prote<strong>in</strong> k<strong>in</strong>ase required for efficient transcription <strong>of</strong> IME1, for<br />

expression <strong>of</strong> early and middle genes such as IME2 and SPS1,2 respectively, and for spore<br />

maturation (Neigeborn and Mitchell, 1991). Expression <strong>of</strong> IME1 from the hetrologous GAL1<br />

promoter suppresses the requirement <strong>of</strong> Mck1 for the transcription <strong>of</strong> EMG and MMG, but spore<br />

maturation rema<strong>in</strong>s defective, suggest<strong>in</strong>g that Mck1 directly regulate IME1 transcription and<br />

spore maturation (Neigeborn and Mitchell, 1991). MCK1 is one <strong>of</strong> 4 <strong>yeast</strong> genes show<strong>in</strong>g<br />

homology to mammalian glycogen synthase k<strong>in</strong>ase 3 (GSK3-β), RIM11, YOL128c, and MRK1.<br />

These homologs are not required for the transcription <strong>of</strong> IME1 (Hajji et al., 1999; Mandel et al.,<br />

1994; Rabitsch et al., 2001), suggest<strong>in</strong>g that the transcription <strong>of</strong> IME1 is only partially dependent<br />

on Mck1.<br />

Mck1 <strong>in</strong>hibits the k<strong>in</strong>ase activity <strong>of</strong> PKA both <strong>in</strong> vitro and <strong>in</strong> vivo, through a mechanism that<br />

does not depend on Mck1 k<strong>in</strong>ase activity (Rayner et al., 2002). This result implies that deletion <strong>of</strong><br />

MCK1 might <strong>in</strong>crease rather than decrease the transcription <strong>of</strong> IME1. Therefore, the effect <strong>of</strong><br />

Mck1 on the transcription <strong>of</strong> IME1 is most probably not mediated by PKA. Mck1 is subject to<br />

15


autophosphorylation, <strong>in</strong>clud<strong>in</strong>g on a conserved tyros<strong>in</strong>e (Y199) (Rayner et al., 2002).<br />

Phosphorylation <strong>of</strong> this tyros<strong>in</strong>e residue <strong>in</strong> Mck1, similar to other GSK3β homologs, is required<br />

for its k<strong>in</strong>ase activity on exogenous substrates, but not for autophosphorylation (Rayner et al.,<br />

2002). The Ptp2 and Ptp3 tyros<strong>in</strong>e phosphatases are required for Mck1 dephosphorylation (Zhan<br />

et al., 2000). Deletion <strong>of</strong> either PTP2 or PTP3 has no effect on <strong>meiosis</strong>, however, the double<br />

mutant ptp2∆ ptp3∆ is sporulation deficient, it arrests <strong>in</strong> <strong>meiosis</strong> prior to premeiotic DNA<br />

replication, with a significant reduction <strong>in</strong> the expression <strong>of</strong> IME1 and IME2, and no expression<br />

<strong>of</strong> middle or late <strong>meiosis</strong>-specific genes (Zhan et al., 2000). A third tyros<strong>in</strong>e phosphatase, Yvh1,<br />

might contribute to this <strong>regulation</strong> (Guan et al., 1992). Deletion <strong>of</strong> YVH1 leads to m<strong>in</strong>or effects on<br />

the transcription <strong>of</strong> IME1, a reduced transcription <strong>of</strong> IME2, and a reduction <strong>in</strong> the efficiency <strong>of</strong><br />

asci formation (Park et al., 1996), but <strong>in</strong> the double mutant yvh1∆ ptp2∆ sporulation is almost<br />

dim<strong>in</strong>ished (Park et al., 1996). Thus, the effect <strong>of</strong> these tyros<strong>in</strong>e phosphatases on <strong>meiosis</strong> may be<br />

partially mediated through Mck1. The transcription <strong>of</strong> YVH1 is <strong>in</strong>duced upon nitrogen depletion,<br />

suggest<strong>in</strong>g that it might be <strong>in</strong>volved with transmitt<strong>in</strong>g the nitrogen signal (Park et al., 1996).<br />

Over expression <strong>of</strong> Mck1 suppresses the sporulation defect <strong>of</strong> cells deleted for TPS1 (De Silva-<br />

Udawatta and Cannon, 2001). TPS1 encodes one <strong>of</strong> the subunits <strong>of</strong> the trehalose phosphate<br />

synthase complex (Re<strong>in</strong>ders et al., 1997). Thus, Tps1 may either regulate Mck1, or suppression <strong>of</strong><br />

mck1∆ by Tps1 is due to unrelated <strong>in</strong>crease <strong>in</strong> the level <strong>of</strong> Ime1. The region <strong>in</strong> IME1 regulated by<br />

Mck1 and its mode <strong>of</strong> function is not known. However, the additive effects <strong>of</strong> mutations <strong>in</strong><br />

MCK1, MDS3, PMD1, RIM1 and IME4 suggest that these genes affect different regulatory<br />

elements <strong>in</strong> the IME1 promoter (Su and Mitchell, 1993b).<br />

2. Tup1/Ssn6. The Tup1/Ssn6 complex functions as a general transcriptional repressor upon<br />

recruitment by specific DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s (Keleher et al., 1992). These prote<strong>in</strong>s are also<br />

negative regulators for the transcription <strong>of</strong> IME1 (Mizuno et al., 1998), but the specific DNAb<strong>in</strong>d<strong>in</strong>g<br />

prote<strong>in</strong> that tethers them to IME1 is not known. Two regions <strong>in</strong> IME1 promoter respond<br />

to Tup1, the first one is from -914 to – 621 and the second from -1215 to –915 (Mizuno et al.,<br />

1998), function<strong>in</strong>g as URS <strong>in</strong> the presence <strong>of</strong> Tup1, and UAS <strong>in</strong> the absence <strong>of</strong> Tup1,<br />

respectively (Mizuno et al., 1998). S<strong>in</strong>ce these regions carry the IREd and IREu repeated<br />

elements (Fig. 3) that function as repression and activation sequences, respectively, it should be<br />

<strong>in</strong>terest<strong>in</strong>g to determ<strong>in</strong>e if the Tup1/Ssn6 complex mediates its effect through these elements, and<br />

whether Sok2 is the DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> that recruits them to these elements.<br />

16


3. Yhp1. YHP1 encodes a homeodoma<strong>in</strong> prote<strong>in</strong> that b<strong>in</strong>ds to the UASv element <strong>in</strong> IME1<br />

promoter, between -702 to –675 (Fig. 3) (Kunoh et al., 2000). This 28 bp element by itself does<br />

not function as a UAS, but when tethered to heterologous reporter genes, it causes a 2-fold<br />

reduction <strong>in</strong> expression, suggest<strong>in</strong>g that Yhp1 functions as a repressor (Kunoh et al., 2000). The<br />

mode by which Yhp1 represses transcription is not known, but its effect is <strong>in</strong>dependent <strong>of</strong> the<br />

Tup1/Ssn6 repression complex (Kunoh et al., 2000). The transcription <strong>of</strong> YHP1 is reduced <strong>in</strong><br />

vegetative cells grown <strong>in</strong> the presence <strong>of</strong> acetate as the sole carbon source <strong>in</strong> comparison to<br />

glucose (Kunoh et al., 2000), imply<strong>in</strong>g that it might transmit a glucose signal. However, deletion<br />

<strong>of</strong> YHP1 has no effect on either the transcription <strong>of</strong> IME1, or on the ability <strong>of</strong> cells to enter and<br />

complete <strong>meiosis</strong> and sporulation (Kunoh et al., 2000).<br />

4. The Swi/Snf chromat<strong>in</strong> remodel<strong>in</strong>g complex. Two components <strong>of</strong> the Swi/Snf<br />

complex, Snf2 and Swi1 are required for high-level expression <strong>of</strong> IME1 (Yoshimoto et al., 1993).<br />

This complex is <strong>in</strong>volved with transcriptional activation <strong>of</strong> genes to whom it is tethered through<br />

<strong>in</strong>teraction with specific activation doma<strong>in</strong>s [for review see (Carlson and Laurent, 1994)].<br />

E. Positive and Negative feedback <strong>regulation</strong>.<br />

The transcription <strong>of</strong> IME1 is subject to positive <strong>regulation</strong>. Over expression <strong>of</strong> Ime1 leads to<br />

<strong>in</strong>crease <strong>in</strong> the level <strong>of</strong> ime1-lacZ, through its effect on the function <strong>of</strong> IREu (Shenhar and Kassir,<br />

2001). The effect <strong>of</strong> Ime1 on additional elements has not yet been determ<strong>in</strong>ed. This<br />

auto<strong>regulation</strong> is required to relieve repression by Sok2. This is deduced from the follow<strong>in</strong>g<br />

results: i. Expression <strong>of</strong> IREu-his4-lacZ is substantially reduced <strong>in</strong> diploid cells deleted for IME1<br />

(Shenhar and Kassir, 2001), ii. Gal4(bd)-Sok2 represses the transcription <strong>of</strong> GAL1UAS-HIS4 UAS -<br />

his4-lacZ <strong>in</strong> SD but has no repression activity <strong>in</strong> SA (Shenhar and Kassir, 2001). However, <strong>in</strong><br />

diploid cells deleted for IME1 the transcription <strong>of</strong> the reporter gene is repressed <strong>in</strong> both SD and<br />

SA media (Shenhar and Kassir, 2001). S<strong>in</strong>ce <strong>in</strong> the presence <strong>of</strong> glucose Ime1 is not expressed, it<br />

is not surpris<strong>in</strong>g that the effect <strong>of</strong> Ime1 is observed only <strong>in</strong> SA. As will be detailed <strong>in</strong> section<br />

IVB2, Ime1 does not b<strong>in</strong>d directly to DNA, and is recruited to the DNA by <strong>in</strong>teract<strong>in</strong>g with a<br />

DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>. S<strong>in</strong>ce the N-term<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> Sok2 is also required to relieve repression<br />

<strong>in</strong> SA, it is tempt<strong>in</strong>g to suggest that Ime1 associates with the N-term<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> Sok2. No<br />

<strong>in</strong>formation on possible associations between these two prote<strong>in</strong>s has yet been reported.<br />

The transcription <strong>of</strong> IME1 as well as the steady-state level <strong>of</strong> Ime1 prote<strong>in</strong> <strong>in</strong> meiotic<br />

conditions is transient, when <strong>in</strong>duced upon a shift to SPM, the level <strong>of</strong> IME1 peaks at about 6-8<br />

17


hours <strong>in</strong> SPM, followed by a decl<strong>in</strong>e (Guttmann-Raviv and Kassir, 2002; Kassir et al., 1988;<br />

Shefer-Vaida et al., 1995). This transient transcription depends on both Ime1 and Ime2<br />

(Guttmann-Raviv and Kassir, 2002; Mitchell et al., 1990; Shefer-Vaida et al., 1995; Smith and<br />

Mitchell, 1989; Yoshida et al., 1990). This negative feedback effect is on transcription <strong>of</strong> IME1<br />

rather than the stability <strong>of</strong> the mRNA, s<strong>in</strong>ce <strong>in</strong>sertions <strong>of</strong> two separate regions, -621 to –924 and –<br />

924 to –1368 (Fig. 3) upstream <strong>of</strong> cyc1-lacZ exhibit the same feedback <strong>regulation</strong> (Shefer-Vaida<br />

et al., 1995). Ime1 is a non-stable prote<strong>in</strong> that is degraded by the proteasome and whose half-life<br />

depends on Ime2 (Guttmann-Raviv and Kassir, 2002). IME2 encodes a <strong>meiosis</strong>-specific prote<strong>in</strong><br />

k<strong>in</strong>ase that <strong>in</strong>teracts with Ime1 and phosphorylates it <strong>in</strong> vitro (Guttmann-Raviv and Kassir, 2002).<br />

We suggest that the effect <strong>of</strong> Ime2 on shutt<strong>in</strong>g down the transcription <strong>of</strong> IME1 is due to its effect<br />

on the stability <strong>of</strong> Ime1 prote<strong>in</strong>. In the presence <strong>of</strong> Ime2, degradation <strong>of</strong> Ime1 would lead to the<br />

establishment <strong>of</strong> Sok2 repression and silenc<strong>in</strong>g. On the other hand, <strong>in</strong> cells deleted for IME2,<br />

accumulation <strong>of</strong> Ime1 leads to cont<strong>in</strong>ues positive auto<strong>regulation</strong>, and <strong>in</strong>crease <strong>in</strong> its transcription,<br />

and consequently the availability <strong>of</strong> Ime1 prote<strong>in</strong> (Guttmann-Raviv and Kassir, 2002).<br />

F. Summary.<br />

IME1 encodes the master regulator required to open a developmental pathway, that <strong>of</strong> <strong>meiosis</strong>, <strong>in</strong><br />

S. cerevisiae. It is not surpris<strong>in</strong>g, therefore, that its transcription is regulated by an unusual large<br />

5’ region that is subject to multiple <strong>regulation</strong>s. Fig. 9 is a summary <strong>of</strong> the results described<br />

above, on the current knowledge on how the meiotic signals are transmitted to IME1. A<br />

comb<strong>in</strong>atorial effect <strong>of</strong> at least 10 dist<strong>in</strong>ct elements ensures that IME1 will be transcribed only<br />

under the appropriate conditions, namely <strong>in</strong> MATa/MATα diploids, the absence <strong>of</strong> glucose, the<br />

presence <strong>of</strong> acetate, and nitrogen depletion. Two negative elements, UCS3 and UCS4 restrict<br />

elevated transcription <strong>of</strong> IME1 to cells express<strong>in</strong>g Mata1 and Matα2 peptides. In vegetative<br />

growth media with glucose as the sole carbon source IME1 is silent. This <strong>regulation</strong> is<br />

accomplished by us<strong>in</strong>g 4 dist<strong>in</strong>ct elements, UCS1 IREu UASru, and UASv, whose function as<br />

repression elements is mediated through dist<strong>in</strong>ct signal transduction pathways. In acetate growth<br />

media a low level <strong>of</strong> transcription is accomplished by the positive activity <strong>of</strong> UASru, IREu,<br />

UASrm and UASv that is opposed by negative <strong>regulation</strong> from UCS1 as well as from three<br />

constitutive URS elements, URSu, URSd and IREd. Upon nitrogen depletion, relief <strong>of</strong> UCS1<br />

repression promotes an <strong>in</strong>crease <strong>in</strong> transcription.<br />

18


IV. <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> early <strong>meiosis</strong>-specific genes (EMG)<br />

A. Silenc<strong>in</strong>g <strong>of</strong> early <strong>meiosis</strong>-specific genes <strong>in</strong> vegetative growth.<br />

In vegetative growth conditions with either glucose or acetate as the sole carbon source <strong>meiosis</strong>specific<br />

genes are silent. Silenc<strong>in</strong>g <strong>of</strong> EMG is not only due to the absence <strong>of</strong> their ma<strong>in</strong><br />

transcriptional activator, Ime1, but is rather due to active repression. The genes required to<br />

promote this silenc<strong>in</strong>g are WTM3, UME2, UME3, SIN3, RPD3, UME5, UME6, and ISW2<br />

(Goldmark et al., 2000; Strich et al., 1989; Strich et al., 1994; Vidal and Gaber, 1991). Except for<br />

Ume6 and Isw2, these genes are not essential for the expression <strong>of</strong> EMG under meiotic conditions<br />

(Goldmark et al., 2000; Strich et al., 1989).<br />

1. The WTM genes. UME1 (WTM3) and its two homologs WTM1 and WTM2 repress the<br />

transcription <strong>of</strong> the EMG IME2 synergistically (Pemberton and Blobel, 1997). These prote<strong>in</strong>s also<br />

function as transcriptional repressors <strong>of</strong> the silent HMR cassette, and when tethered artificially to<br />

heterologous genes. These genes are non-essential; the wtm1 wtm2 wmt3 triple mutant is viable<br />

(Pemberton and Blobel, 1997). The Wtm prote<strong>in</strong>s are expressed constitutively <strong>in</strong> both mitotic and<br />

meiotic conditions, and are localized to the nucleus (Pemberton and Blobel, 1997). The mode by<br />

which they cause repression is not known.<br />

2. The UME2 gene. UME2 (SRB9, SSN2) is a component <strong>of</strong> the SRB subcomplex <strong>of</strong> RNA<br />

polymerase II holoenzyme (Kornberg, 1999). Although it is a negative regulator <strong>of</strong> EMG under<br />

vegetative growth conditions (Strich et al., 1989), when fused to lexA it activates transcription<br />

(Song and Carlson, 1998).<br />

3. The UME3 and UME5 genes. The UME3 (SRB11, SSN8) and UME5 (SRB10, SSN3)<br />

encode <strong>in</strong>tegral components <strong>of</strong> the SRB subcomplex <strong>of</strong> RNA polymerase II holoenzyme (Cooper<br />

and Strich, 1998; Kornberg, 1999). UME3 encodes a cycl<strong>in</strong> C homolog that associates with the<br />

cycl<strong>in</strong> dependent k<strong>in</strong>ase, Ume5 (Cooper et al., 1997; Cooper and Strich, 1998). Ume5<br />

phosphorylates the C-term<strong>in</strong>al CTD repeats <strong>of</strong> RNA polymerase II, but this event is <strong>in</strong>dependent<br />

<strong>of</strong> its repression activity (Cooper and Strich, 1998), thus, it is not know how Ume5 represses the<br />

transcription <strong>of</strong> EMG. Ume3 is degraded <strong>in</strong> meiotic conditions, and this degradation, that is<br />

<strong>in</strong>dependent <strong>of</strong> Ume5, is required for complete relief <strong>of</strong> repression <strong>of</strong> EMG (Cooper et al., 1997).<br />

19


4. The SIN3, UME6, and RPD3 genes. UME6 encodes a C6Zn2 prote<strong>in</strong> that b<strong>in</strong>ds the<br />

URS1 sequence present <strong>in</strong> many genes <strong>in</strong>clud<strong>in</strong>g EMG (Anderson et al., 1995; Strich et al.,<br />

1994). The C6Zn2 doma<strong>in</strong> is sufficient for b<strong>in</strong>d<strong>in</strong>g, and is required for its repression activity<br />

(Anderson et al., 1995; Strich et al., 1994). Deletion <strong>of</strong> UME6 causes high expression <strong>of</strong> EMG, as<br />

well as additional non-meiotic genes carry<strong>in</strong>g the URS1 element, <strong>in</strong> vegetative growth conditions<br />

(Bowdish and Mitchell, 1993; Lopes et al., 1993; Park et al., 1992; Strich et al., 1994). In<br />

addition, Ume6 functions as a transcriptional repressor when tethered to heterologous genes<br />

follow<strong>in</strong>g its fusion to lexA (Kadosh and Struhl, 1997). This repression activity <strong>of</strong> Ume6 depends<br />

on the availability <strong>of</strong> S<strong>in</strong>3 (Ume4, Rpd1) and Rpd3 (Kadosh and Struhl, 1997).<br />

Coimmunoprecipitation and two-hybrid assays demonstrate physical association between<br />

Ume6(515-530) and S<strong>in</strong>3(426-472) (the PAH2 doma<strong>in</strong>), as well as between S<strong>in</strong>3 and Rpd3<br />

(Kadosh and Struhl, 1997; Kasten et al., 1997; Washburn and Esposito, 2001).<br />

S<strong>in</strong>3 functions as a transcriptional repressor when tethered to heterologous genes follow<strong>in</strong>g its<br />

fusion to lexA (Wang and Stillman, 1993). The repression activity <strong>of</strong> S<strong>in</strong>3 depends on Rpd3<br />

(Kadosh and Struhl, 1997; Kasten et al., 1997). In accord, the double mutant s<strong>in</strong>3 rpd3 have the<br />

same phenotype as either s<strong>in</strong>gle mutant (Vidal and Gaber, 1991).<br />

RPD3 encodes histone deacetylase whose deletion, similar to SIN3 deletion, causes an <strong>in</strong>crease<br />

<strong>in</strong> acetylation <strong>of</strong> lys<strong>in</strong>es 5 and 12 <strong>in</strong> histone H4 <strong>in</strong> various genes carry<strong>in</strong>g the URS1 element<br />

(Burgess et al., 1999; Rundlett et al., 1998). Recruitment <strong>of</strong> the S<strong>in</strong>3/Rpd3 complex by Ume6 to<br />

DNA results <strong>in</strong> decreased acetylation <strong>of</strong> histone H3 and H4 <strong>in</strong> a restricted region, up to two<br />

nucleosomes from the b<strong>in</strong>d<strong>in</strong>g site <strong>of</strong> Ume6 (Kadosh and Struhl, 1998).<br />

5. The ISW2 gene. Isw2 <strong>in</strong> a complex with Itc1 functions as an ATP dependent chromat<strong>in</strong>remodel<strong>in</strong>g<br />

factor that is required for the repression <strong>of</strong> EMG under vegetative growth conditions<br />

(Goldmark et al., 2000). Isw2 functions <strong>in</strong> parallel to the S<strong>in</strong>3/Rpd3 histone deacetylase complex,<br />

as deduced from the observation that the level <strong>of</strong> expression <strong>of</strong> EMG are dramatically <strong>in</strong>creased<br />

<strong>in</strong> the s<strong>in</strong>3 isw2 double mutant <strong>in</strong> comparison to the s<strong>in</strong>gle mutants (Goldmark et al., 2000). Isw2<br />

physically associates with Ume6, and s<strong>in</strong>ce it’s b<strong>in</strong>d<strong>in</strong>g to the URS1 element depends on Ume6,<br />

it was concluded that Ume6 recruits the Isw2 complex to URS1 (Goldmark et al., 2000). DNaseI<br />

and micrococcal nuclease digestion <strong>of</strong> the chromat<strong>in</strong> near the EMG REC104 show dependence on<br />

Isw2, suggest<strong>in</strong>g that the Isw2 complex forms an <strong>in</strong>accessible chromat<strong>in</strong> structure near the URS1<br />

element (Goldmark et al., 2000). These results suggest that Isw2 would be localized to the<br />

nucleus <strong>in</strong> vegetative growth media. However, this is not the case, Isw2 is localized to the<br />

cytoplasm <strong>in</strong> vegetative growth cultures, and to the nucleus, the cytoplasmic and sp<strong>in</strong>dle<br />

microtubuli, <strong>in</strong> meiotic cultures (Trachtulcova et al., 2000). There is no good explanation to<br />

20


clarify these contradict<strong>in</strong>g results. Isw2 is required for <strong>meiosis</strong>, cells deleted for ISW2 arrest under<br />

meiotic conditions prior to premeiotic DNA replication, and asci are not formed (Trachtulcova et<br />

al., 2000). It is not known whether Isw2 is required for the transcription <strong>of</strong> EMG under meiotic<br />

conditions, nor why cells deleted for ISW2 are sporulation deficient.<br />

B. Expression <strong>of</strong> early MSG under meiotic conditions.<br />

1. Ume6 is also a positive regulator. Ume6 is required for the transcription <strong>of</strong> EMG under<br />

meiotic conditions (Bowdish and Mitchell, 1993; Bowdish et al., 1995; Steber and Esposito,<br />

1995; Strich et al., 1994; Szent-Gyorgyi, 1995). In addition, S<strong>in</strong>3 and Rpd3 are required for their<br />

high level <strong>of</strong> expression (Lamb and Mitchell, 2001). Ume6, S<strong>in</strong>3 and Rpd3 are present and<br />

physically associate <strong>in</strong> both vegetative growth and meiotic conditions (Lamb and Mitchell, 2001).<br />

When tethered to heterologous genes, Ume6 can serve as either a transcriptional repressor or<br />

activator, depend<strong>in</strong>g on S<strong>in</strong>3/Rpd3 and Ime1, respectively (Bowdish et al., 1995; Kadosh and<br />

Struhl, 1997). Different doma<strong>in</strong>s <strong>in</strong> Ume6 are required for transcriptional repression and<br />

activation (Bowdish et al., 1995; Washburn and Esposito, 2001). The Ume6T99N mutant prote<strong>in</strong><br />

is competent <strong>in</strong> repression <strong>of</strong> EMG <strong>in</strong> growth conditions, but is defective <strong>in</strong> their expression<br />

under meiotic conditions (Bowdish et al., 1995). On the other hand, the M530T mutation <strong>in</strong><br />

UME6 has a defect only <strong>in</strong> repression (Washburn and Esposito, 2001). In agreement, the<br />

transcriptional activator, Ime1 associates with Ume6(1-232) (Rub<strong>in</strong>-Bejerano et al., 1996), while<br />

the transcriptional repressor, S<strong>in</strong>3, with Ume6(515-530) (Kadosh and Struhl, 1997; Kasten et al.,<br />

1997; Washburn and Esposito, 2001). The positive and negative roles <strong>of</strong> Ume6 are mediated<br />

through its b<strong>in</strong>d<strong>in</strong>g to the URS1 element whose presence is required for complete expression <strong>of</strong><br />

EMG <strong>in</strong> starved cells (Bowdish et al., 1995).<br />

2. The function <strong>of</strong> Ime1. IME1 encodes as positive regulator <strong>of</strong> <strong>meiosis</strong>. Over-expression<br />

<strong>of</strong> Ime1 bypasses the requirement for MATa1 and MATα2 gene products for sporulation (Kassir<br />

et al., 1988). Diploid cells deleted for IME1 arrest under meiotic conditions <strong>in</strong> G1, as unbudded<br />

cells (Foiani et al., 1996; Kassir et al., 1988), prior to execution <strong>of</strong> any meiotic event, namely,<br />

expression <strong>of</strong> <strong>meiosis</strong>-specific genes, premeiotic DNA replication, commitment to meiotic<br />

recomb<strong>in</strong>ation, meiotic divisions, and spore formation (Foiani et al., 1996; Kassir et al., 1988;<br />

Mitchell et al., 1990; Smith and Mitchell, 1989). The requirement <strong>of</strong> IME1 for transcription<br />

suggests that it might encode a transcription factor. However, its predicted am<strong>in</strong>o acid sequence<br />

does not show any homology to known DNA-b<strong>in</strong>d<strong>in</strong>g motifs (Sherman et al., 1993; Smith et al.,<br />

21


1993). Nevertheless, when Ime1 is tethered to heterologous genes it functions as a potent<br />

transcriptional activator (Mandel et al., 1994; Smith et al., 1993).<br />

As described above, the transcription <strong>of</strong> IME1 is subject to extensive <strong>regulation</strong> by the MAT<br />

alleles and nutrients. On top <strong>of</strong> it, translation <strong>of</strong> IME1 mRNA is regulated by nutrients (Sherman<br />

et al., 1993). In vegetative growth media with acetate as the sole carbon source low but<br />

substantial levels <strong>of</strong> IME1 or ime1-lacZ mRNAs are observed, but Ime1-lacZ prote<strong>in</strong> is not<br />

detected. On the other hand, upon nitrogen depletion, the level <strong>of</strong> IME1 mRNA is not <strong>in</strong>duced <strong>in</strong><br />

MATa/MATa cells, but Ime1-lacZ prote<strong>in</strong> is readily observed (Sherman et al., 1993). Furthermore,<br />

α-factor and heat-shock treatment <strong>in</strong>creases the transcription <strong>of</strong> IME1, but translation occurs only<br />

<strong>in</strong> cells arrested <strong>in</strong> G1 by either α-factor, or the CDC mutations cdc28-4 and cdc4-3 (Sherman et<br />

al., 1993). These results suggest that nitrogen and/or cell cycle progression <strong>in</strong>hibits translation <strong>of</strong><br />

IME1 mRNA. S<strong>in</strong>ce nitrogen depletion leads to a G1 arrest, it is possible that the effect <strong>of</strong><br />

nitrogen is <strong>in</strong>direct, and that a G1 arrest is a prerequisite for efficient translation. IME1 has an<br />

atypical 5 UTR (untranslated region), 229 bp long (Sherman et al., 1993), which might mediate<br />

this <strong>regulation</strong>. Sequence analysis reveals that this RNA region can form stem-and-loop<br />

secondary structure that might <strong>in</strong>hibit translation. However, deletion <strong>of</strong> this region has no effect<br />

on the translation <strong>of</strong> IME1 mRNA (Ben- Dov, 1994). Thus, it is not known how nitrogen and/or<br />

G1 phase controls the efficiency <strong>of</strong> translation <strong>of</strong> IME1 mRNA.<br />

Over expression <strong>of</strong> Ime1 <strong>in</strong> acetate growth media does not <strong>in</strong>duce <strong>meiosis</strong> and sporulation <strong>in</strong><br />

logarithmic cultures <strong>of</strong> wild-type diploids (Colom<strong>in</strong>a et al., 1999; Sherman et al., 1993). On the<br />

other hand, <strong>in</strong> diploid cells arrested <strong>in</strong> G1 by recessive temperature sensitive mutations <strong>in</strong><br />

CDC28, CDC4, CDC25, CDC35 (CYR1), or concomitant deletion <strong>of</strong> the three CLN genes, high<br />

percentages <strong>of</strong> asci are formed (Colom<strong>in</strong>a et al., 1999; Sherman et al., 1993; Shilo et al., 1978),<br />

suggest<strong>in</strong>g that post-translational modification <strong>of</strong> Ime1 or another factor required for <strong>meiosis</strong><br />

depends on lack <strong>of</strong> function <strong>of</strong> these prote<strong>in</strong>s. Indeed, <strong>in</strong> vegetative cultures, depend<strong>in</strong>g on the<br />

Cdc28/Cln function, Ime1 is phosphorylated and sequestered from the nucleus (Colom<strong>in</strong>a et al.,<br />

1999). The localization <strong>of</strong> Ime1 <strong>in</strong> the cytoplasm prevents its transcriptional activation function,<br />

and entry <strong>in</strong>to <strong>meiosis</strong>. Cdc4 is an F-box prote<strong>in</strong> [for review see (Patton et al., 1998)] required for<br />

degradation <strong>of</strong> specific targets <strong>in</strong> G1. Interest<strong>in</strong>gly, one <strong>of</strong> its substrates is Far1, an <strong>in</strong>hibitor <strong>of</strong><br />

Cln/Cdc28 function (Henchoz et al., 1997). Thus, the effect <strong>of</strong> Cdc4 on sporulation may be<br />

mediated through Cdc28/Cln function. The role <strong>of</strong> Cdc25 and Cdc35, the two positive regulators<br />

<strong>of</strong> PKA on the function <strong>of</strong> Ime1 is most probably through their effect on the association <strong>of</strong> Ime1<br />

with Ume6 (see below).<br />

22


It is not clear if the activity <strong>of</strong> Ime1 as a transcriptional activator is subject to <strong>regulation</strong>. Us<strong>in</strong>g<br />

the SK1 stra<strong>in</strong> and a LexA-Ime1 fusion, it was shown that transcriptional activation <strong>of</strong> the<br />

reporter gene lexAop-lacZ depends on nitrogen depletion and the presence <strong>of</strong> a prote<strong>in</strong> k<strong>in</strong>ase,<br />

Rim11 (Smith et al., 1993). Genetic analysis showed that <strong>in</strong> this system Rim11 was required to<br />

relieve a repression activity <strong>of</strong> Ime1 that was modulated by the C-term<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> Ime1. This<br />

conclusion was based on the follow<strong>in</strong>g results: i. A lexA-Ime1 fusion truncated for the C-term<strong>in</strong>al<br />

66 am<strong>in</strong>o acids activated transcription <strong>of</strong> lexAop-lacZ <strong>in</strong> rim11∆ cells (Smith et al., 1993). ii.<br />

LexA-Ime1L321F mutant prote<strong>in</strong> is impaired <strong>in</strong> both association with Rim11 and transcriptional<br />

activation (Malathi et al., 1997). On the other hand, us<strong>in</strong>g the S288C stra<strong>in</strong> and a Gal4(bd)-Ime1<br />

fusion, it was shown that transcriptional activation <strong>of</strong> the reporter gene gal1-lacZ is <strong>in</strong>dependent<br />

<strong>of</strong> growth conditions or Rim11 (Mandel et al., 1994; Rub<strong>in</strong>-Bejerano et al., 1996). In both stra<strong>in</strong>s<br />

Rim11 is required for the transcription <strong>of</strong> EMG and sporulation (Mandel et al., 1994; Mitchell<br />

and Bowdish, 1992). The reasons for the disparity between these reports are not known, but<br />

several possibilities can be suggested. i. The use <strong>of</strong> different stra<strong>in</strong> backgrounds, S288C and SK1.<br />

ii. The b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> the Gal4(bd) and lexA to the DNA requires its dimerization. The gal4(bd)-<br />

IME1 gene <strong>in</strong>cludes the Gal4 dimerization signal, whereas the lexA-IME1 gene might lack the<br />

lexA dimerization sequence (Golemis and Brent, 1992). Under meiotic conditions Ime1 can<br />

oligomerize, and this activity depends on Rim11 (Rub<strong>in</strong>-Bejerano et al., 1996). Therefore, it is<br />

possible that the ability <strong>of</strong> the lexA-Ime1 prote<strong>in</strong> to activate transcription only under meiotic<br />

conditions and only <strong>in</strong> the RIM11 stra<strong>in</strong> is due to the ability <strong>of</strong> the prote<strong>in</strong> to dimerize and b<strong>in</strong>d<br />

the DNA only under these conditions.<br />

Deletion and mutation analysis reveals that Ime1 is composed <strong>of</strong> at least two doma<strong>in</strong>s essential<br />

for <strong>meiosis</strong>: a transcriptional activation doma<strong>in</strong> (ad) (am<strong>in</strong>o acids 165-228), and an <strong>in</strong>teraction<br />

doma<strong>in</strong> (id) (am<strong>in</strong>o acids 270-360) (Mandel et al., 1994; Smith et al., 1993). The N-term<strong>in</strong>al 160<br />

am<strong>in</strong>o acids are not essential for <strong>meiosis</strong> (Mandel et al., 1994). However, an Ime1 prote<strong>in</strong><br />

truncated for this doma<strong>in</strong> gives rise to lower levels <strong>of</strong> asci <strong>in</strong> comparison to cells express<strong>in</strong>g this<br />

truncated prote<strong>in</strong> fused to the Gal4(bd), suggest<strong>in</strong>g that the later might either provide a nuclear<br />

localization signal or <strong>in</strong>crease the stability <strong>of</strong> the prote<strong>in</strong> (Mandel et al., 1994).<br />

Two hybrid assays reveal that Ime1 <strong>in</strong>teracts with Ume6, and that this <strong>in</strong>teraction is through<br />

am<strong>in</strong>o acids 270-360 <strong>of</strong> Ime1 [Ime1(id)] and am<strong>in</strong>o acids 1-232 <strong>of</strong> Ume6 [Ume6(id)] (Colom<strong>in</strong>a<br />

et al., 1999; Rub<strong>in</strong>-Bejerano et al., 1996; Xiao and Mitchell, 2000). The validity <strong>of</strong> this<br />

<strong>in</strong>teraction is evident from the use <strong>of</strong> different stra<strong>in</strong> backgrounds and different two-hybrid<br />

systems, namely, the Gal4(bd)-Ime1(id) with Ume6(id)-Gal4(ad) (Rub<strong>in</strong>-Bejerano et al., 1996;<br />

Xiao and Mitchell, 2000), and the tetR-Ime1(id) with Ume6(id)-VP16 (Colom<strong>in</strong>a et al., 1999).<br />

23


However, direct demonstration <strong>of</strong> physical association between Ime1 and Ume6 is still miss<strong>in</strong>g.<br />

The use <strong>of</strong> the two-hybrid assay enables the demonstration that this <strong>in</strong>teraction is negatively<br />

regulated by both glucose and nitrogen. The <strong>in</strong>teraction is absent <strong>in</strong> vegetative growth conditions<br />

with glucose as the sole carbon source, and excessive <strong>in</strong>teraction takes place under meiotic<br />

conditions (SPM media) (Colom<strong>in</strong>a et al., 1999; Rub<strong>in</strong>-Bejerano et al., 1996). Shift<strong>in</strong>g glucose<br />

grown stationary cells to acetate growth media leads to partial <strong>in</strong>teraction (Rub<strong>in</strong>-Bejerano et al.,<br />

1996), suggest<strong>in</strong>g that the <strong>in</strong>teraction <strong>of</strong> Ime1 with Ume6 depends on the absence <strong>of</strong> both glucose<br />

and nitrogen. Furthermore, this <strong>in</strong>teraction is absolutely dependent on Rim11 (Colom<strong>in</strong>a et al.,<br />

1999; Rub<strong>in</strong>-Bejerano et al., 1996), and partially dependent on Rim15 (Vidan and Mitchell,<br />

1997).<br />

Two models were suggested for the role <strong>of</strong> Ime1 Ume6 and Rim11 <strong>in</strong> the transcription <strong>of</strong><br />

EMG. I. Bowdish et al proposed that Ume6 is converted from a repressor to an activator<br />

follow<strong>in</strong>g phosphorylation by Rim11, and that Ime1, which associates with Rim11 (Bowdish et<br />

al., 1994; Rub<strong>in</strong>-Bejerano et al., 1996), is required to recruit Rim11 to Ume6 (Bowdish et al.,<br />

1995). ii. Rub<strong>in</strong>-Bejerano et al proposed that Ume6 recruits Ime1 to the URS1 element present <strong>in</strong><br />

EMG, and that this recruitment promotes the Ime1 transcriptional activation doma<strong>in</strong> to <strong>in</strong>duce the<br />

transcription <strong>of</strong> EMG. Accord<strong>in</strong>g to the later, Rim11 is required for the association between Ime1<br />

and Ume6 (Rub<strong>in</strong>-Bejerano et al., 1996). The follow<strong>in</strong>g results support the second model: i. Ime1<br />

functions as a transcriptional activator (Mandel et al., 1994; Smith et al., 1993). ii. Fusion <strong>of</strong> an<br />

heterologous transcriptional activation doma<strong>in</strong>, that <strong>of</strong> Gal4, to Ime1(id), leads to the expression<br />

<strong>of</strong> the EMG, IME2, and sporulation (Mandel et al., 1994). Moreover, this Ime1(id)-Gal4(ad)<br />

fusion prote<strong>in</strong> promotes <strong>meiosis</strong> <strong>of</strong> ime1∆ diploid cells only when galactose is added to the<br />

sporulation media (Mandel et al., 1994). S<strong>in</strong>ce galactose is required to relieve repression <strong>of</strong> Gal80<br />

on the transcriptional activation <strong>of</strong> Gal4(ad) (Johnston and Carlson, 1992), it was suggested that<br />

the normal function <strong>of</strong> Ime1 is to activate transcription (Mandel et al., 1994). iii. A Gal4(ad)-<br />

Ume6(159-836) fusion prote<strong>in</strong> that is expressed from the IME1 promoter bypasses the<br />

requirement for IME1 for the transcription <strong>of</strong> EMG and sporulation: it leads to expression <strong>of</strong> the<br />

EMG, HOP1, and to 40% sporulation (Rub<strong>in</strong>-Bejerano et al., 1996). iv. A Gal4(ad)-Ume6 fusion<br />

prote<strong>in</strong> can activate the transcription <strong>of</strong> the EMG SPO13 <strong>in</strong> SD media if it carries mutations that<br />

prevent its association with S<strong>in</strong>3 (Washburn and Esposito, 2001). In the absence <strong>of</strong> the Gal4(ad),<br />

these ume6 mutants do not promote expression <strong>of</strong> EMG (Washburn and Esposito, 2001). These<br />

results suggest that by itself Ume6 does not function as a transcriptional activator, and its<br />

conversion <strong>in</strong>to a positive regulator depends on the recruitment <strong>of</strong> the transcriptional activation<br />

doma<strong>in</strong> <strong>of</strong> Ime1 (Rub<strong>in</strong>-Bejerano et al., 1996; Washburn and Esposito, 2001). Currently, this is<br />

24


the established model, however, the ultimate pro<strong>of</strong> would be the demonstration that Ime1 is<br />

present on the complex formed on URS1, and/or that it associates with the transcription<br />

mach<strong>in</strong>ery. Both models assume that s<strong>in</strong>ce Ime1 functions through Ume6, deletions <strong>of</strong> either one<br />

<strong>of</strong> these genes would have a similar phenotype <strong>in</strong> meiotic cultures. However, ime1∆ diploids<br />

arrest <strong>in</strong> G1, whereas ume6∆ cells ma<strong>in</strong>ly arrest at G2. We assume that the expression <strong>of</strong> EMG <strong>in</strong><br />

vegetative cultures promotes the entry and progression through some meiotic events.<br />

Ime1 has at least two functions, it supplies a transcriptional activation doma<strong>in</strong>, but it is also<br />

required to relieve repression <strong>of</strong> S<strong>in</strong>3/Rpd3 (Washburn and Esposito, 2001). Diploid cells deleted<br />

for IME1 and express<strong>in</strong>g Gal4(ad)-Ume6 from the ADH1 promoter are sporulation deficient<br />

(Washburn and Esposito, 2001), but the Gal4(ad)-Ume6(M530T) mutant prote<strong>in</strong> promotes low<br />

levels <strong>of</strong> sporulation (10%). S<strong>in</strong>ce the latter prote<strong>in</strong> is defective <strong>in</strong> association with S<strong>in</strong>3, it was<br />

concluded that Ime1 is required to relieve repression mediated by the S<strong>in</strong>3/Rpd3 complex<br />

(Washburn and Esposito, 2001).<br />

The <strong>in</strong>teraction between Ime1 and Ume6 is regulated by nutrients, Rim11 and Rim15. S<strong>in</strong>ce<br />

two-hybrid assays were used to determ<strong>in</strong>e this <strong>in</strong>teraction, the regulated expression <strong>of</strong> the reporter<br />

genes might result from <strong>regulation</strong> at any <strong>of</strong> the follow<strong>in</strong>g levels: i. Regulated physical<br />

association between Ime1 and Ume6, or ii. Regulated transcriptional activation by the<br />

Ume6/Ime1 complex. The latter model assumes that Ime1 and Ume6 physically associate under<br />

all growth conditions, but <strong>in</strong> glucose growth media the result<strong>in</strong>g complex does not function as a<br />

transcriptional activator. This hypothesis is supported by the follow<strong>in</strong>g results. i. In vegetative<br />

growth media association <strong>of</strong> Ume6 with the S<strong>in</strong>3/Rpd3 complex represses transcription (Kadosh<br />

and Struhl, 1997). ii. A po<strong>in</strong>t mutation <strong>in</strong> Ume6 that prevents its association with S<strong>in</strong>3 promote<br />

transcriptional activation when the mutant prote<strong>in</strong> is fused to Gal4(ad) (Washburn and Esposito,<br />

2001). Direct demonstration <strong>of</strong> regulated (or not) physical association between Ime1 and Ume6 is<br />

required to dist<strong>in</strong>guish between these models.<br />

In accord with the role <strong>of</strong> Ime1 as a transcriptional activator, the prote<strong>in</strong> is localized to the<br />

nucleus (Colom<strong>in</strong>a et al., 1999; Rub<strong>in</strong>-Bejerano et al., 1996; Smith et al., 1993). Localization <strong>of</strong><br />

Ime1 is <strong>in</strong>dependent <strong>of</strong> Rim11 (Rub<strong>in</strong>-Bejerano et al., 1996), but as described above, it is<br />

<strong>in</strong>hibited by the Cln/Cdc28 function (Colom<strong>in</strong>a et al., 1999). Fig. 10 schematically illustrates how<br />

nutrients control the availability and activity <strong>of</strong> Ime1<br />

25


3. Prote<strong>in</strong>s required for the association <strong>of</strong> Ime1 with Ume6.<br />

3.1. The function <strong>of</strong> Rim11 and its homologs Mck1 and Mrk1. Rim11 is absolutely<br />

required for the <strong>in</strong>teraction between Ime1 and Ume6 (Colom<strong>in</strong>a et al., 1999; Rub<strong>in</strong>-Bejerano et<br />

al., 1996), the expression <strong>of</strong> EMG, and spore formation (Bowdish et al., 1994; Rub<strong>in</strong>-Bejerano et<br />

al., 1996). Rim11 associates with Ime1(id) under all growth conditions (Bowdish et al., 1994;<br />

Rub<strong>in</strong>-Bejerano et al., 1996), and <strong>in</strong> vitro it phosphorylates both Ime1 (Bowdish et al., 1994) and<br />

Ume6 (Malathi et al., 1997). The Rim11 homologs Mck1 and Mrk1 have only m<strong>in</strong>or effects on<br />

these processes (Neigeborn and Mitchell, 1991; Rabitsch et al., 2001).<br />

The transcription <strong>of</strong> RIM11 is constitutive <strong>in</strong> vegetative and meiotic conditions; however, <strong>in</strong><br />

ime1∆ cells its expression <strong>in</strong> meiotic cultures is reduced, probably reflect<strong>in</strong>g the use <strong>of</strong> the URS1<br />

like site present <strong>in</strong> its promoter (Bowdish et al., 1994). The level <strong>of</strong> Rim11 prote<strong>in</strong> is slightly<br />

reduced <strong>in</strong> glucose growth media <strong>in</strong> comparison to acetate (Bowdish et al., 1994). Rim11<br />

functions as a k<strong>in</strong>ase that shows autophosphorylation (Bowdish et al., 1994). Rim11 is<br />

phosphorylated on tyros<strong>in</strong>e 199 (Zhan et al., 2000). Tyros<strong>in</strong>e to phenylalan<strong>in</strong>e mutation <strong>in</strong> this<br />

residue results <strong>in</strong> impaired expression <strong>of</strong> ime2-lacZ <strong>in</strong> vivo, and <strong>in</strong> a defect <strong>in</strong> phosphorylation<br />

Ume6 <strong>in</strong> vitro (Zhan et al., 2000). This mutation does not impair autophosphorylation (Zhan et<br />

al., 2000), suggest<strong>in</strong>g different requirement for phosphorylation <strong>of</strong> exogenous substrates (see<br />

section IIID1).<br />

Rim11 associates with the C-term<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> Ime1, am<strong>in</strong>o acids 270-360 (Malathi et al.,<br />

1999; Rub<strong>in</strong>-Bejerano et al., 1996). In vitro, Rim11 phosphorylates the C-term<strong>in</strong>al 20 am<strong>in</strong>o<br />

acids residues <strong>in</strong> this region <strong>of</strong> Ime1 (Malathi et al., 1999). Simultaneous mutations <strong>of</strong> 8 ser<strong>in</strong>e<br />

threon<strong>in</strong>e and tyros<strong>in</strong>e <strong>of</strong> these residues have the follow<strong>in</strong>g effects: i. In vitro phosphorylation <strong>of</strong><br />

Ime1 by Rim11 is absent, ii. Ime1 does not <strong>in</strong>teract with Ume6, iii. IME2 is not expressed, and iv.<br />

Asci are not formed (Malathi et al., 1999). Ime1 is detected by antibodies directed aga<strong>in</strong>st<br />

phosphotyros<strong>in</strong>es, and tyros<strong>in</strong>e to phenylalan<strong>in</strong>e mutation <strong>in</strong> this doma<strong>in</strong> results <strong>in</strong> impaired<br />

expression <strong>of</strong> ime2-lacZ <strong>in</strong> vivo, suggest<strong>in</strong>g that Rim11 phosphorylates these tyros<strong>in</strong>e resides <strong>in</strong><br />

Ime1.<br />

Rim11 controls <strong>meiosis</strong> by regulat<strong>in</strong>g not only Ime1, but also Ume6. Us<strong>in</strong>g the two-hybrid<br />

method, coIP, and <strong>in</strong> vitro k<strong>in</strong>ase assay, Malathi et al show that Rim11 physically associates with<br />

Ume6, it phosphorylates Ume6(1-232), and the association between Rim11 and Ume6 is<br />

<strong>in</strong>dependent <strong>of</strong> Ime1 (Malathi et al., 1997). In addition, the two-hybrid assay reveals <strong>in</strong>teraction<br />

between Ume6 and Mck1 (Xiao and Mitchell, 2000). In vivo, Ume6 is a phosphoprote<strong>in</strong>, whose<br />

level <strong>of</strong> phosphorylation is <strong>in</strong>creased <strong>in</strong> vegetative growth media with acetate as the sole carbon<br />

source <strong>in</strong> comparison to glucose (Xiao and Mitchell, 2000). Moreover, hyper phosphorylation is<br />

26


observed upon nitrogen starvation (Xiao and Mitchell, 2000). S<strong>in</strong>gle deletion <strong>of</strong> RIM11, MCK1 or<br />

MRK1 has no effect on the <strong>in</strong> vivo pattern <strong>of</strong> phosphorylation <strong>of</strong> Ume6 (Xiao and Mitchell, 2000),<br />

but a substantial reduction <strong>in</strong> the pattern <strong>of</strong> phosphorylation is observed for the rim11∆ mck1∆<br />

mrk1∆ triple mutant (Xiao and Mitchell, 2000). The redundant function <strong>of</strong> these GSK3-β<br />

homologs is also evident <strong>in</strong> cells carry<strong>in</strong>g the partially active rim11K68R allele. Efficient<br />

sporulation is observed for cells carry<strong>in</strong>g the MCK1 MRK1 wild type alleles, and no sporulation<br />

<strong>in</strong> either mck1∆ MRK1 or MCK1 mrk1∆ stra<strong>in</strong>s (Xiao and Mitchell, 2000). Ume6 sequence<br />

reveals several consensus sites for phosphorylation by the mammalian GSK3-β homolog.<br />

Simultaneous substitution <strong>of</strong> ser<strong>in</strong>e and threon<strong>in</strong>e residues to alan<strong>in</strong>e <strong>in</strong> one such site (T99A<br />

T103A T107) leads to a reduction <strong>in</strong> phosphorylation, and concomitantly a dramatic reduction <strong>in</strong><br />

the ability <strong>of</strong> Ume6(1-232) to <strong>in</strong>teract with Ime1, suggest<strong>in</strong>g that phosphorylation is a prerequisite<br />

for <strong>in</strong>teraction (Xiao and Mitchell, 2000).<br />

The two-hybrid method demonstrates that the <strong>in</strong>teraction between Rim11 and Ume6 is<br />

regulated by the carbon source; the <strong>in</strong>teraction is low <strong>in</strong> glucose growth media and it is <strong>in</strong>creased<br />

<strong>in</strong> acetate growth media (Malathi et al., 1997). However, coIP shows physical association<br />

between Ume6 and Rim11 that is <strong>in</strong>dependent on nutrients (Malathi et al., 1997). The authors<br />

suggest that this is due to lower levels <strong>of</strong> Galbd-Ume6 <strong>in</strong> glucose versus acetate growth media<br />

(Malathi et al., 1997). However, it is also possible that this is due to the <strong>in</strong>ability <strong>of</strong> the Gal4(bd)-<br />

Rim11/Gal4(ad)-Ume6 complex to activate transcription, due to the activity <strong>of</strong> S<strong>in</strong>3/Rpd3.<br />

Nevertheless, the ability <strong>of</strong> Rim11 to <strong>in</strong> vitro phosphorylate Ume6 is <strong>in</strong>creased when both<br />

prote<strong>in</strong>s are isolated from acetate grown cells, suggest<strong>in</strong>g that the carbon source regulates either<br />

the activity <strong>of</strong> Rim11, or the ability <strong>of</strong> its substrate, Ume6, to be phosphorylated (Malathi et al.,<br />

1997).<br />

The association <strong>of</strong> Ime1 with Ume6, and the association <strong>of</strong> Rim11 with both Ime1 and Ume6<br />

suggest that one association may serve as a scaffold for the second association. There are no<br />

direct evidence for association <strong>of</strong> Rim11 with Ime1 and Ume6 <strong>in</strong> cells deleted for UME6 or<br />

IME1, respectively. However, a Gal4ad-Ume6T99N mutant prote<strong>in</strong> shows no <strong>in</strong>teraction with<br />

Rim11 <strong>in</strong> a two-hybrid assay, and reduced <strong>in</strong>teraction with Ime1 (Malathi et al., 1997). These<br />

results suggest that either the association between Ume6 and Rim11 is required for the <strong>in</strong>teraction<br />

between Ime1 and Ume6 (Malathi et al., 1997), or that the Ume6 T99 residue is required for the<br />

association <strong>of</strong> Ume6 with both Rim11 and Ime1.<br />

3.2. The function <strong>of</strong> Rim15. Rim15 is absolutely required for the transcription <strong>of</strong> EMG<br />

such as IME2, HOP1 and SPO13 (Vidan and Mitchell, 1997). As described above (section<br />

IIIC1.2), Rim15 is required for the transcription <strong>of</strong> IME1, however, expression <strong>of</strong> IME1 from an<br />

27


heterologous promoter does not suppress this phenotype, suggest<strong>in</strong>g that Rim15 is required for an<br />

additional event. Indeed, Rim15 is required for efficient <strong>in</strong>teraction between Ime1 and Ume6, <strong>in</strong><br />

cells deleted for RIM15 5-fold reduction <strong>in</strong> their <strong>in</strong>teraction is observed (Vidan and Mitchell,<br />

1997). This effect is probably due to the effect <strong>of</strong> Rim15 on Ume6. It is required for complete <strong>in</strong><br />

vivo phosphorylation <strong>of</strong> Ume6 <strong>in</strong> SA and SPM media (Xiao and Mitchell, 2000). Rim15 is not<br />

required for the <strong>in</strong> vitro k<strong>in</strong>ase activity <strong>of</strong> Rim11 on Ime1 (Vidan and Mitchell, 1997).<br />

4. The function <strong>of</strong> Gcn5. The conversion <strong>of</strong> URS1 from a repression to activation element<br />

requires the function <strong>of</strong> the histone acetylase, Gcn5 (Burgess et al., 1999). Diploid cells carry<strong>in</strong>g<br />

a recessive mutation <strong>in</strong> GCN5 arrest under meiotic conditions prior to premeiotic DNA<br />

replication and meiotic recomb<strong>in</strong>ation (Burgess et al., 1999). These cells express IME1, but are<br />

defective <strong>in</strong> the transcription <strong>of</strong> IME2 (Burgess et al., 1999). Us<strong>in</strong>g antibodies directed aga<strong>in</strong>st<br />

acetylated histones H3 and H4, Burges et al., show that under meiotic conditions, depend<strong>in</strong>g on<br />

Gcn5, specific hyper acetylation on histone H3 is observed <strong>in</strong> the IME2 promoter (Burgess et al.,<br />

1999). Deletion <strong>of</strong> RPD3 that is required for deacetylation <strong>of</strong> Histone H4 <strong>in</strong> the IME2 promote,<br />

does not suppress gcn5, suggest<strong>in</strong>g that acetylation <strong>of</strong> histones is a prerequisite for the<br />

transcription <strong>of</strong> EMG (Burgess et al., 1999). It is not known how Gcn5 is recruited to the<br />

promoter <strong>of</strong> IME2 or additional EMG.<br />

5. The transcription <strong>of</strong> EMG is also regulated by positive elements. In addition to the<br />

negative element, URS1, the promoters <strong>of</strong> EMG carry positive elements required for their<br />

transcription under meiotic conditions: UASH element <strong>in</strong> HOP1, and the T4C sequence <strong>in</strong> IME2<br />

(Bowdish and Mitchell, 1993; Vershon et al., 1992). Homologous sequences are found <strong>in</strong><br />

additional EMG (Chu et al., 1998). Deletion or mutations <strong>of</strong> these elements prevent high-level<br />

expression <strong>of</strong> EMG (Bowdish and Mitchell, 1993; Vershon et al., 1992). Abf1 b<strong>in</strong>ds to the UASH<br />

element (Gailus-Durner et al., 1996). ABF1 encodes an essential DNA-b<strong>in</strong>d<strong>in</strong>g transcriptional<br />

activator required for the transcription <strong>of</strong> various genes as well as for DNA replication; it b<strong>in</strong>ds to<br />

orig<strong>in</strong>s <strong>of</strong> DNA replication (Svetlov and Cooper, 1995). The prote<strong>in</strong> that b<strong>in</strong>ds the T4C element <strong>in</strong><br />

IME2 is not known.<br />

6. Additional positive regulators. Ime2, Rim4, and Ndt80 are early and early late genes<br />

required for high-level transcription <strong>of</strong> EMG. Detailed description on their functions is given <strong>in</strong><br />

section VA.<br />

28


7. The role <strong>of</strong> premeiotic DNA replication and/or recomb<strong>in</strong>ation <strong>in</strong> controll<strong>in</strong>g EMG<br />

expression. Hydroxyurea (HU) is rout<strong>in</strong>ely used to <strong>in</strong>hibit DNA synthesis as it <strong>in</strong>hibits<br />

ribonucleotide reductase (Elford, 1968; Slater, 1973). Shift<strong>in</strong>g cells to meiotic conditions <strong>in</strong> the<br />

presence <strong>of</strong> 40 mM and 200 mM hydroxyurea leads to a reduction or no expression, respectively,<br />

<strong>of</strong> EMG (Davis et al., 2001; Lamb and Mitchell, 2001). It was suggested that perturbation <strong>of</strong><br />

premeiotic DNA replication <strong>in</strong>hibits the transcription <strong>of</strong> EMG. However, cells deleted for CLB5<br />

along with CLB6 or MUM2/SPOT8 arrest prior to premeiotic DNA replication, with complete<br />

expression <strong>of</strong> EMG (Davis et al., 2001; Dirick et al., 1998; Stuart and Wittenberg, 1998). These<br />

results imply that HU blocks DNA replication at different po<strong>in</strong>t than clb5 clb6 and mum2, and<br />

that only the HU block transmits a checkpo<strong>in</strong>t signal to repress transcription. However, it is also<br />

possible, that the effect <strong>of</strong> HU is not mediated through DNA replication. The latter hypothesis is<br />

strengthened by the observation that treatment <strong>of</strong> <strong>yeast</strong> cells with HU leads to a substantial<br />

decrease <strong>in</strong> RNA and prote<strong>in</strong> synthesis (Slater, 1973).<br />

The effect <strong>of</strong> HU is mediated through Rpd3 and S<strong>in</strong>3 whose presence is required for the<br />

reduction <strong>in</strong> transcription (Lamb and Mitchell, 2001). In addition, <strong>in</strong> the presence <strong>of</strong> HU there is a<br />

reduction <strong>in</strong> phosphorylation <strong>of</strong> Ume6, a phenomenon that is associated with reduction <strong>in</strong> the<br />

activity <strong>of</strong> Ume6 (Lamb and Mitchell, 2001).<br />

8. The choice between silenc<strong>in</strong>g and expression <strong>of</strong> EMG. Silenc<strong>in</strong>g and expression <strong>of</strong><br />

EMG are dependent on the URS1 element present <strong>in</strong> their promoter. URS1 serves as a silenc<strong>in</strong>g<br />

element <strong>in</strong> vegetative growth media with glucose as the sole carbon source, and is converted <strong>in</strong>to<br />

an activation element under meiotic conditions, i.e. nitrogen depletion <strong>in</strong> the presence <strong>of</strong> acetate.<br />

Fig. 11 summarizes the current knowledge and model for how this <strong>regulation</strong> is accomplished.<br />

Under all growth conditions Ume6 b<strong>in</strong>ds to URS1. However, nutrients control the prote<strong>in</strong><br />

complexes that are recruited by Ume6 to the URS1 element. In glucose growth media Ume6<br />

recruits two repression complexes, the HDAC S<strong>in</strong>3/Rpd3 that leads to deacetylation <strong>of</strong> lys<strong>in</strong>es <strong>in</strong><br />

histone H4, and the Isw2 complex that leads to chromat<strong>in</strong> remodel<strong>in</strong>g. These two activities are<br />

required for complete silenc<strong>in</strong>g <strong>of</strong> EMG. The S<strong>in</strong>3/Rpd3 complex is conserved, and functions as a<br />

transcriptional silencer <strong>in</strong> all eukaryotes (Ng and Bird, 2000; Paz<strong>in</strong> and Kadonaga, 1997).<br />

However, Ume6 is not conserved, and <strong>in</strong> mammals, it is the Mad/Max and nuclear receptors that<br />

recruit S<strong>in</strong>3/Rpd3 to the DNA (Ng and Bird, 2000; Paz<strong>in</strong> and Kadonaga, 1997). Similarly, the<br />

Isw2 complex is also conserved and functional <strong>in</strong> all eukaryotes [for review see (Cairns, 1998)].<br />

Under meiotic conditions (acetate media and nitrogen depletion) the S<strong>in</strong>3/Rpd3 and Isw2<br />

repression complexes are not functional. Relief <strong>of</strong> repression <strong>of</strong> S<strong>in</strong>3 depends on Ime1 that is<br />

29


expressed only <strong>in</strong> the absence <strong>of</strong> glucose. The mode by which Ime1 relieves repression <strong>of</strong><br />

S<strong>in</strong>3/Rpd3 and how Isw2 does not repress transcription under these conditions are not known.<br />

Relief <strong>of</strong> repression does not suffice for transcriptional activation, and requires Gcn5, the histone<br />

acetylase, as well as a transcriptional activation prote<strong>in</strong>, Ime1. Ume6 recruits Ime1 to URS1, and<br />

the <strong>in</strong>teraction between these two prote<strong>in</strong>s depends on the absence <strong>of</strong> glucose and nitrogen. These<br />

nutrient signals are transmitted to both Ime1 and Ume6, by two prote<strong>in</strong> k<strong>in</strong>ases, Rim15 and<br />

Rim11. S<strong>in</strong>ce Rim15 is required for complete phosphorylation <strong>of</strong> Ume6 <strong>in</strong> SA and SPM, and<br />

s<strong>in</strong>ce its k<strong>in</strong>ase activity is <strong>in</strong>hibited <strong>in</strong> the presence <strong>of</strong> glucose by PKA, it is assumed that the<br />

glucose signal that <strong>in</strong>hibits the association <strong>of</strong> Ime1 with Ume6 is transmitted, at least partially,<br />

through Rim15. S<strong>in</strong>ce Rim15 is only partially required for the <strong>in</strong>teraction <strong>of</strong> Ime1 with Ume6, the<br />

glucose signal must be transmitted through an additional prote<strong>in</strong>. Rim11 phosphorylates both<br />

Ime1 and Ume6, and this phosphorylation is essential for their <strong>in</strong>teraction. It is not known how<br />

nutrients regulate the function <strong>of</strong> Rim11.<br />

V. <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> middle <strong>meiosis</strong>-specific genes (MMG).<br />

The regulated transcription <strong>of</strong> middle <strong>meiosis</strong>-specific genes (MMG) depends on the presence <strong>of</strong><br />

two positive elements, MSE (Middle Sporulation Element) (gNCRCAAAA/T) and an Abf1<br />

b<strong>in</strong>d<strong>in</strong>g site (Chu et al., 1998; Chu and Herskowitz, 1998; Hepworth et al., 1995; Hepworth et al.,<br />

1998; Ozsarac et al., 1995; Ozsarac et al., 1997; Pierce et al., 1998). The MSE elements present <strong>in</strong><br />

different MMG are not identical, some, for example that <strong>in</strong> CLB1, function only as positive<br />

elements, whereas others, for example that <strong>of</strong> SMK1 or NDT80 (designated <strong>in</strong>here as MSE*),<br />

function also as repression elements <strong>in</strong> vegetative growth conditions and early meiotic times<br />

(Pierce et al., 1998; Xie et al., 1999). Schematic illustration on the <strong>regulation</strong> <strong>of</strong> MMG is<br />

illustrated <strong>in</strong> Fig. 13, and discussed <strong>in</strong> the text below.<br />

A. Positive regulators <strong>of</strong> MMG.<br />

Ndt80 and Ime2 are two <strong>meiosis</strong>-specific positive regulators absolutely required for the<br />

transcription <strong>of</strong> MMG (Chu et al., 1998; Chu and Herskowitz, 1998; Hepworth et al., 1998;<br />

Ozsarac et al., 1997). The sequential transcription <strong>of</strong> MMG follow<strong>in</strong>g transcription <strong>of</strong> the early<br />

genes is due to the regulated transcription <strong>of</strong> both NDT80 and IME2 (Chu and Herskowitz, 1998;<br />

Hepworth et al., 1998).<br />

30


1. Ndt80. NDT80 encodes a transcriptional activator that b<strong>in</strong>ds to the MSE and MSE* elements<br />

and activates transcription (Chu and Herskowitz, 1998). Cells deleted for NDT80 arrest follow<strong>in</strong>g<br />

the completion <strong>of</strong> premeiotic DNA replication and meiotic recomb<strong>in</strong>ation, at pachytene, as mono<br />

nucleate cells (Chu and Herskowitz, 1998; Hepworth et al., 1998; Xu et al., 1995). Ectopic<br />

expression <strong>of</strong> Ndt80 <strong>in</strong> vegetative growth conditions is sufficient to <strong>in</strong>duce the transcription <strong>of</strong><br />

most MMG (probably the ones carry<strong>in</strong>g the MSE rather than the MSE* element) (Chu et al.,<br />

1998; Chu and Herskowitz, 1998), suggest<strong>in</strong>g that under growth conditions, the second positive<br />

regulator, Ime2, is non-essential. Ime2 is also non-essential under meiotic conditions. When<br />

Ndt80 is expressed from a heterologous, IME2 <strong>in</strong>dependent promoter, the MMG, SPS4, is<br />

expressed <strong>in</strong> both IME2 and ime2∆ cells (Pak and Segall, 2002a). Nevertheless, <strong>in</strong> the latter,<br />

expression <strong>of</strong> SPS4 is both delayed and lower than <strong>in</strong> the IME2 diploid cells, suggest<strong>in</strong>g that Ime2<br />

may be required for high and efficient activity <strong>of</strong> Ndt80 (Pak and Segall, 2002a). This hypothesis<br />

is supported by the observation that Ime2 phosphorylates Ndt80 (Benjam<strong>in</strong> et al., 2002). Ndt80 is<br />

a phosphoprote<strong>in</strong> (Benjam<strong>in</strong> et al., 2002; Tung et al., 2000), whose <strong>in</strong> vivo extent <strong>of</strong><br />

phosphorylation depends on IME2 (Benjam<strong>in</strong> et al., 2002). Furthermore, <strong>in</strong> vitro k<strong>in</strong>ase assay<br />

demonstrates that Ime2 directly phosphorylates Ndt80 (Benjam<strong>in</strong> et al., 2002).<br />

Ndt80 is also required for high levels <strong>of</strong> transcription <strong>of</strong> IME1 as well as for its transient<br />

expression (Hepworth et al., 1998). This effect might be mediated through Ime2, s<strong>in</strong>ce Ndt80<br />

regulates the transcription <strong>of</strong> IME2, through an MSE element present <strong>in</strong> its promoter (Chu and<br />

Herskowitz, 1998).<br />

NDT80 is a <strong>meiosis</strong>-specific gene whose transcription is dependent on Ime1 and Ime2 (Fig. 13)<br />

(Chu and Herskowitz, 1998; Hepworth et al., 1998; Pak and Segall, 2002a). This regulated<br />

transcription is mediated by two URS1 and two MSE elements (Chu and Herskowitz, 1998;<br />

Hepworth et al., 1998; Pak and Segall, 2002a). One MSE element functions either as a repression<br />

or activation element under growth and meiotic condition, respectively, whereas the second one<br />

functions only as an activation element (Pak and Segall, 2002a; Xie et al., 1999). Furthermore,<br />

both URS1 elements contribute to repression <strong>of</strong> NDT80 <strong>in</strong> vegetative cultures and for its high<br />

level expression under meiotic conditions [(Pak and Segall, 2002a) and Fig. 13]. The presence <strong>of</strong><br />

the MSE elements delays NDT80 transcription by the Ume6/Ime1 complex <strong>in</strong> comparison to<br />

other EMG that carry only the URS1 element (Chu and Herskowitz, 1998; Hepworth et al., 1998;<br />

Mitchell et al., 1990; Pak and Segall, 2002a; Yoshida et al., 1990).<br />

2. Ime2. IME2 encodes a <strong>meiosis</strong>-specific prote<strong>in</strong> k<strong>in</strong>ase that shows 58.9% similarity and<br />

37% identity to the human cycl<strong>in</strong>-dependent k<strong>in</strong>ase, CDK2 (Kom<strong>in</strong>ami et al., 1993). The crystal<br />

31


structure <strong>of</strong> hCDK2 reveals that cycl<strong>in</strong> A b<strong>in</strong>ds to two regions: PSTAIRE and the T-loop (Jeffrey<br />

et al., 1995). This b<strong>in</strong>d<strong>in</strong>g is required to <strong>in</strong>duce a conformational change that promotes activation<br />

<strong>of</strong> the k<strong>in</strong>ase. Ime2 does not carry the PSTAIRE sequence, but it shows 67% similarity and<br />

52.3% identity to the T-loop region <strong>of</strong> hCDK2 (Fig. 12). This homology raises the possibility that<br />

Ime2 might be activated by b<strong>in</strong>d<strong>in</strong>g to a cycl<strong>in</strong> like molecule, and that <strong>in</strong> the meiotic cycle it<br />

might replace Cdc28, the <strong>yeast</strong> CDK that regulates <strong>in</strong>itiation and progression <strong>in</strong> the cell cycle<br />

(Lew et al., 1997). The follow<strong>in</strong>g results support the hypothesis that <strong>in</strong> the meiotic cycle Ime2<br />

might replace Cdc28: i. Similarly to Cdc28, phosphorylation by Ime2 affects stability <strong>of</strong> Ime1<br />

Cdh1 and Sic1, the latter two prote<strong>in</strong>s are known substrates <strong>of</strong> Cdc28 <strong>in</strong> the mitotic cell cycle<br />

(Bolte et al., 2002; Dirick et al., 1998; Guttmann-Raviv and Kassir, 2002). ii. Ime2 is absolutely<br />

required for premeiotic DNA replication and meiotic recomb<strong>in</strong>ation <strong>in</strong> cdc28-4 cells <strong>in</strong>cubated at<br />

the non-permissive temperature (Guttmann-Raviv et al., 2001). However, there are no reports on<br />

any <strong>yeast</strong> prote<strong>in</strong>s that b<strong>in</strong>d to the T-loop like sequence <strong>in</strong> Ime2 and are required for its k<strong>in</strong>ase<br />

activity. Moreover, a recomb<strong>in</strong>ant Ime2 prote<strong>in</strong> isolated from E. coli is active <strong>in</strong> phosphorylat<strong>in</strong>g<br />

histone H1 (Donzeau and Bandlow, 1999), suggest<strong>in</strong>g that unlike Cdc28, Ime2 is active <strong>in</strong> the<br />

absence <strong>of</strong> additional <strong>yeast</strong> prote<strong>in</strong>s. However, phosphorylation <strong>of</strong> its native substrate, Gpa2 (see<br />

below) requires the isolation <strong>of</strong> Ime2 from <strong>yeast</strong> cells (Donzeau and Bandlow, 1999), suggest<strong>in</strong>g<br />

that specificity, for at least for some substrates, might require post-translation modification or the<br />

presence <strong>of</strong> an activator.<br />

Ime2 is a positive regulator <strong>of</strong> <strong>meiosis</strong> required for multiple functions: the correct tim<strong>in</strong>g and<br />

high level transcription <strong>of</strong> early <strong>meiosis</strong>-specific genes, the transcription <strong>of</strong> middle and late genes<br />

(Mitchell et al., 1990; Yoshida et al., 1990), the correct tim<strong>in</strong>g <strong>of</strong> entry <strong>in</strong>to premeiotic DNA<br />

replication meiotic recomb<strong>in</strong>ation and nuclear division (Benjam<strong>in</strong> et al., 2002; Foiani et al.,<br />

1996), as well as for ascus formation (Benjam<strong>in</strong> et al., 2002; Foiani et al., 1996; Mitchell et al.,<br />

1990; Yoshida et al., 1990). When Ime2 is over expressed, but only <strong>in</strong> the SK1 background, it<br />

bypasses the requirement for Ime1 for the transcription <strong>of</strong> EMG and sporulation (Mitchell et al.,<br />

1990), suggest<strong>in</strong>g an Ime1 <strong>in</strong>dependent pathway for activat<strong>in</strong>g the transcription <strong>of</strong> these genes<br />

(Mitchell et al., 1990). It is not known how Ime2, that encodes a k<strong>in</strong>ase, replaces a transcriptional<br />

activator, nor if under normal conditions this pathway has any contribution to the transcription <strong>of</strong><br />

these genes.<br />

Ime2 functions also as a negative regulator for the follow<strong>in</strong>g functions: i. Phosphorylation <strong>of</strong><br />

Ime1 by Ime2 leads to its degradation by the 26S proteasome (Guttmann-Raviv and Kassir,<br />

2002). It is possible that the absence <strong>of</strong> Ime1 reestablishes repression by Sok2 and S<strong>in</strong>3 [see<br />

sections IIIC1.1.2, IVB1.2, and (Shenhar and Kassir, 2001; Washburn and Esposito, 2001)],<br />

32


esult<strong>in</strong>g <strong>in</strong> the transient transcription <strong>of</strong> IME1 and EMG. Accord<strong>in</strong>g to this model, <strong>in</strong> cells<br />

deleted for IME2, the cont<strong>in</strong>uous presence <strong>of</strong> Ime1 leads to the non-transient expression <strong>of</strong> IME1<br />

and EMG (Mitchell et al., 1990; Yoshida et al., 1990). ii. In the meiotic cycle, upon completion <strong>of</strong><br />

premeiotic DNA replication and the presence <strong>of</strong> Ime2, two subunits <strong>of</strong> the DNA polymerase αprimase<br />

complex, Pol1 and Pol12, are degraded (Foiani et al., 1996). iii. The level <strong>of</strong> the<br />

Clb/Cdc28 <strong>in</strong>hibitor, Sic1, is reduced under meiotic conditions, a process regulated by Ime2<br />

(Dirick et al., 1998). iv. When Ime2 is ectopically expressed <strong>in</strong> the cell cycle it phosphorylates<br />

Cdh1, an event caus<strong>in</strong>g dissociation <strong>of</strong> Cdh1 from APC (anaphase promot<strong>in</strong>g complex – the<br />

ubiquit<strong>in</strong> ligase required for proteolysis <strong>of</strong> specific substrates), and consequently stabilization <strong>of</strong><br />

its substrates, such as Clb2 and Cdc5 (Bolte et al., 2002). It is not known if Ime2 phosphorylate<br />

Cdh1 <strong>in</strong> the meiotic cycle, and what are the consequences <strong>of</strong> this event for entry <strong>in</strong>to the meiotic<br />

divisions. v. In the meiotic cycle Ime2 is required to limit premeiotic DNA replication and<br />

nuclear division to one and two rounds, respectively (Foiani et al., 1996). It is assumed that this<br />

phenomenon results from stabilization <strong>of</strong> specific prote<strong>in</strong>s or regulators <strong>of</strong> DNA replication and<br />

nuclear divisions.<br />

The availability and function <strong>of</strong> Ime2 is regulated <strong>in</strong> several levels:<br />

<strong>Transcriptional</strong> <strong>regulation</strong>: The transcription <strong>of</strong> IME2 is silent <strong>in</strong> vegetative growth conditions by<br />

the b<strong>in</strong>d<strong>in</strong>g and activity <strong>of</strong> the Ume6/S<strong>in</strong>3/Rpd3 and Ume6/Isw2 complexes to the URS1 element<br />

present <strong>in</strong> its promoter. Under meiotic condition, recruitment <strong>of</strong> Ime1 by Ume6 promotes the<br />

transcription <strong>of</strong> Ime2. An additional histone deacetylase activity, that <strong>of</strong> the Set3/Hos2 complex,<br />

regulates the transcription <strong>of</strong> IME2 as well as NDT80 early <strong>in</strong> <strong>meiosis</strong> (Pijnappel et al., 2001).<br />

Deletion <strong>of</strong> SET3 or HOS2 has no effect dur<strong>in</strong>g vegetative growth (although this might be due to<br />

repression by S<strong>in</strong>3/Rpd3), but it leads to <strong>in</strong>creased and advanced transcription <strong>of</strong> IME2 and<br />

NDT80, as well as moderate <strong>in</strong>crease <strong>in</strong> the transcription <strong>of</strong> IME1 (Pijnappel et al., 2001).<br />

Consequently, these cells progress faster through MI, MII and asci formation (Pijnappel et al.,<br />

2001). Ime2 is subject to positive auto<strong>regulation</strong> that is most probably mediated by the MSE<br />

element present <strong>in</strong> its promoter and Ndt80 (see above, VA1).<br />

Prote<strong>in</strong> stability: Ime2 is an extremely non-stable prote<strong>in</strong> (Bolte et al., 2002; Guttmann-Raviv and<br />

Kassir, 2002) with a half-life <strong>of</strong> about 5 m<strong>in</strong>utes (Guttmann-Raviv and Kassir, 2002). It is not<br />

known if the stability <strong>of</strong> Ime2 is subject to any <strong>regulation</strong>.<br />

Activity <strong>regulation</strong>: Ime2 is subject to phosphorylation (Benjam<strong>in</strong> et al., 2002; Guttmann-Raviv<br />

and Kassir, 2002). In vitro k<strong>in</strong>ase assays reveals autophosphorylation (Benjam<strong>in</strong> et al., 2002;<br />

Guttmann-Raviv and Kassir, 2002; Kom<strong>in</strong>ami et al., 1993), but its effect on the activity and/or<br />

stability <strong>of</strong> Ime2 is not known. The k<strong>in</strong>ase activity <strong>of</strong> Ime2 is negatively regulated by its C-<br />

33


term<strong>in</strong>al non-essential doma<strong>in</strong> (Kom<strong>in</strong>ami et al., 1993). This is deduced from the follow<strong>in</strong>g<br />

observations: i. Over expression <strong>of</strong> an Ime2 prote<strong>in</strong> truncated for this doma<strong>in</strong> promotes <strong>meiosis</strong> <strong>in</strong><br />

the presence <strong>of</strong> either nitrogen or glucose (Kom<strong>in</strong>ami et al., 1993), and ii. The <strong>in</strong> vitro-k<strong>in</strong>ase<br />

activity <strong>of</strong> Ime2 is higher for a truncated Ime2 prote<strong>in</strong> <strong>in</strong> comparison to the wt prote<strong>in</strong> (Kom<strong>in</strong>ami<br />

et al., 1993). Thus, <strong>in</strong> the presence <strong>of</strong> nutrients Ime2 is less, or non-active. The nutrient signal is<br />

transmitted to the C-term<strong>in</strong>al doma<strong>in</strong> through the Gα prote<strong>in</strong>, Gpa2 (Donzeau and Bandlow,<br />

1999). Ime2 specifically associates with the GTP bound Gpa2 (Donzeau and Bandlow, 1999), a<br />

form whose level is <strong>in</strong>creased <strong>in</strong> the presence <strong>of</strong> glucose [for review see (Versele et al., 2001)]. In<br />

addition, the association between Gpa2 and Ime2 requires the presence <strong>of</strong> nitrogen (Donzeau and<br />

Bandlow, 1999). Cells deleted for GPA2 show similar phenotypes to cells over express<strong>in</strong>g the<br />

truncated Ime2 prote<strong>in</strong>, namely, sporulation <strong>in</strong> the presence <strong>of</strong> glucose and nitrogen (Donzeau<br />

and Bandlow, 1999). Gpa2 associates with the glucose receptor, Gpr1, and transmits the glucose<br />

signal to adenylate cyclase [for review see (Pan et al., 2000)], however, the effect <strong>of</strong> Gpa2 on<br />

Ime2 is <strong>in</strong>dependent <strong>of</strong> PKA (Donzeau and Bandlow, 1999). This is deduced from the<br />

observation that the bcy1-tpk1 w1 mutation that leads to low activity <strong>of</strong> PKA does not suppress the<br />

reduction <strong>in</strong> sporulation exhibited by cells express<strong>in</strong>g the constitutive active Gpa2G132V prote<strong>in</strong><br />

(Donzeau and Bandlow, 1999). The <strong>in</strong> vitro k<strong>in</strong>ase activity <strong>of</strong> Ime2 on histone H1 is decreased by<br />

the addition <strong>of</strong> recomb<strong>in</strong>ant Gpa2γS (Donzeau and Bandlow, 1999), suggest<strong>in</strong>g that Gpa2 <strong>in</strong>hibits<br />

the k<strong>in</strong>ase activity <strong>of</strong> Ime2 (Donzeau and Bandlow, 1999). In vitro k<strong>in</strong>ase assays also demonstrate<br />

that Ime2 phosphorylates Gpa2 (Donzeau and Bandlow, 1999), but the role <strong>of</strong> this<br />

phosphorylation is not known.<br />

Ime2 k<strong>in</strong>ase activity fluctuates <strong>in</strong> the meiotic cycle, it first peaks concomitantly with<br />

premeiotic DNA replication and then, concomitantly with nuclear division (Benjam<strong>in</strong> et al.,<br />

2002). The second <strong>in</strong>crease <strong>in</strong> activity is regulated by Cdc28 (Benjam<strong>in</strong> et al., 2002).<br />

The activity <strong>of</strong> Ime2 is also regulated by Ids2 (Sia and Mitchell, 1995). Over expression <strong>of</strong><br />

Ime2 <strong>in</strong> vegetative growth media is toxic (Bolte et al., 2002; Guttmann-Raviv and Kassir, 2002;<br />

Sia and Mitchell, 1995), and this toxicity is relived <strong>in</strong> cells deleted for IDS2 (Sia and Mitchell,<br />

1995), suggest<strong>in</strong>g that Ids2 is a positive regulator <strong>of</strong> Ime2. Ids2 is not required for <strong>meiosis</strong> (Sia<br />

and Mitchell, 1995), however, <strong>in</strong> its absence over expression <strong>of</strong> Ime2 (<strong>in</strong> the SK1 stra<strong>in</strong>) leads to<br />

the transcription <strong>of</strong> only the early <strong>meiosis</strong>-specific genes, while middle and late genes rema<strong>in</strong><br />

silent and spores are not formed (Sia and Mitchell, 1995). This result suggests that the activity <strong>of</strong><br />

Ime2 is dist<strong>in</strong>ctly regulated at early and middle meiotic times (Sia and Mitchell, 1995), <strong>in</strong><br />

agreement with the f<strong>in</strong>d<strong>in</strong>g <strong>of</strong> Benjam<strong>in</strong> et al., (Benjam<strong>in</strong> et al., 2002) reported above. The effect<br />

<strong>of</strong> Ids2 is mediated through the C-term<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> Ime2: Over expression <strong>of</strong> Ime2 truncated<br />

34


for this doma<strong>in</strong> suppresses ime1∆ <strong>in</strong> both wt and ids2∆ cells (Sia and Mitchell, 1995). The steadystate<br />

level <strong>of</strong> Ids2 is decreased and then disappears <strong>in</strong> cells shifted to meiotic conditions (Sia and<br />

Mitchell, 1995).<br />

RIM4 is an EMG required for high-level expression <strong>of</strong> EMG, premeiotic DNA replication,<br />

timely and efficient commitment to meiotic recomb<strong>in</strong>ation, nuclear division, and spore formation<br />

(Deng and Saunders, 2001; Soushko and Mitchell, 2000). Rim4 function is mediated through<br />

both the Ime1 and Ime2 transcriptional activation pathways (Soushko and Mitchell, 2000). rim4∆<br />

diploids over-express<strong>in</strong>g Ime2 are sporulation pr<strong>of</strong>icient, suggest<strong>in</strong>g that Rim4 functions<br />

upstream <strong>of</strong> Ime2 (Soushko and Mitchell, 2000). Rim4 carries two RNA recognition motifs that<br />

are important for its role <strong>in</strong> <strong>meiosis</strong> (Soushko and Mitchell, 2000). The mode by which Rim4<br />

affects transcription is not known, though it was proposed that it might stabilize IME2 mRNA<br />

(Deng and Saunders, 2001; Soushko and Mitchell, 2000).<br />

3. S<strong>in</strong>3 and Rpd3. Random screen<strong>in</strong>g for mutations prevent<strong>in</strong>g expression <strong>of</strong> MMG<br />

identified NDT80 and SIN3 (Hepworth et al., 1998). S<strong>in</strong>3 as well as Rpd3 are required for the<br />

transcription <strong>of</strong> MMG, suggest<strong>in</strong>g that they function also as positive regulators, or that their<br />

function as positive regulators might be due to repression <strong>of</strong> a negative regulator (Hepworth et al.,<br />

1998). Diploid cells deleted for SIN3 or RPD3 arrest at the same po<strong>in</strong>t as ndt80 cells, namely,<br />

follow<strong>in</strong>g the completion <strong>of</strong> premeiotic DNA replication, as mono nucleate cells (Hepworth et al.,<br />

1998). Po<strong>in</strong>t <strong>of</strong> arrest is not due to defects <strong>in</strong> recomb<strong>in</strong>ation, s<strong>in</strong>ce concomitant deletion <strong>of</strong> SPO13<br />

and SPO11 does not allow a bypass <strong>of</strong> the arrest, and the triple mutant cells rema<strong>in</strong> sporulation<br />

deficient (Hepworth et al., 1998; Xu et al., 1995).<br />

B. Negative regulators <strong>of</strong> MMG.<br />

SMK1 is a middle gene subject to silenc<strong>in</strong>g by b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> Sum1 to the MSE* element <strong>in</strong> its<br />

promoter (Xie et al., 1999). Sum1 also represses the transcription <strong>of</strong> NDT80, <strong>in</strong> cells deleted for<br />

SUM1 premature expression <strong>of</strong> NDT80 is observed (Pak and Segall, 2002a). Sum1 associates<br />

with Hst1, and the result<strong>in</strong>g complex, conta<strong>in</strong><strong>in</strong>g additional prote<strong>in</strong>s, exhibits NAD + dependent<br />

histone deacetylase activity (Pierce et al., 1998; Pijnappel et al., 2001; Rusche and R<strong>in</strong>e, 2001).<br />

Deletion <strong>of</strong> either SUM1 or HST1 leads to expression <strong>of</strong> SMK1 <strong>in</strong> vegetative growth conditions<br />

(Xie et al., 1999). This expression is probably <strong>in</strong>dependent <strong>of</strong> Ndt80, s<strong>in</strong>ce NDT80 rema<strong>in</strong>s silent<br />

<strong>in</strong> sum1∆ or hst1∆ cells (Xie et al., 1999). Relief <strong>of</strong> repression is regulated by Sum1 availability:<br />

The transcription <strong>of</strong> SUM1 is constitutive, but the level <strong>of</strong> Sum1 prote<strong>in</strong> is reduced prior to<br />

expression <strong>of</strong> MMG, and then it is <strong>in</strong>duced aga<strong>in</strong>. As described above, the transcription <strong>of</strong> NDT80<br />

35


is dependent on Ime2, but <strong>in</strong> cells deleted for both IME2 and SUM1, NDT80 is expressed (Pak<br />

and Segall, 2002a). These results suggest that Ime2 abrogates Sum1 repression (Pak and Segall,<br />

2002a). It is possible, as <strong>in</strong> the case <strong>of</strong> Ime1 (Guttmann-Raviv and Kassir, 2002), that<br />

phosphorylation <strong>of</strong> Sum1 by Ime2 is required for the regulated degradation <strong>of</strong> Sum1. The<br />

Sum1/Hst1 repression activity can be bypassed <strong>in</strong> vegetative growth media by over expression <strong>of</strong><br />

Ndt80 (Xie et al., 1999). Sum1 and Hst1 are not required for <strong>meiosis</strong>, cells deleted for either gene<br />

complete <strong>meiosis</strong> and form viable spores (L<strong>in</strong>dgren et al., 2000).<br />

The S<strong>in</strong>3/Rpd3/Ume6 and Ssn6/Tup1 repression complexes that regulate transcription <strong>of</strong> EMG<br />

are not <strong>in</strong>volved with repression activity <strong>of</strong> MMG (Pierce et al., 1998).<br />

C. The recomb<strong>in</strong>ation checkpo<strong>in</strong>t.<br />

In <strong>meiosis</strong>, a checkpo<strong>in</strong>t mechanism consist<strong>in</strong>g <strong>of</strong> Rad17, Rad24, Mek1, Pch2, Mec3, and Ddc1<br />

prevents the transcription <strong>of</strong> MMG and entry <strong>in</strong>to the first meiotic division <strong>in</strong> cells that have not<br />

properly completed synapsis <strong>of</strong> homologs and meiotic recomb<strong>in</strong>ation (Roeder and Bailis, 2000).<br />

Diploid cells deleted for DMC1 are impaired <strong>in</strong> both meiotic recomb<strong>in</strong>ation (Bishop et al., 1992)<br />

and expression <strong>of</strong> MMG (Chu and Herskowitz, 1998; Hepworth et al., 1998), whereas dmc1 cells<br />

carry<strong>in</strong>g mutations <strong>in</strong> RAD17, or MEK1 express MMG (Chu and Herskowitz, 1998; Hepworth et<br />

al., 1998). Three targets <strong>of</strong> the pachytene checkpo<strong>in</strong>t were identified; these are the prote<strong>in</strong> k<strong>in</strong>ase<br />

Swe1 that phosphorylates and <strong>in</strong>activates Cdc28 (Leu and Roeder, 1999), the transcriptional<br />

activator Ndt80 (Pak and Segall, 2002b; Tung et al., 2000), and the transcriptional repressor<br />

Sum1 (L<strong>in</strong>dgren et al., 2000; Pak and Segall, 2002b). This is concluded from the observations<br />

that <strong>in</strong> the double mutants dmc1 swe1 and dmc1 sum1 MMG are expressed and cells enter the<br />

meiotic nuclear division (Leu and Roeder, 1999; L<strong>in</strong>dgren et al., 2000; Pak and Segall, 2002b;<br />

Tung et al., 2000). Over expression <strong>of</strong> Ndt80 leads to expression <strong>of</strong> MMG and partial entry <strong>in</strong>to<br />

nuclear division (Pak and Segall, 2002b). In this review we focus only on how this surveillance<br />

mechanism affects transcription [for a detailed review on the checkpo<strong>in</strong>t pathway see (Roeder<br />

and Bailis, 2000)].<br />

The pachytene checkpo<strong>in</strong>t leads to a reduction <strong>in</strong> the transcription <strong>of</strong> NDT80 (L<strong>in</strong>dgren et<br />

al., 2000; Pak and Segall, 2002b). However, contradict<strong>in</strong>g results, namely, no effect, were<br />

reported for other different stra<strong>in</strong>s (Chu and Herskowitz, 1998; Hepworth et al., 1998; Tung et al.,<br />

2000). Sum1 mediates the reduction <strong>in</strong> the transcription <strong>of</strong> NDT80. The availability <strong>of</strong> Sum1 is<br />

regulated by the pachytene checkpo<strong>in</strong>t as evident from the observations that <strong>in</strong> dmc1 cells the<br />

steady-state level <strong>of</strong> Sum1 is constitutive throughout the meiotic pathway, whereas <strong>in</strong> the double<br />

mutant dmc1 rad17 the level <strong>of</strong> Sum1 is reduced <strong>in</strong> meiotic prophase, as <strong>in</strong> the wild type stra<strong>in</strong><br />

36


(L<strong>in</strong>dgren et al., 2000). S<strong>in</strong>ce Ime2 mediates the availability <strong>of</strong> Sum1 under normal meiotic<br />

conditions, it is possible that the effect <strong>of</strong> the checkpo<strong>in</strong>t system on Sum1 is mediated through<br />

Ime2. Pak and Segall (Pak and Segall, 2002b) suggest that the effect <strong>of</strong> Sum1 is mediated only<br />

through regulat<strong>in</strong>g the transcription <strong>of</strong> NDT80, s<strong>in</strong>ce the triple mutant dmc1 sum1 ndt80 does not<br />

enter nuclear division (Pak and Segall, 2002b). Morovere, they propose that <strong>in</strong> the dmc1 sum1<br />

cells, expression <strong>of</strong> the B-type cycl<strong>in</strong>s is responsible for entry <strong>in</strong>to meiotic nuclear division (Pak<br />

and Segall, 2002b). This hypothesis is supported by the observation that entry <strong>in</strong>to nuclear<br />

division <strong>in</strong> swe1 hop2 cells is delayed <strong>in</strong> comparison to wild type cells or to swe1 hop2 cells over<br />

express<strong>in</strong>g Clb1 (Leu and Roeder, 1999). However, this model cannot expla<strong>in</strong> why ectopic<br />

expression <strong>of</strong> Ndt80 <strong>in</strong> dmc1 cells leads to <strong>in</strong>efficient entry <strong>in</strong>to nuclear division <strong>in</strong> comparison to<br />

the dmc1 sum1 cells (20% versus 60%) (Pak and Segall, 2002b). These results suggest that an<br />

additional target <strong>of</strong> Sum1 might be required for proper meiotic arrest (L<strong>in</strong>dgren et al., 2000).<br />

S<strong>in</strong>ce the CLB genes are not direct targets <strong>of</strong> Sum1, they are not the genes directly regulated by<br />

Sum1 (L<strong>in</strong>dgren et al., 2000).<br />

The pachytene checkpo<strong>in</strong>t <strong>in</strong>hibits phosphorylation <strong>of</strong> Ntd80 (Tung et al., 2000). In cells<br />

deleted for ZIP1 [a component <strong>of</strong> the synaptonemal complex required for synapsis <strong>of</strong> homologs<br />

(Sym et al., 1993)] Ndt80 is partially phosphorylated (Tung et al., 2000), whereas <strong>in</strong> the double<br />

mutants zip1 pch2 and zip1 mek1 it is highly phosphorylated (Tung et al., 2000). It was suggested<br />

that activation <strong>of</strong> transcription by Ndt80 requires its phosphorylation (Benjam<strong>in</strong> et al., 2002; Chu<br />

and Herskowitz, 1998; Hepworth et al., 1998; Tung et al., 2000). Therefore, lack <strong>of</strong> expression <strong>of</strong><br />

MMG is also due to the reduction <strong>in</strong> the phosphorylation <strong>of</strong> Ndt80. S<strong>in</strong>ce Ime2 is required for<br />

phosphorylation <strong>of</strong> Ndt80, it should be <strong>in</strong>terest<strong>in</strong>g to determ<strong>in</strong>e if Ime2 is the target for the<br />

pachytene check po<strong>in</strong>t system.<br />

In addition, the pachytene checkpo<strong>in</strong>t mediates the transcription <strong>of</strong> MMG through the Cdc28<br />

<strong>in</strong>hibitor, Swe1. In swe1 hop2 diploid cells CLB1 transcription is <strong>in</strong>creased to its normal level,<br />

although with a substantial delay, <strong>in</strong> comparison to wild type or hop2 cells (Leu and Roeder,<br />

1999). It should be <strong>in</strong>terest<strong>in</strong>g to determ<strong>in</strong>e whether this effect <strong>of</strong> Swe1 on Cdc28 is mediated<br />

through Ime2, s<strong>in</strong>ce Cdc28 is required for the activity <strong>of</strong> Ime2 that co<strong>in</strong>cides with nuclear<br />

divisions (Benjam<strong>in</strong> et al., 2002), and Ime2 is absolutely required for the transcription <strong>of</strong> the CLB<br />

genes.<br />

37


VI. <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> late <strong>meiosis</strong>-specific genes<br />

A. Early late genes.<br />

SGA1 is an early-late gene encod<strong>in</strong>g a glucoamylase whose transcription <strong>in</strong> growth media is<br />

silenced by a negative element, NRE SGA , which exhibits no UAS activity (Kihara et al., 1991;<br />

Mai and Breeden, 2000). Expression under meiotic conditions depends on a UAS element present<br />

<strong>in</strong> its promoter region (Kihara et al., 1991). In addition, it carries 4 STRE elements (Mai and<br />

Breeden, 2000), but their role <strong>in</strong> the transcription <strong>of</strong> SGA1 is not known. Relief <strong>of</strong> repression by<br />

the NRE SGA element depends on Ime1 and Ime2. However, the activity <strong>of</strong> the UAS element is<br />

<strong>in</strong>dependent <strong>of</strong> these regulators, because it does not require the presence <strong>of</strong> both MATa and<br />

MATα alleles that are required for the expression <strong>of</strong> Ime1 (Kihara et al., 1991). The activity <strong>of</strong> the<br />

positive element requires nitrogen depletion (Kihara et al., 1991), demonstrat<strong>in</strong>g that the effect <strong>of</strong><br />

nitrogen on <strong>meiosis</strong> is mediated not only through Ime1.<br />

Silenc<strong>in</strong>g <strong>of</strong> SGA1 <strong>in</strong> vegetative growth media depends on the chromat<strong>in</strong>-remodel<strong>in</strong>g factor<br />

Isw2 (Goldmark et al., 2000). As described above (section IVA5), Isw2 is also a negative<br />

regulator <strong>of</strong> EMG, and it is recruited by Ume6 to their promoters. S<strong>in</strong>ce the URS1 element to<br />

which Ume6 b<strong>in</strong>ds is not present <strong>in</strong> SGA1, it is not known how Isw2 is recruited to SGA1. The<br />

identities <strong>of</strong> the positive regulators promot<strong>in</strong>g transcriptional activation <strong>of</strong> SGA1 are not known.<br />

However, one positive regulator, SME2, when present on a multi-copy plasmid, <strong>in</strong>creases the<br />

transcription <strong>of</strong> SGA1 <strong>in</strong> the presence <strong>of</strong> nitrogen, thus bypass<strong>in</strong>g also the nitrogen signal<br />

<strong>in</strong>hibit<strong>in</strong>g spore formation (Kawaguchi et al., 1992). Deletion <strong>of</strong> SME2 has no effect on the<br />

regulated expression <strong>of</strong> SGA1, or on the efficiency <strong>of</strong> sporulation. The sequence and mode <strong>of</strong><br />

function <strong>of</strong> SME2 are not known.<br />

B. Mid-late genes. The transcription <strong>of</strong> the mid-late sporulation-specific genes DIT1 and DIT2<br />

is <strong>in</strong>duced upon completion <strong>of</strong> meiotic divisions (Friesen et al., 1997). Time <strong>of</strong> expression is<br />

compatible with function, s<strong>in</strong>ce these divergently transcribed genes (Friesen et al., 1997) encode<br />

enzymes required for biosynthesis <strong>of</strong> the dityros<strong>in</strong>e precursor that is <strong>in</strong>corporated <strong>in</strong>to the<br />

outermost layer <strong>of</strong> the spore wall (Briza et al., 1990). A negative element, designated NRE DIT is<br />

required for repression dur<strong>in</strong>g vegetative growth (Friesen et al., 1997) (Note that NRE DIT is not<br />

identical to NRE SGA1 ). This region carries the middle sporulation-like element MSE, and the DRE<br />

element (Bogengruber et al., 1998; Friesen et al., 1997) that are required for repression <strong>in</strong> growth<br />

conditions and activation under meiotic conditions. Two additional positive elements are required<br />

for high expression under meiotic conditions (Friesen et al., 1997).<br />

38


<strong>Transcriptional</strong> activation <strong>of</strong> the mid late genes depends on three positive regulators, Ndt80<br />

and Ime2 whose effects are most probably mediated through the MSE like sequence (Friesen et<br />

al., 1997), and Rim1 (Rim101) through an unidentified region (Bogengruber et al., 1998). The<br />

effect <strong>of</strong> Rim1 may be <strong>in</strong>direct (see IIIC3). The repression activity <strong>of</strong> NRE DIT depends on Ssn6,<br />

Tup1, Rox3, S<strong>in</strong>4 and Spe3 (Friesen et al., 1997; Friesen et al., 1998). Rox3 and S<strong>in</strong>4 are<br />

subunits <strong>of</strong> the mediator complex <strong>of</strong> RNA polymerase II (Gustafsson et al., 1997; Myer and<br />

Young, 1998) that repress transcription <strong>of</strong> many genes (Hanna-Rose and Hansen, 1996). SPE3<br />

encodes Spermid<strong>in</strong>e synthase (Hamasaki-Katagiri et al., 1997), and accord<strong>in</strong>gly, addition <strong>of</strong><br />

spermid<strong>in</strong>e to the media leads to <strong>in</strong>creased repression (Friesen et al., 1998). The way by which<br />

these complexes are specifically recruited to NRE DIT is not known.<br />

XBP1 is a mid-late gene whose transcription is <strong>in</strong>itiated at the same time as that <strong>of</strong> DIT1,2, but<br />

unlike these genes its transcription is non-transient (Mai and Breeden, 2000). Accord<strong>in</strong>gly, its<br />

regulated transcription is not mediated via a NRE DIT element (Mai and Breeden, 2000). XBP1<br />

encodes a DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> that functions as a repressor when fused to lexA. Deletion <strong>of</strong><br />

XBP1 leads to a delay and reduction <strong>in</strong> the percentage <strong>of</strong> asci, while meiotic divisions are<br />

properly controlled (Mai and Breeden, 2000). It is not known how this effect <strong>of</strong> Xbp1 is<br />

mediated.<br />

C. Late genes<br />

The transcription <strong>of</strong> the 5 late meiotic genes peaks follow<strong>in</strong>g completion <strong>of</strong> meiotic divisions at<br />

8-12 hours <strong>in</strong> sporulation media (Law and Segall, 1988). A representative <strong>of</strong> these genes is<br />

SPS100 that is <strong>in</strong>volved <strong>in</strong> spore wall formation (Law and Segall, 1988). The transcription <strong>of</strong><br />

SPS100 depends on Ime1, Ndt80, and Ama1 (Cooper et al., 2000; Hepworth et al., 1998). AMA1<br />

encodes a <strong>meiosis</strong>-specific subunit <strong>of</strong> APC/C that is required for degradation <strong>of</strong> CLB1 <strong>in</strong> the<br />

meiotic cycle (Cooper et al., 2000). The transcription <strong>of</strong> AMA1 is <strong>in</strong>duced at the same time as<br />

MMG, but s<strong>in</strong>ce it does not carry the Ndt80 bid<strong>in</strong>g site (Chu et al., 1998) its mode <strong>of</strong> <strong>regulation</strong> is<br />

not known. In addition, <strong>meiosis</strong>-specific splic<strong>in</strong>g by the EMG Mer1 (Engebrecht and Roeder,<br />

1990) determ<strong>in</strong>es the accumulation <strong>of</strong> Ama1 prote<strong>in</strong> (Cooper et al., 2000). Diploid cells deleted<br />

for AMA1 arrest with a short sp<strong>in</strong>dle prior to the first meiotic division (Cooper et al., 2000).<br />

Genetic analysis suggests that Ama1 is dispensable for the second meiotic division, but is<br />

required for spore formation (Cooper et al., 2000). Further work is required to determ<strong>in</strong>e if Ama1<br />

is directly <strong>in</strong>volved with transcriptional <strong>regulation</strong> <strong>of</strong> late genes, for example by degradation <strong>of</strong> a<br />

transcriptional repressor, or if the effect on transcription is mediated by a checkpo<strong>in</strong>t mechanism.<br />

39


The transcriptional activator that b<strong>in</strong>ds to and activates transcription <strong>of</strong> the late genes is not<br />

known.<br />

The late <strong>meiosis</strong>-specific genes SPS100 and DTT1 show basal levels <strong>of</strong> expression <strong>in</strong><br />

vegetative growth conditions (Jung and Lev<strong>in</strong>, 1999). This expression depends on a functional<br />

MAPK cascade that transmits the cell wall <strong>in</strong>tegrity signal (Jung and Lev<strong>in</strong>, 1999). Activation <strong>of</strong><br />

the MAPK Slt2/Mpk1, or deletion <strong>of</strong> Rlm1, the transcription factor whose activity is regulated by<br />

it, reduces the expression the late <strong>meiosis</strong> specific genes (Jung and Lev<strong>in</strong>, 1999). It is not known<br />

whether Rlm1 b<strong>in</strong>ds the promoter region <strong>of</strong> late genes, or if it is required for their expression<br />

under meiotic conditions.<br />

VII. A feedback loop controll<strong>in</strong>g <strong>meiosis</strong>.<br />

The transcription <strong>of</strong> all <strong>meiosis</strong>-specific genes is transient, reflect<strong>in</strong>g the need for the function <strong>of</strong><br />

these genes for a limited period. Nevertheless, it is important to po<strong>in</strong>t that transient transcription<br />

does not necessarily reflect transient availability <strong>of</strong> prote<strong>in</strong>s. For <strong>in</strong>stance, <strong>in</strong> the mitotic cell<br />

cycle, the transcription <strong>of</strong> genes <strong>in</strong>volved <strong>in</strong> DNA replication, such as CLB5, POL1, or POL12 is<br />

periodic, and identically regulated (Lowndes et al., 1991; Schwob and Nasmyth, 1993). But the<br />

steady state level <strong>of</strong> Pol1 and Pol12 is constant (Foiani et al., 1995), whereas the level <strong>of</strong> Clb5 is<br />

periodic (Irniger and Nasmyth, 1997). Periodicity <strong>of</strong> Clb5 is accomplished by its regulated<br />

degradation (Irniger and Nasmyth, 1997). The pattern <strong>of</strong> prote<strong>in</strong> expression <strong>of</strong> many <strong>meiosis</strong>specific<br />

genes is not known, however, the two ma<strong>in</strong> positive regulators <strong>of</strong> <strong>meiosis</strong>, Ime1 and<br />

Ime2 are non-stable prote<strong>in</strong>s whose prote<strong>in</strong> levels mimic the level <strong>of</strong> their RNA. This suggests<br />

that at least <strong>in</strong> the case <strong>of</strong> Ime1 and Ime2, the transient availability might be a sign <strong>of</strong> requirement<br />

for only a short period. In agreement with this view, over expression <strong>of</strong> Ime1 <strong>in</strong> meiotic cultures<br />

leads (<strong>in</strong> some stra<strong>in</strong>s) to <strong>in</strong>efficient sporulation, a substantial <strong>in</strong>crease <strong>in</strong> non-disjunction, the<br />

formation <strong>of</strong> 2-spored rather than 4-spored asci, and a reduction <strong>in</strong> sporulation (Shefer-Vaida et<br />

al., 1995; Sherman, 1992).<br />

Positive feedback loops control the transcription <strong>of</strong> both IME1 and IME2 (Fig. 14). Positive<br />

auto<strong>regulation</strong> by Ime1 is required to relieve repression mediated by Sok2, and thus lead to an<br />

<strong>in</strong>crease <strong>in</strong> Ime1 level (Shenhar and Kassir, 2001). Positive auto<strong>regulation</strong> accelerates the<br />

availability <strong>of</strong> Ime1 and Ime2, and this enhanced expression might be essential for proper entry<br />

and progression through the meiotic cycle.<br />

The transient transcription <strong>of</strong> all <strong>meiosis</strong>-specific genes is expla<strong>in</strong>ed by the transient<br />

availability <strong>of</strong> Ime1, Ime2, and Ndt80. Phosphorylation <strong>of</strong> Ime1 by Ime2 leads to its degradation<br />

40


y the 26S proteasome (Guttmann-Raviv and Kassir, 2002). In the absence <strong>of</strong> Ime1, Sok2<br />

represses the transcription <strong>of</strong> IME1, lead<strong>in</strong>g to a decrease <strong>in</strong> IME1 mRNA. It is not known<br />

whether the negative feedback <strong>regulation</strong> <strong>of</strong> Ime2 on the transcription <strong>of</strong> IME1 is mediated only<br />

through its effect on Ime1. S<strong>in</strong>ce the transcription <strong>of</strong> IME2 depends on Ime1 (Mitchell et al.,<br />

1990; Yoshida et al., 1990) and the half-life <strong>of</strong> Ime2 is very short (Bolte et al., 2002; Guttmann-<br />

Raviv and Kassir, 2002), there are sufficient levels <strong>of</strong> Ime1 to relieve repression by S<strong>in</strong>3<br />

(Washburn and Esposito, 2001) and to activate transcription <strong>of</strong> EMG. In the absence <strong>of</strong> Ime1 lack<br />

<strong>of</strong> these two functions leads to shutt<strong>in</strong>g down the transcription <strong>of</strong> EMG. The transcription <strong>of</strong><br />

MMG and LMG depends on Ime2 (Chu et al., 1998; Chu and Herskowitz, 1998; Friesen et al.,<br />

1997; Hepworth et al., 1998; Kihara et al., 1991; Ozsarac et al., 1997; Smith and Mitchell, 1989).<br />

In cells depleted for Ime1, Ime2 is absent, and the transcription <strong>of</strong> the early late and late genes is<br />

closed.<br />

VIII. Conclud<strong>in</strong>g remark. The choice between developmental pathways<br />

S. cerevisiae is a simple eukaryote with limited developmental options. Diploid cells may chose<br />

between growth <strong>in</strong> <strong>yeast</strong> or pseudohyphal forms, and <strong>meiosis</strong>. Two crucial mechanisms regard<strong>in</strong>g<br />

developmental decisions are required for faithful growth, entry <strong>in</strong>to a specific developmental<br />

pathway should take place only <strong>in</strong> response to specific signals, and should be a signal to block<br />

entry <strong>in</strong>to an alternative pathway. Concomitant entry <strong>in</strong>to two developmental pathways could lead<br />

to chaos. In S. cerevisiae, the MATa and MATα gene products as well as the pachytene<br />

surveillance mechanisms ensure that only diploid cells can <strong>in</strong>itiate <strong>meiosis</strong>. This is a crucial<br />

decision, because <strong>meiosis</strong> <strong>in</strong> haploid cells would lead to the formation <strong>of</strong> uneuploid gametes.<br />

In S. cerevisiae, the meiotic cycle ends with the formation <strong>of</strong> dormant spores resistant to<br />

hazardous conditions. The decision to exit the mitotic cell cycle depends on the availability <strong>of</strong><br />

nutrients. Entry <strong>in</strong>to the meiotic cycle <strong>in</strong> the presence <strong>of</strong> nutrients can be a “waste”. Several<br />

redundant mechanisms ensure that <strong>in</strong> the presence <strong>of</strong> glucose the meiotic pathway will be<br />

repressed: i. The transcription <strong>of</strong> IME1 is repressed <strong>in</strong> the presence <strong>of</strong> glucose due to the presence<br />

<strong>of</strong> dist<strong>in</strong>ct negative elements (i.e. UASru, IREu, UCS1), each respond<strong>in</strong>g to a different signal<br />

pathway. Thus, if one signal pathway is malfunction<strong>in</strong>g, the activity <strong>of</strong> the additional signal<br />

pathways will ensure repression. Moreover, several UAS elements (i.e. UASru, IREu, UASrm),<br />

each regulated <strong>in</strong> a dist<strong>in</strong>ct manner, activates transcription <strong>in</strong> the absence <strong>of</strong> glucose. ii. The<br />

activity <strong>of</strong> Ime1 is repressed <strong>in</strong> the presence <strong>of</strong> glucose by two dist<strong>in</strong>ct mechanisms. a.<br />

Phosphorylation <strong>of</strong> Ime1 by the Cln/Cdc28 k<strong>in</strong>ase sequesters the prote<strong>in</strong> from the nucleus, and b.<br />

41


Functional <strong>in</strong>teraction between Ime1 and Ume6 that promote transcriptional activation is<br />

<strong>in</strong>hibited <strong>in</strong> the presence <strong>of</strong> glucose and nitrogen. The presence <strong>of</strong> nitrogen also prevents<br />

translation <strong>of</strong> IME1 mRNA and entry <strong>of</strong> Ime1 <strong>in</strong>to the nucleus.<br />

The presence <strong>of</strong> glucose also <strong>in</strong>hibits the function <strong>of</strong> Ime2. Normally Ime2 is available only<br />

under meiotic conditions, but cells developed mechanisms to ensure that <strong>in</strong> the presence <strong>of</strong><br />

nutrients accidental expression <strong>of</strong> Ime2 will not lead to entry and completion <strong>of</strong> <strong>meiosis</strong>. In the<br />

presence <strong>of</strong> glucose activated Gpa2 b<strong>in</strong>ds to and <strong>in</strong>hibits the function <strong>of</strong> Ime2. Furthermore, Ime2<br />

is a non-stabile prote<strong>in</strong>, thus under growth conditions the low level <strong>of</strong> Ime2 does not suffice for<br />

<strong>in</strong>itiation <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> the absence <strong>of</strong> Ime1.<br />

Yeast cells use the same regulators, but <strong>in</strong> reverse directions, to control alternative<br />

developmental pathways. Ime2 is a positive regulator <strong>of</strong> <strong>meiosis</strong> that prevents pseudohyphae<br />

development <strong>in</strong> growth media with acetate as the sole carbon source (Donzeau and Bandlow,<br />

1999). The activity <strong>of</strong> the cAMP/PKA signal pathway is a major player <strong>in</strong> determ<strong>in</strong><strong>in</strong>g the<br />

developmental choice <strong>yeast</strong> cells make. Entry <strong>in</strong>to mitosis with either the <strong>yeast</strong> form or<br />

filamentous morphology requires high activity <strong>of</strong> this pathway [for reviews see (Gancedo, 2001)],<br />

whereas entry <strong>in</strong>to <strong>meiosis</strong> requires low PKA activity (Matsumoto et al., 1983). Sok2 is a positive<br />

regulator <strong>of</strong> mitosis (Ward et al., 1995) and a negative regulator for the transcription <strong>of</strong> IME1,<br />

<strong>meiosis</strong>, and filamentous growth (Shenhar and Kassir, 2001; Ward et al., 1995). The negative role<br />

<strong>of</strong> Sok2 <strong>in</strong> pseudohyphae formation is not clear s<strong>in</strong>ce its C. albicans homolog, Efg1, that is a<br />

negative regulator <strong>of</strong> <strong>meiosis</strong> when expressed <strong>in</strong> S. cerevisiae [complement<strong>in</strong>g sok2∆ (Shenhar<br />

and Kassir, 2001)], is a positive regulator <strong>of</strong> filamentous growth <strong>in</strong> both S. cerevisiae and C.<br />

albicans (Stoldt et al., 1997). Msn2/4 exhibit reverse tasks, it functions as a negative regulator <strong>of</strong><br />

the mitotic cell cycle (Smith et al., 1998) and filamentous growth (Stanhill et al., 1999) and as a<br />

positive regulator for the transcription <strong>of</strong> IME1 and <strong>meiosis</strong> (Shenhar and Kassir, 2001). The<br />

activity <strong>of</strong> Sok2 and Msn2/4 <strong>in</strong> grow<strong>in</strong>g cells requires their phosphorylation by PKA (Gorner et<br />

al., 1998; Shenhar and Kassir, 2001; Smith et al., 1998; Ward et al., 1995). The use <strong>of</strong> the same<br />

regulators, but <strong>in</strong> reverse directions for different developmental pathways, ensures that one<br />

pathway will be an alternative to the other one. When cells enter mitosis, <strong>meiosis</strong> will be<br />

repressed, whereas when cells enter <strong>meiosis</strong>, mitosis will be blocked. Thus, a s<strong>in</strong>gle signal<br />

transduction pathway, the cAMP/PKA, is sufficient to control two alternative developmental<br />

pathways.<br />

Acknowledgment. We thank M. Foiani for his hospitality while writ<strong>in</strong>g this review, for fruitful<br />

discussions and critical read<strong>in</strong>g <strong>of</strong> the review.<br />

42


References<br />

Anderson, S. F., Steber, C. M., Esposito, R. E., and Coleman, J. E. (1995). UME6, a negative<br />

regulator <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae, conta<strong>in</strong>s a C-term<strong>in</strong>al Zn2Cys6 b<strong>in</strong>uclear<br />

cluster that b<strong>in</strong>ds the URS1 DNA sequence <strong>in</strong> a z<strong>in</strong>c-dependent manner. Prote<strong>in</strong> Sci 4, 1832-<br />

1843.<br />

Beck, T., and Hall, M. N. (1999). The TOR signall<strong>in</strong>g pathway controls nuclear localization <strong>of</strong><br />

nutrient- regulated transcription factors. Nature 402, 689-692.<br />

Ben- Dov, N. (1994) The translational <strong>regulation</strong> <strong>of</strong> IME1 <strong>in</strong> the <strong>yeast</strong> Saccharomyces<br />

cerevisiae. M.Sc thesis, Technion, Haifa Israel., M.Sc, Technion, Haifa.<br />

Benjam<strong>in</strong>, K. R., Zhang, C., Shokat, K. M., and Herskowitz, I. (2002). Control <strong>of</strong> landmark<br />

events <strong>in</strong> <strong>meiosis</strong> by the CDK Cdc28 and the<br />

<strong>meiosis</strong>-specific k<strong>in</strong>ase Ime2. Submitted.<br />

Benni, M. L., and Neigeborn, L. (1997). Identification <strong>of</strong> a new class <strong>of</strong> negative regulators<br />

affect<strong>in</strong>g sporulation-specific gene expression <strong>in</strong> <strong>yeast</strong> [published erratum appears <strong>in</strong> Genetics<br />

1998 Jan;148(1):537]. Genetics 147, 1351-1366.<br />

Berger, K. H., and Yaffe, M. P. (2000). Mitochondrial DNA <strong>in</strong>heritance <strong>in</strong> Saccharomyces<br />

cerevisiae. Trends Microbiol 8, 508-513.<br />

Bishop, D. K., Park, D., Xu, L., and Kleckner, N. (1992). DMC1: a <strong>meiosis</strong>-specific <strong>yeast</strong><br />

homolog <strong>of</strong> E. coli recA required for recomb<strong>in</strong>ation, synaptonemal complex formation, and cell<br />

cycle progression. Cell 69, 439-456.<br />

Blumental-Perry, a., Li, w., Simchen, G., and Mitchell, A. P. (2002). Repression and activation<br />

doma<strong>in</strong>s <strong>of</strong> Rme1p structurally overlap, but differ <strong>in</strong> genetic requirements. Mol Biol Cell 13,<br />

1709-1721.<br />

Bogengruber, E., Eichberger, T., Briza, P., Dawes, I. W., Breitenbach, M., and Schricker, R.<br />

(1998). Sporulation-specific expression <strong>of</strong> the <strong>yeast</strong> DIT1/DIT2 promoter is controlled by a<br />

newly identified repressor element and the short form <strong>of</strong> Rim101p. Eur J Biochem 258, 430-436.<br />

Boger-Nadjar, E. (2000) Regulation <strong>of</strong> DNA replication <strong>in</strong> mitosis and <strong>meiosis</strong> <strong>in</strong> the budd<strong>in</strong>g<br />

<strong>yeast</strong> Saccharomyces cerevisiae, Ph.D, Technion - Israel Institute <strong>of</strong> Technology, Haifa.<br />

Bolte, M., Steigemann, P., Braus, G. H., and Irniger, S. (2002). Inhibition <strong>of</strong> APC-mediated<br />

proteolysis by the <strong>meiosis</strong>-specific prote<strong>in</strong> k<strong>in</strong>ase Ime2. Proc Natl Acad Sci U S A 99, 4385-<br />

4390.<br />

43


Bowdish, K. S., and Mitchell, A. P. (1993). Bipartite structure <strong>of</strong> an early meiotic upstream<br />

activation sequence from Saccharomyces cerevisiae. Mol Cell Biol 13, 2172-2181.<br />

Bowdish, K. S., Yuan, H. E., and Mitchell, A. P. (1994). Analysis <strong>of</strong> RIM11, a <strong>yeast</strong> prote<strong>in</strong><br />

k<strong>in</strong>ase that phosphorylates the meiotic activator IME1. Mol Cell Biol 14, 7909-7919.<br />

Bowdish, K. S., Yuan, H. E., and Mitchell, A. P. (1995). Positive control <strong>of</strong> <strong>yeast</strong> meiotic genes<br />

by the negative regulator UME6. Mol Cell Biol 15, 2955-2961.<br />

Briza, P., Breitenbach, M., Ell<strong>in</strong>ger, A., and Segall, J. (1990). Isolation <strong>of</strong> two developmentally<br />

regulated genes <strong>in</strong>volved <strong>in</strong> spore wall maturation <strong>in</strong> Saccharomyces cerevisiae. Genes Dev 4,<br />

1775-1789.<br />

Broach, J. R. (1991). RAS genes <strong>in</strong> Saccharomyces cerevisiae: signal transduction <strong>in</strong> search <strong>of</strong> a<br />

pathway. Trends Genet 7, 28-33.<br />

Broek, D., Toda, T., Michaeli, T., Lev<strong>in</strong>, L., Birchmeier, C., Zoller, M., Powers, S., and Wigler,<br />

M. (1987). The S. cerevisiae CDC25 gene product regulates the RAS/adenylate cyclase pathway.<br />

Cell 48, 789-799.<br />

Burgess, S. M., Ajimura, M., and Kleckner, N. (1999). GCN5-dependent histone H3 acetylation<br />

and RPD3-dependent histone H4 deacetylation have dist<strong>in</strong>ct, oppos<strong>in</strong>g effects on IME2<br />

transcription, dur<strong>in</strong>g <strong>meiosis</strong> and dur<strong>in</strong>g vegetative growth, <strong>in</strong> budd<strong>in</strong>g <strong>yeast</strong>. Proc Natl Acad Sci<br />

U S A 96, 6835-6840.<br />

Cairns, B. (1998). Chromat<strong>in</strong> remodel<strong>in</strong>g mach<strong>in</strong>es: similar motors, ulterior motives. Trends<br />

Biochem Sci 23, 20-25.<br />

Carlson, M., and Laurent, B. C. (1994). The SNF/SWI family <strong>of</strong> global transcriptional activators.<br />

Curr Op<strong>in</strong> Cell Biol 6, 396-402.<br />

Chu, S., DeRisi, J., Eisen, M., Mulholland, J., Botste<strong>in</strong>, D., Brown, P. O., and Herskowitz, I.<br />

(1998). The transcriptional program <strong>of</strong> sporulation <strong>in</strong> budd<strong>in</strong>g <strong>yeast</strong> [published erratum appears <strong>in</strong><br />

Science 1998 Nov 20;282(5393):1421]. Science 282, 699-705.<br />

Chu, S., and Herskowitz, I. (1998). Gametogenesis <strong>in</strong> <strong>yeast</strong> is regulated by a transcriptional<br />

cascade dependent on Ndt80. Mol Cell 1, 685-696.<br />

Colom<strong>in</strong>a, N., Gari, E., Gallego, C., Herrero, E., and Aldea, M. (1999). G1 cycl<strong>in</strong>s block the<br />

Ime1 pathway to make mitosis and <strong>meiosis</strong> <strong>in</strong>compatible <strong>in</strong> budd<strong>in</strong>g <strong>yeast</strong>. Embo J 18, 320-329.<br />

Cooper, K. F., Mallory, M. J., Egeland, D. B., Jarnik, M., and Strich, R. (2000). Ama1p is a<br />

<strong>meiosis</strong>-specific regulator <strong>of</strong> the anaphase promot<strong>in</strong>g complex/cyclosome <strong>in</strong> <strong>yeast</strong>. Proc Natl<br />

Acad Sci U S A 97, 14548-14553.<br />

Cooper, K. F., Mallory, M. J., Smith, J. B., and Strich, R. (1997). Stress and developmental<br />

<strong>regulation</strong> <strong>of</strong> the <strong>yeast</strong> C-type cycl<strong>in</strong> Ume3p (Srb11p/Ssn8p). EMBO J 16, 4665-4675.<br />

44


Cooper, K. F., and Strich, R. (1998). Functional analysis <strong>of</strong> the Ume3p/ Srb11p-RNA polymerase<br />

II holoenzyme <strong>in</strong>teraction. Gene Expr 8, 43-57.<br />

Covitz, P. A., Herskowitz, I., and Mitchell, A. P. (1991). The <strong>yeast</strong> RME1 gene encodes a<br />

putative z<strong>in</strong>c f<strong>in</strong>ger prote<strong>in</strong> that is directly repressed by a1-alpha 2. Genes Dev 5, 1982-1989.<br />

Covitz, P. A., and Mitchell, A. P. (1993). Repression by the <strong>yeast</strong> meiotic <strong>in</strong>hibitor RME1. Genes<br />

Dev 7, 1598-1608.<br />

Covitz, P. A., Song, W., and Mitchell, A. P. (1994). Requirement for RGR1 and SIN4 <strong>in</strong> RME1dependent<br />

repression <strong>in</strong> Saccharomyces cerevisiae. Genetics 138, 577-586.<br />

Davis, L., Barbera, M., McDonnell, A., McIntyre, K., Sternglanz, R., J<strong>in</strong>, Q., Loidl, J., and<br />

Engebrecht, J. (2001). The Saccharomyces cerevisiae MUM2 gene <strong>in</strong>teracts with the DNA<br />

replication mach<strong>in</strong>ery and is required for meiotic levels <strong>of</strong> double strand breaks. Genetics 157,<br />

1179-1189.<br />

De Silva-Udawatta, M. N., and Cannon, J. F. (2001). Roles <strong>of</strong> trehalose phosphate synthase <strong>in</strong><br />

<strong>yeast</strong> glycogen metabolism and sporulation. Mol Microbiol 40, 1345-1356.<br />

Deng, C., and Saunders, W. S. (2001). RIM4 encodes a meiotic activator required for early events<br />

<strong>of</strong> <strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae. Mol Genet Genomics 266, 497-504.<br />

Dirick, L., Goetsch, L., Ammerer, G., and Byers, B. (1998). Regulation <strong>of</strong> meiotic S phase by<br />

Ime2 and a Clb5,6-associated k<strong>in</strong>ase <strong>in</strong> Saccharomyces cerevisiae. Science 281, 1854-1857.<br />

Donzeau, M., and Bandlow, W. (1999). The <strong>yeast</strong> trimeric guan<strong>in</strong>e nucleotide-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong><br />

alpha subunit, Gpa2p, controls the <strong>meiosis</strong>-specific k<strong>in</strong>ase Ime2p activity <strong>in</strong> response to nutrients.<br />

Mol Cell Biol 19, 6110-6119.<br />

Elford, H. L. (1968). Effect <strong>of</strong> hydroxyurea on ribonucleotide reductase. Biochem Biophys Res<br />

Commun 33, 129-135.<br />

Engebrecht, J., Masse, S., Davis, L., Rose, K., and Kessel, T. (1998). Yeast meiotic mutants<br />

pr<strong>of</strong>icient for the <strong>in</strong>duction <strong>of</strong> ectopic recomb<strong>in</strong>ation. Genetics 148, 581-598.<br />

Engebrecht, J., and Roeder, G. S. (1990). MER1, a <strong>yeast</strong> gene required for chromosome pair<strong>in</strong>g<br />

and genetic recomb<strong>in</strong>ation, is <strong>in</strong>duced <strong>in</strong> <strong>meiosis</strong>. Mol Cell Biol 10, 2379-2389.<br />

Estruch, F., and Carlson, M. (1993). Two homologous z<strong>in</strong>c f<strong>in</strong>ger genes identified by multicopy<br />

suppression <strong>in</strong> a SNF1 prote<strong>in</strong> k<strong>in</strong>ase mutant <strong>of</strong> Saccharomyces cerevisiae. Mol Cell Biol 13,<br />

3872-3881.<br />

Fields, S. (1990). Pheromone response <strong>in</strong> <strong>yeast</strong>. Trends Biochem Sci 15, 270-273.<br />

Foiani, M., Liberi, G., Lucch<strong>in</strong>i, G., and Plevani, P. (1995). Cell cycle-dependent phosphorylation<br />

and dephosphorylation <strong>of</strong> the <strong>yeast</strong> DNA polymerase alpha-primase B subunit. Mol Cell Biol 15,<br />

883-891.<br />

45


Foiani, M., Nadjar-Boger, E., Capone, R., Sagee, S., Hashimshoni, T., and Kassir, Y. (1996). A<br />

<strong>meiosis</strong>-specific prote<strong>in</strong> k<strong>in</strong>ase, Ime2, is required for the correct tim<strong>in</strong>g <strong>of</strong> DNA replication and<br />

for spore formation <strong>in</strong> <strong>yeast</strong> <strong>meiosis</strong>. Mol Gen Genet 253, 278-288.<br />

Frenz, L. M., Johnson, A. L., and Johnston, L. H. (2001). Rme1, which controls CLN2 expression<br />

<strong>in</strong> Saccharomyces cerevisiae, is a nuclear prote<strong>in</strong> that is cell cycle regulated. Mol Genet<br />

Genomics 266, 374-384.<br />

Friesen, H., Hepworth, S. R., and Segall, J. (1997). An Ssn6-Tup1-dependent negative regulatory<br />

element controls sporulation- specific expression <strong>of</strong> DIT1 and DIT2 <strong>in</strong> Saccharomyces cerevisiae.<br />

Mol Cell Biol 17, 123-134.<br />

Friesen, H., Tanny, J. C., and Segall, J. (1998). Spe3, which encodes spermid<strong>in</strong>e synthase, is<br />

required for full repression through NRE(DIT) <strong>in</strong> Saccharomyces cerevisiae. Genetics 150, 59-73.<br />

Gailus-Durner, V., Xie, J., Ch<strong>in</strong>tamaneni, C., and Vershon, A. K. (1996). Participation <strong>of</strong> the<br />

<strong>yeast</strong> activator Abf1 <strong>in</strong> <strong>meiosis</strong>-specific expression <strong>of</strong> the HOP1 gene. Mol Cell Biol 16, 2777-<br />

2786.<br />

Gancedo, J. M. (2001). Control <strong>of</strong> pseudohyphae formation <strong>in</strong> Saccharomyces cerevisiae. FEMS<br />

Microbiol Rev 25, 107-123.<br />

Goetsch, L., and Byers, B. (1982). Meiotic cytology <strong>of</strong> Saccharomyces cerevisiae <strong>in</strong> protoplast<br />

lysates. Mol Gen Genet 187, 54-60.<br />

Goldmark, J. P., Fazzio, T. G., Estep, P. W., Church, G. M., and Tsukiyama, T. (2000). The Isw2<br />

chromat<strong>in</strong> remodel<strong>in</strong>g complex represses early meiotic genes upon recruitment by Ume6p [In<br />

Process Citation]. Cell 103, 423-433.<br />

Golemis, E. A., and Brent, R. (1992). Fused prote<strong>in</strong> doma<strong>in</strong>s <strong>in</strong>hibit DNA b<strong>in</strong>d<strong>in</strong>g by LexA. Mol<br />

Cell Biol 12, 3006-3014.<br />

Gorner, W., Durchschlag, E., Mart<strong>in</strong>ez-Pastor, M. T., Estruch, F., Ammerer, G., Hamilton, B.,<br />

Ruis, H., and Schuller, C. (1998). Nuclear localization <strong>of</strong> the C2H2 z<strong>in</strong>c f<strong>in</strong>ger prote<strong>in</strong> Msn2p is<br />

regulated by stress and prote<strong>in</strong> k<strong>in</strong>ase A activity. Genes Dev 12, 586-597.<br />

Gorner, W., Durchschlag, E., Wolf, J., Brown, E. L., Ammerer, G., Ruis, H., and Schuller, C.<br />

(2002). Acute glucose starvation activates the nuclear localization signal <strong>of</strong> a stress-specific <strong>yeast</strong><br />

transcription factor. EMBO J 21, 135-144.<br />

Goutte, C., and Johnson, A. D. (1988). a1 prote<strong>in</strong> alters the DNA b<strong>in</strong>d<strong>in</strong>g specificity <strong>of</strong> alpha 2<br />

repressor. Cell 52, 875-882.<br />

Granot, D., Margolskee, J. P., and Simchen, G. (1989). A long region upstream <strong>of</strong> the IME1 gene<br />

regulates <strong>meiosis</strong> <strong>in</strong> <strong>yeast</strong>. Mol Gen Genet 218, 308-314.<br />

46


Guan, K., Hakes, D. J., Wang, Y., Park, H. D., Cooper, T. G., and Dixon, J. E. (1992). A <strong>yeast</strong><br />

prote<strong>in</strong> phosphatase related to the vacc<strong>in</strong>ia virus VH1 phosphatase is <strong>in</strong>duced by nitrogen<br />

starvation. Proc Natl Acad Sci U S A 89, 12175-12179.<br />

Gustafsson, C. M., Myers, L. C., Li, Y., Redd, M. J., Lui, M., Erdjument-Bromage, H., Tempst,<br />

P., and Kornberg, R. D. (1997). Identification <strong>of</strong> Rox3 as a component <strong>of</strong> mediator and RNA<br />

polymerase II holoenzyme. J Biol Chem 272, 48-50.<br />

Guttmann-Raviv, N., Boger-Nadjar, E., Edri, I., and Kassir, Y. (2001). Cdc28 and Ime2 Possess<br />

Redundant Functions <strong>in</strong> Promot<strong>in</strong>g Entry Into Premeiotic DNA Replication <strong>in</strong> Saccharomyces<br />

cerevisiae. Genetics 159, 1547-1558.<br />

Guttmann-Raviv, N., and Kassir, Y. (2002). Ime2, a <strong>meiosis</strong>-specific k<strong>in</strong>ase <strong>in</strong> <strong>yeast</strong>, is required<br />

for destabilization <strong>of</strong> its transcriptional activator, Ime1. Mol Cell Biol 22, 2047-2056.<br />

Hajji, K., Clotet, J., and Ar<strong>in</strong>o, J. (1999). Disruption and phenotypic analysis <strong>of</strong> seven ORFs from<br />

the left arm <strong>of</strong> chromosome XV <strong>of</strong> Saccharomyces cerevisiae. Yeast 15, 435-441.<br />

Hamasaki-Katagiri, N., Tabor, C. W., and Tabor, H. (1997). Spermid<strong>in</strong>e biosynthesis <strong>in</strong><br />

Saccharomyces cerevisae: polyam<strong>in</strong>e requirement <strong>of</strong> a null mutant <strong>of</strong> the SPE3 gene (spermid<strong>in</strong>e<br />

synthase). Gene 187, 35-43.<br />

Hanna-Rose, W., and Hansen, U. (1996). Active repression mechanisms <strong>of</strong> eukaryotic<br />

transcription repressors. Trends Genet 12, 229-234.<br />

Henchoz, S., Chi, Y., Catar<strong>in</strong>, B., Herskowitz, I., Deshaies, R. J., and Peter, M. (1997).<br />

Phosphorylation- and ubiquit<strong>in</strong>-dependent degradation <strong>of</strong> the cycl<strong>in</strong>- dependent k<strong>in</strong>ase <strong>in</strong>hibitor<br />

Far1p <strong>in</strong> budd<strong>in</strong>g <strong>yeast</strong>. Genes Dev 11, 3046-3060.<br />

Hepworth, S. R., Ebisuzaki, L. K., and Segall, J. (1995). A 15-base-pair element activates the<br />

SPS4 gene midway through sporulation <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 15, 3934-<br />

3944.<br />

Hepworth, S. R., Friesen, H., and Segall, J. (1998). NDT80 and the meiotic recomb<strong>in</strong>ation<br />

checkpo<strong>in</strong>t regulate expression <strong>of</strong> middle sporulation-specific genes <strong>in</strong> saccharomyces cerevisiae.<br />

Mol Cell Biol 18, 5750-5761.<br />

Herskowitz, I. (1995). MAP k<strong>in</strong>ase pathways <strong>in</strong> <strong>yeast</strong>: for mat<strong>in</strong>g and more. Cell 80, 187-197.<br />

Honigberg, S. M., and Lee, R. H. (1998). Snf1 k<strong>in</strong>ase connects nutritional pathways controll<strong>in</strong>g<br />

<strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 18, 4548-4555.<br />

Irniger, S., and Nasmyth, K. (1997). The anaphase-promot<strong>in</strong>g complex is required <strong>in</strong> G1 arrested<br />

<strong>yeast</strong> cells to <strong>in</strong>hibit B-type cycl<strong>in</strong> accumulation and to prevent uncontrolled entry <strong>in</strong>to S-phase. J<br />

Cell Sci 110, 1523-1531.<br />

47


Jeffrey, P. D., Russo, A. A., Polyak, K., Gibbs, E., Hurwitz, J., Massague, J., and Pavletich, N. P.<br />

(1995). Mechanism <strong>of</strong> CDK activation revealed by the structure <strong>of</strong> a cycl<strong>in</strong>A-CDK2 complex.<br />

Nature 376, 313-320.<br />

Jiang, R., and Carlson, M. (1996). Glucose regulates prote<strong>in</strong> <strong>in</strong>teractions with<strong>in</strong> the <strong>yeast</strong> SNF1<br />

prote<strong>in</strong> k<strong>in</strong>ase complex. Genes Dev 10, 3105-3115.<br />

Johnson, A. D., and Herskowitz, I. (1985). A repressor (MAT alpha 2 Product) and its operator<br />

control expression <strong>of</strong> a set <strong>of</strong> cell type specific genes <strong>in</strong> <strong>yeast</strong>. Cell 42, 237-247.<br />

Johnston, L. H., Johnson, A. L., and Barker, D. G. (1986). The expression <strong>in</strong> <strong>meiosis</strong> <strong>of</strong> genes<br />

which are transcribed periodically <strong>in</strong> the mitotic cell cycle <strong>of</strong> budd<strong>in</strong>g <strong>yeast</strong>. Exp Cell Res 165,<br />

541-549.<br />

Johnston, M., and Carlson, M., eds. (1992). Regulation <strong>of</strong> carbon and phosphate utilization.<br />

Jung, U. S., and Lev<strong>in</strong>, D. E. (1999). Genome-wide analysis <strong>of</strong> gene expression regulated by the<br />

<strong>yeast</strong> cell wall <strong>in</strong>tegrity signall<strong>in</strong>g pathway. Mol Microbiol 34, 1049-1057.<br />

Kadosh, D., and Struhl, K. (1997). Repression by Ume6 <strong>in</strong>volves recruitment <strong>of</strong> a complex<br />

conta<strong>in</strong><strong>in</strong>g S<strong>in</strong>3 corepressor and Rpd3 histone deacetylase to target promoters. Cell 89, 365-371.<br />

Kadosh, D., and Struhl, K. (1998). Targeted recruitment <strong>of</strong> the S<strong>in</strong>3-rpd3 histone deacetylase<br />

complex generates a highly localized doma<strong>in</strong> <strong>of</strong> repressed chromat<strong>in</strong> In vivo. Mol Cell Biol 18,<br />

5121-5127.<br />

Kao, G., Shah, J. C., and Clancy, M. J. (1990). An RME1-<strong>in</strong>dependent pathway for sporulation<br />

control <strong>in</strong> Saccharomyces cerevisiae acts through IME1 transcript accumulation. Genetics 126,<br />

823-835.<br />

Kassir, Y., Granot, D., and Simchen, G. (1988). IME1, a positive regulator gene <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> S.<br />

cerevisiae. Cell 52, 853-862.<br />

Kassir, Y., and Simchen, G. (1976). Regulation <strong>of</strong> mat<strong>in</strong>g and <strong>meiosis</strong> <strong>in</strong> <strong>yeast</strong> by the mat<strong>in</strong>g-type<br />

region. Genetics 82, 187-206.<br />

Kassir, Y., and Simchen, G. (1985). Mutations lead<strong>in</strong>g to expression <strong>of</strong> the cryptic HMRa locus<br />

<strong>in</strong> the <strong>yeast</strong> Saccharomyces cerevisiae [published erratum appears <strong>in</strong> Genetics 1987<br />

May;116(1):181]. Genetics 109, 481-492.<br />

Kassir, Y., and Simchen, G. (1991). Monitor<strong>in</strong>g <strong>meiosis</strong> and sporulation <strong>in</strong> Saccharomyces<br />

cerevisiae. Methods Enzymol 194, 94-110.<br />

Kasten, M. M., Dorland, S., and Stillman, D. J. (1997). A large prote<strong>in</strong> complex conta<strong>in</strong><strong>in</strong>g the<br />

<strong>yeast</strong> S<strong>in</strong>3p and Rpd3p transcriptional regulators. Mol Cell Biol 17, 4852-4858.<br />

Kawaguchi, H., Yoshida, M., and Yamashita, I. (1992). Nutritional <strong>regulation</strong> <strong>of</strong> <strong>meiosis</strong>-specific<br />

gene expression <strong>in</strong> Saccharomyces cerevisiae. Biosci Biotech Biochem 56, 289-297.<br />

48


Keleher, C. A., Redd, M. J., Schultz, J., Carlson, M., and Johnson, A. D. (1992). Ssn6-Tup1 is a<br />

general repressor <strong>of</strong> transcription <strong>in</strong> <strong>yeast</strong>. Cell 68, 709-719.<br />

Kelley, L. A., MacCallum, R. M., and Sternberg, M. J. (2000). Enhanced genome annotation<br />

us<strong>in</strong>g structural pr<strong>of</strong>iles <strong>in</strong> the program 3D-PSSM. J Mol Biol 229, 499-520.<br />

Kihara, K., Nakamura, M., Akada, R., and Yamashita, I. (1991). Positive and negative elements<br />

upstream <strong>of</strong> the <strong>meiosis</strong>-specific glucoamylase gene <strong>in</strong> Saccharomyces cerevisiae. Mol Gen<br />

Genet 226, 383-392.<br />

Kom<strong>in</strong>ami, K., Sakata, Y., Sakai, M., and Yamashita, I. (1993). Prote<strong>in</strong> k<strong>in</strong>ase activity associated<br />

with the IME2 gene product, a meiotic <strong>in</strong>ducer <strong>in</strong> the <strong>yeast</strong> Saccharomyces cerevisiae. Biosci<br />

Biotechnol Biochem 57, 1731-1735.<br />

Kornberg, R. D. (1999). Eukaryotic transcriptional control. Trends <strong>in</strong> Cell Biology 9, M46-49.<br />

Kunoh, T., Kaneko, Y., and Harashima, S. (2000). YHP1 encodes a new homeoprote<strong>in</strong> that b<strong>in</strong>ds<br />

to the IME1 promoter <strong>in</strong> Saccharomyces cerevisiae. Yeast 16, 439-449.<br />

Kupiec, M., Byers, B., Esposito, R. E., and Mitchell, A. P. (1997). Meiosis and sporulation <strong>in</strong><br />

Saccharomyces cerevisiae, Vol 3 (Cold Spr<strong>in</strong>g Harbor, Cold Spr<strong>in</strong>g Harbor Laboratory Press).<br />

Lamb, T. M., and Mitchell, A. P. (2001). Coupl<strong>in</strong>g <strong>of</strong> Saccharomyces cerevisiae Early Meiotic<br />

Gene Expression to DNA Replication Depends Upon RPD3 and SIN3. Genetics 157, 545-556.<br />

Law, D. T., and Segall, J. (1988). The SPS100 gene <strong>of</strong> Saccharomyces cerevisiae is activated late<br />

<strong>in</strong> the sporulation process and contributes to spore wall maturation. Mol Cell Biol 8, 912-922.<br />

Lengeler, K. B., Davidson, R. C., D'Souza, C., Harashima, T., Shen, W. C., Wang, P., Pan, X.,<br />

Waugh, M., and Heitman, J. (2000). Signal transduction cascades regulat<strong>in</strong>g fungal development<br />

and virulence. Microbiol Mol Biol Rev 64, 746-785.<br />

Lesage, P., Yang, X., and Carlson, M. (1996). Yeast SNF1 prote<strong>in</strong> k<strong>in</strong>ase <strong>in</strong>teracts with SIP4, a<br />

C6 z<strong>in</strong>c cluster transcriptional activator: a new role for SNF1 <strong>in</strong> the glucose response. Mol Cell<br />

Biol 16, 1921-1928.<br />

Leu, J. Y., and Roeder, G. S. (1999). The pachytene checkpo<strong>in</strong>t <strong>in</strong> S. cerevisiae depends on<br />

Swe1-mediated phosphorylation <strong>of</strong> the cycl<strong>in</strong>-dependent k<strong>in</strong>ase Cdc28. Mol Cell 4, 805-814.<br />

Lew, D. J., We<strong>in</strong>ert, T., and Pr<strong>in</strong>gle, J. R. (1997). Cell cycle control <strong>in</strong> Saccharomyces cerevisiae,<br />

Vol 3 (Cold Spr<strong>in</strong>g Harbor, Cold Spr<strong>in</strong>g Harbor Laboratory Press).<br />

Li, T., Stark, M. R., Johnson, A. D., and Wolberger, C. (1995). Crystal structure <strong>of</strong> the<br />

MATa1/MAT alpha 2 homeodoma<strong>in</strong> heterodimer bound to DNA. Science 270, 262-269.<br />

Li, W., and Mitchell, A. P. (1997). Proteolytic activation <strong>of</strong> Rim1p, a positive regulator <strong>of</strong> <strong>yeast</strong><br />

sporulation and <strong>in</strong>vasive growth. Genetics 145, 63-73.<br />

49


L<strong>in</strong>dgren, A., Bungard, D., Pierce, M., Xie, J., Vershon, A., and W<strong>in</strong>ter, E. (2000). The pachytene<br />

checkpo<strong>in</strong>t <strong>in</strong> Saccharomyces cerevisiae requires the Sum1 transcriptional repressor. Embo J 19,<br />

6489-6497.<br />

Lopes, J. M., Schulze, K. L., Yates, J. W., Hirsch, J. P., and Henry, S. A. (1993). The INO1<br />

promoter <strong>of</strong> Saccharomyces cerevisiae <strong>in</strong>cludes an upstream repressor sequence (URS1) common<br />

to a diverse set <strong>of</strong> <strong>yeast</strong> genes. J Bacteriol 175, 4235-4238.<br />

Lowndes, N. F., Johnson, A. L., and Johnston, L. H. (1991). Coord<strong>in</strong>ation <strong>of</strong> expression <strong>of</strong> DNA<br />

synthesis genes <strong>in</strong> budd<strong>in</strong>g <strong>yeast</strong> by a cell-cycle regulated trans factor. Nature 350, 247-250.<br />

Mai, B., and Breeden, L. (2000). CLN1 and its repression by Xbp1 are important for efficient<br />

sporulation <strong>in</strong> budd<strong>in</strong>g <strong>yeast</strong>. Mol Cell Biol 20, 478-487.<br />

Malathi, K., Xiao, Y., and Mitchell, A. P. (1997). Interaction <strong>of</strong> <strong>yeast</strong> repressor-activator prote<strong>in</strong><br />

Ume6p with glycogen synthase k<strong>in</strong>ase 3 homolog Rim11p. Mol Cell Biol 17, 7230-7236.<br />

Malathi, K., Xiao, Y., and Mitchell, A. P. (1999). Catalytic Roles <strong>of</strong> Yeast GSK3beta/Shaggy<br />

Homolog Rim11p <strong>in</strong> Meiotic Activation. Genetics 153, 1145-1152.<br />

Mandel, S., Robzyk, K., and Kassir, Y. (1994). IME1 gene encodes a transcription factor which is<br />

required to <strong>in</strong>duce <strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae. Dev Genet 15, 139-147.<br />

Mart<strong>in</strong>ez-Pastor, M. T., Marchler, G., Schuller, C., Marchler-Bauer, A., Ruis, H., and Estruch, F.<br />

(1996). The Saccharomyces cerevisiae z<strong>in</strong>c f<strong>in</strong>ger prote<strong>in</strong>s Msn2p and Msn4p are required for<br />

transcriptional <strong>in</strong>duction through the stress response element (STRE). Embo J 15, 2227-2235.<br />

Matsumoto, K., Uno, I., and Ishikawa, T. (1983). Initiation <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> <strong>yeast</strong> mutants defective<br />

<strong>in</strong> adenylate cyclase and cyclic AMP-dependent prote<strong>in</strong> k<strong>in</strong>ase. Cell 32, 417-423.<br />

Matsuura, A., Tre<strong>in</strong><strong>in</strong>, M., Mitsuzawa, H., Kassir, Y., Uno, I., and Simchen, G. (1990). The<br />

adenylate cyclase/prote<strong>in</strong> k<strong>in</strong>ase cascade regulates entry <strong>in</strong>to <strong>meiosis</strong> <strong>in</strong> Saccharomyces<br />

cerevisiae through the gene IME1. Embo J 9, 3225-3232.<br />

Mitchell, A. P., and Bowdish, K. S. (1992). Selection for early meiotic mutants <strong>in</strong> <strong>yeast</strong>. Genetics<br />

131, 65-72.<br />

Mitchell, A. P., Driscoll, S. E., and Smith, H. E. (1990). Positive control <strong>of</strong> sporulation-specific<br />

genes by the IME1 and IME2 products <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 10, 2104-<br />

2110.<br />

Mitchell, A. P., and Herskowitz, I. (1986). Activation <strong>of</strong> <strong>meiosis</strong> and sporulation by repression <strong>of</strong><br />

the RME1 product <strong>in</strong> <strong>yeast</strong>. Nature 319, 738-742.<br />

Mizuno, T., Nakazawa, N., Remgsamrarn, P., Kunoh, T., Oshima, Y., and Harashima, S. (1998).<br />

The Tup1-ssn6 general repressor is <strong>in</strong>volved <strong>in</strong> repression <strong>of</strong> IME1 encod<strong>in</strong>g a transcriptional<br />

activator <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> saccharomyces cerevisiae [In Process Citation]. Curr Genet 33, 239-247.<br />

50


Myer, V. E., and Young, R. A. (1998). RNA polymerase II holoenzymes and subcomplexes. J<br />

Biol Chem 273, 27757-27760.<br />

Nadjar-Boger, E. (2000). Regulation <strong>of</strong> DNA replication <strong>in</strong> mitosis and <strong>meiosis</strong> <strong>in</strong> the budd<strong>in</strong>g<br />

<strong>yeast</strong> Saccharomyces cerevisiae. PhD thesis, Technion - Israel Institute <strong>of</strong> Technology.<br />

Nehl<strong>in</strong>, J. O., Carlberg, M., and Ronne, H. (1991). Control <strong>of</strong> <strong>yeast</strong> GAL genes by MIG1<br />

repressor: a transcriptional cascade <strong>in</strong> the glucose response. Embo J 10, 3373-3377.<br />

Neigeborn, L., and Mitchell, A. P. (1991). The <strong>yeast</strong> MCK1 gene encodes a prote<strong>in</strong> k<strong>in</strong>ase<br />

homolog that activates early meiotic gene expression. Genes Dev 5, 533-548.<br />

Ng, H. H., and Bird, A. (2000). Histone deacetylases: silencers for hire. Trends Biochem Sci 25,<br />

121-126.<br />

Ozsarac, N., Bhattacharyya, M., Dawes, I. W., and Clancy, M. J. (1995). The SPR3 gene encodes<br />

a sporulation-specific homologue <strong>of</strong> the <strong>yeast</strong> CDC3/10/11/12 family <strong>of</strong> bud neck micr<strong>of</strong>ilaments<br />

and is regulated by ABFI. Gene 164, 157-162.<br />

Ozsarac, N., Straffon, M. J., Dalton, H. E., and Dawes, I. W. (1997). Regulation <strong>of</strong> gene<br />

expression dur<strong>in</strong>g <strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae: SPR3 is controlled by both ABFI and a<br />

new sporulation control element. Mol Cell Biol 17, 1152-1159.<br />

Pak, J., and Segall, J. (2002a). Regulation <strong>of</strong> the Premiddle and Middle Phases <strong>of</strong> Expression<br />

<strong>of</strong> the NDT80 Gene dur<strong>in</strong>g Sporulation <strong>of</strong> Saccharomyces cerevisiae. Mol Cell Biol.<br />

Pak, J., and Segall, J. (2002b). Role <strong>of</strong> Ndt80, Sum1, and Swe1 as Targets <strong>of</strong> the Meiotic<br />

Recomb<strong>in</strong>ation Checkpo<strong>in</strong>t that Control Exit from Pachytene and Spore Formation <strong>in</strong><br />

Saccharomyces cerevisiae. Mol Cell Biol.<br />

Pan, X., Harashima, T., and Heitman, J. (2000). Signal transduction cascades regulat<strong>in</strong>g<br />

pseudohyphal differentiation <strong>of</strong> Saccharomyces cerevisiae. Curr Op<strong>in</strong> Microbiol 3, 567-572.<br />

Pan, X., and Heitman, J. (2000). Sok2 regulates <strong>yeast</strong> pseudohyphal differentiation via a<br />

transcription factor cascade that regulates cell-cell adhesion. Mol Cell Biol 20, 8364-8372.<br />

Park, H. D., Beeser, A. E., Clancy, M. J., and Cooper, T. G. (1996). The S. cerevisiae nitrogen<br />

starvation-<strong>in</strong>duced Yvh1p and Ptp2p phosphatases play a role <strong>in</strong> control <strong>of</strong> sporulation. Yeast 12,<br />

1135-1151.<br />

Park, H. D., Luche, R. M., and Cooper, T. G. (1992). The <strong>yeast</strong> UME6 gene product is required<br />

for transcriptional repression mediated by the CAR1 URS1 repressor b<strong>in</strong>d<strong>in</strong>g site. Nucleic Acids<br />

Res 20, 1909-1915.<br />

Patton, E. E., Willems, A. R., and Tyers, M. (1998). Comb<strong>in</strong>atorial control <strong>in</strong> ubiquit<strong>in</strong>-dependent<br />

proteolysis: don't Skp the F-box hypothesis. Trends Genet 14, 236-243.<br />

51


Paz<strong>in</strong>, M. J., and Kadonaga, J. T. (1997). What's up and down with histone deacetylation and<br />

transcription? Cell 89, 325-328.<br />

Pemberton, L. F., and Blobel, G. (1997). Characterization <strong>of</strong> the Wtm prote<strong>in</strong>s, a novel family <strong>of</strong><br />

Saccharomyces cerevisiae transcriptional modulators with roles <strong>in</strong> meiotic <strong>regulation</strong> and<br />

silenc<strong>in</strong>g. Mol Cell Biol 17, 4830-4841.<br />

Pierce, M., Wagner, M., Xie, J., Gailus-Durner, V., Six, J., Vershon, A. K., and W<strong>in</strong>ter, E.<br />

(1998). <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> the SMK1 mitogen-activated prote<strong>in</strong> k<strong>in</strong>ase gene dur<strong>in</strong>g<br />

meiotic development <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 18, 5970-5980.<br />

Pijnappel, W. W., Schaft, D., Roguev, A., Shevchenko, A., Tekotte, H., Wilm, M., Rigaut, G.,<br />

Seraph<strong>in</strong>, B., Aasland, R., and Stewart, A. F. (2001). The S. cerevisiae SET3 complex <strong>in</strong>cludes<br />

two histone deacetylases, Hos2 and Hst1, and is a meiotic-specific repressor <strong>of</strong> the sporulation<br />

gene program. Genes Dev 15, 2991-3004.<br />

Primig, M., Williams, R. M., W<strong>in</strong>zeler, E. A., Tevzadze, G. G., Conway, A. R., Hwang, S. Y.,<br />

Davis, R. W., and Esposito, R. E. (2000). The core meiotic transcriptome <strong>in</strong> budd<strong>in</strong>g <strong>yeast</strong>s. Nat<br />

Genet 26, 415-423.<br />

Rabitsch, K. P., Toth, A., Galova, M., Schleiffer, A., Schaffner, G., Aigner, E., Rupp, C.,<br />

Penkner, A. M., Moreno-Borchart, A. C., Primig, M., et al. (2001). A screen for genes required<br />

for <strong>meiosis</strong> and spore formation based on whole-genome expression. Curr Biol 11, 1001-1009.<br />

Rayner, T. F., Gray, J. V., and Thorner, J. W. (2002). Direct and novel <strong>regulation</strong> <strong>of</strong> cAMPdependent<br />

prote<strong>in</strong> k<strong>in</strong>ase by Mck1p, a <strong>yeast</strong> glycogen synthase k<strong>in</strong>ase-3. J Biol Chem 277,<br />

16814-16822.<br />

Re<strong>in</strong>ders, A., Burckert, N., Boller, T., Wiemken, A., and De Virgilio, C. (1998). Saccharomyces<br />

cerevisiae cAMP-dependent prote<strong>in</strong> k<strong>in</strong>ase controls entry <strong>in</strong>to stationary phase through the<br />

Rim15p prote<strong>in</strong> k<strong>in</strong>ase. Genes Dev 12, 2943-2955.<br />

Re<strong>in</strong>ders, A., Burckert, N., Hohmann, S., Thevele<strong>in</strong>, J. M., Boller, T., Wiemken, A., and De<br />

Virgilio, C. (1997). Structural analysis <strong>of</strong> the subunits <strong>of</strong> the trehalose-6-phosphate<br />

synthase/phosphatase complex <strong>in</strong> Saccharomyces cerevisiae and their function dur<strong>in</strong>g heat shock.<br />

Mol Microbiol 24, 687-695.<br />

R<strong>in</strong>e, J., and Herskowitz, I. (1987). Four genes responsible for a position effect on expression<br />

from HML and HMR <strong>in</strong> Saccharomyces cerevisiae. Genetics 116, 9-22.<br />

R<strong>in</strong>e, J., Sprague, G. F., Jr., and Herskowitz, I. (1981). rme1 Mutation <strong>of</strong> Saccharomyces<br />

cerevisiae: map position and bypass <strong>of</strong> mat<strong>in</strong>g type locus control <strong>of</strong> sporulation. Mol Cell Biol 1,<br />

958-960.<br />

Roeder, G. S. (1997). Meiotic chromosomes: it takes two to tango. Genes Dev 11, 2600-2621.<br />

52


Roeder, G. S., and Bailis, J. M. (2000). The pachytene checkpo<strong>in</strong>t. Trends Genet 16, 395-403.<br />

Roman, H., and Sands, S. (1953). Heterogeneity <strong>of</strong> clones <strong>of</strong> Saccharomyces derived from<br />

haploid ascospores. Proc Natl Acad/ Science 39, 171-179.<br />

Roth, R., and Fogel, S. (1971). A system selective for <strong>yeast</strong> mutants deficient <strong>in</strong> meiotic<br />

recomb<strong>in</strong>ation. Mol Gen Genet 112, 295-305.<br />

Rub<strong>in</strong>-Bejerano, I., Mandel, S., Robzyk, K., and Kassir, Y. (1996). Induction <strong>of</strong> <strong>meiosis</strong> <strong>in</strong><br />

Saccharomyces cerevisiae depends on conversion <strong>of</strong> the transcriptional represssor Ume6 to a<br />

positive regulator by its regulated association with the transcriptional activator Ime1. Mol Cell<br />

Biol 16, 2518-2526.<br />

Rundlett, S. E., Carmen, A. A., Suka, N., Turner, B. M., and Grunste<strong>in</strong>, M. (1998).<br />

<strong>Transcriptional</strong> repression by UME6 <strong>in</strong>volves deacetylation <strong>of</strong> lys<strong>in</strong>e 5 <strong>of</strong> histone H4 by RPD3.<br />

Nature 392, 831-835.<br />

Rusche, L. N., and R<strong>in</strong>e, J. (2001). Conversion <strong>of</strong> a gene-specific repressor to a regional silencer.<br />

Genes Dev 15, 955-967.<br />

Sagee, S., Sherman, A., Shenhar, G., Robzyk, K., Ben-Doy, N., Simchen, G., and Kassir, Y.<br />

(1998). Multiple and dist<strong>in</strong>ct activation and repression sequences mediate the regulated<br />

transcription <strong>of</strong> IME1, a transcriptional activator <strong>of</strong> <strong>meiosis</strong>- specific genes <strong>in</strong> Saccharomyces<br />

cerevisiae [published erratum appears <strong>in</strong> Mol Cell Biol 1998 Jun;18(6):3645]. Mol Cell Biol 18,<br />

1985-1995.<br />

Schmitt, A. P., and McEntee, K. (1996). Msn2p, a z<strong>in</strong>c f<strong>in</strong>ger DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>, is the<br />

transcriptional activator <strong>of</strong> the multistress response <strong>in</strong> Saccharomyces cerevisiae. Proc Natl Acad<br />

Sci U S A 93, 5777-5782.<br />

Schwob, E., and Nasmyth, K. (1993). CLB5 and CLB6, a new pair <strong>of</strong> B cycl<strong>in</strong>s <strong>in</strong>volved <strong>in</strong> DNA<br />

replication <strong>in</strong> Saccharomyces cerevisiae. Genes Dev 7, 1160-1175.<br />

Segal, M., Clarke, D. J., Maddox, P., Salmon, E. D., Bloom, K., and Reed, S. I. (2000).<br />

Coord<strong>in</strong>ated sp<strong>in</strong>dle assembly and orientation requires Clb5p-dependent k<strong>in</strong>ase <strong>in</strong> budd<strong>in</strong>g <strong>yeast</strong>.<br />

J Cell Biol 148, 441-452.<br />

Shah, J. C., and Clancy, M. J. (1992). IME4, a gene that mediates MAT and nutritional control <strong>of</strong><br />

<strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 12, 1078-1086.<br />

Shefer-Vaida, M., Sherman, A., Ashkenazi, T., Robzyk, K., and Kassir, Y. (1995). Positive and<br />

negative feedback loops affect the transcription <strong>of</strong> IME1, a positive regulator <strong>of</strong> <strong>meiosis</strong> <strong>in</strong><br />

Saccharomyces cerevisiae. Dev Genet 16, 219-228.<br />

53


Shenhar, G. (2001). TRANSCRIPTIONAL REGULATION OF IME1, THE MASTER<br />

REGULATOR OF MEIOSIS IN BUDDING YEAST. PhD thesis, Technion - Israel Institute <strong>of</strong><br />

Technology.<br />

Shenhar, G., and Kassir, Y. (2001). A positive regulator <strong>of</strong> mitosis, Sok2, functions as a negative<br />

regulator <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 21, 1603-1612.<br />

Sherman, A. (1992). The <strong>regulation</strong> and time <strong>of</strong> action <strong>of</strong> the gene IME1 - <strong>in</strong>ducer <strong>of</strong> <strong>meiosis</strong> <strong>in</strong><br />

the <strong>yeast</strong> Saccharomyces cerevisiae., Ph. D, Hebrew University, Jerusalem.<br />

Sherman, A., Shefer, M., Sagee, S., and Kassir, Y. (1993). Post-transcriptional <strong>regulation</strong> <strong>of</strong><br />

IME1 determ<strong>in</strong>es <strong>in</strong>itiation <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae. Mol Gen Genet 237, 375-<br />

384.<br />

Shilo, V., Simchen, G., and Shilo, B. (1978). Initiation <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> cell cycle <strong>in</strong>itiation mutants<br />

<strong>of</strong> Saccharomyces cerevisiae. Exp Cell Res 112, 241-248.<br />

Shimizu, M., Hara, M., Murase, A., Sh<strong>in</strong>do, H., and Mitchell, A. P. (1997a). Dissection <strong>of</strong> the<br />

DNA b<strong>in</strong>d<strong>in</strong>g doma<strong>in</strong> <strong>of</strong> <strong>yeast</strong> Zn-f<strong>in</strong>ger prote<strong>in</strong> Rme1p, a repressor <strong>of</strong> meiotic activator IME1<br />

[In Process Citation]. Nucleic Acids Symp Ser, 175-176.<br />

Shimizu, M., Li, W., Covitz, P. A., Hara, M., Sh<strong>in</strong>do, H., and Mitchell, A. P. (1998). Genomic<br />

footpr<strong>in</strong>t<strong>in</strong>g <strong>of</strong> the <strong>yeast</strong> z<strong>in</strong>c f<strong>in</strong>ger prote<strong>in</strong> Rme1p and its roles <strong>in</strong> repression <strong>of</strong> the meiotic<br />

activator IME1. Nucleic Acids Res 26, 2329-2336.<br />

Shimizu, M., Li, W., Sh<strong>in</strong>do, H., and Mitchell, A. P. (1997b). <strong>Transcriptional</strong> repression at a<br />

distance through exclusion <strong>of</strong> activator b<strong>in</strong>d<strong>in</strong>g <strong>in</strong> vivo. Proc Natl Acad Sci U S A 94, 790-795.<br />

Sia, R. A., and Mitchell, A. P. (1995). Stimulation <strong>of</strong> later functions <strong>of</strong> the <strong>yeast</strong> meiotic prote<strong>in</strong><br />

k<strong>in</strong>ase Ime2p by the IDS2 gene product. Mol Cell Biol 15, 5279-5287.<br />

Simchen, G. (1974). Are mitotic functions required <strong>in</strong> <strong>meiosis</strong>? Genetics 76, 745-753.<br />

Slater, M. L. (1973). Effect <strong>of</strong> reversible <strong>in</strong>hibition <strong>of</strong> deoxyribonucleic acid synthesis on the<br />

<strong>yeast</strong> cell cycle. J Bacteriol 113, 263-270.<br />

Smith, A., Ward, M. P., and Garrett, S. (1998). Yeast PKA represses Msn2p/Msn4p-dependent<br />

gene expression to regulate growth, stress response and glycogen accumulation. Embo J 17,<br />

3556-3564.<br />

Smith, H. E., Driscoll, S. E., Sia, R. A., Yuan, H. E., and Mitchell, A. P. (1993). Genetic evidence<br />

for transcriptional activation by the <strong>yeast</strong> IME1 gene product. Genetics 133, 775-784.<br />

Smith, H. E., and Mitchell, A. P. (1989). A transcriptional cascade governs entry <strong>in</strong>to <strong>meiosis</strong> <strong>in</strong><br />

Saccharomyces cerevisiae. Mol Cell Biol 9, 2142-2152.<br />

Smith, H. E., Su, S. S., Neigeborn, L., Driscoll, S. E., and Mitchell, A. P. (1990). Role <strong>of</strong> IME1<br />

expression <strong>in</strong> <strong>regulation</strong> <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 10, 6103-6113.<br />

54


Song, W., and Carlson, M. (1998). Srb/mediator prote<strong>in</strong>s <strong>in</strong>teract functionally and physically with<br />

transcriptional repressor Sfl1. EMBO J 17, 5757-5765.<br />

Soushko, M., and Mitchell, A. P. (2000). An RNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> homologue that promotes<br />

sporulation-specific gene expression <strong>in</strong> Saccharomyces cerevisiae. Yeast 16, 631-639.<br />

Sprague, G. F., Jr. (1991). Signal transduction <strong>in</strong> <strong>yeast</strong> mat<strong>in</strong>g: receptors, transcription factors,<br />

and the k<strong>in</strong>ase connection. Trends Genet 7, 393-398.<br />

Stanhill, A., Schick, N., and Engelberg, D. (1999). The <strong>yeast</strong> ras/cyclic AMP pathway <strong>in</strong>duces<br />

<strong>in</strong>vasive growth by suppress<strong>in</strong>g the cellular stress response. Mol Cell Biol 19, 7529-7538.<br />

Stark, M. R., and Johnson, A. D. (1994). Interaction between two homeodoma<strong>in</strong> prote<strong>in</strong>s is<br />

specified by a short C- term<strong>in</strong>al tail. Nature 371, 429-432.<br />

Steber, C. M., and Esposito, R. E. (1995). UME6 is a central component <strong>of</strong> a developmental<br />

regulatory switch controll<strong>in</strong>g <strong>meiosis</strong>-specific gene expression. Proc Natl Acad Sci U S A 92,<br />

12490-12494.<br />

Stoldt, V. R., Sonneborn, A., Leuker, C. E., and Ernst, J. F. (1997). Efg1p, an essential regulator<br />

<strong>of</strong> morphogenesis <strong>of</strong> the human pathogen Candida albicans, is a member <strong>of</strong> a conserved class <strong>of</strong><br />

bHLH prote<strong>in</strong>s regulat<strong>in</strong>g morphogenetic processes <strong>in</strong> fungi. Embo J 16, 1982-1991.<br />

Strathern, J., Hicks, J., and Herskowitz, I. (1981). Control <strong>of</strong> cell type <strong>in</strong> <strong>yeast</strong> by the mat<strong>in</strong>g type<br />

locus. The alpha 1- alpha 2 hypothesis. J Mol Biol 147, 357-372.<br />

Strich, R., Slater, M. R., and Esposito, R. E. (1989). Identification <strong>of</strong> negative regulatory genes<br />

that govern the expression <strong>of</strong> early meiotic genes <strong>in</strong> <strong>yeast</strong>. Proc Natl Acad Sci U S A 86, 10018-<br />

10022.<br />

Strich, R., Surosky, R. T., Steber, C., Dubois, E., Messenguy, F., and Esposito, R. E. (1994).<br />

UME6 is a key regulator <strong>of</strong> nitrogen repression and meiotic development. Genes Dev 8, 796-810.<br />

Stuart, D., and Wittenberg, C. (1998). CLB5 and CLB6 are required for premeiotic DNA<br />

replication and activation <strong>of</strong> the meiotic S/M checkpo<strong>in</strong>t. Genes Dev 12, 2698-2710.<br />

Su, S. S., and Mitchell, A. P. (1993a). Identification <strong>of</strong> functionally related genes that stimulate<br />

early meiotic gene expression <strong>in</strong> <strong>yeast</strong>. Genetics 133, 67-77.<br />

Su, S. S., and Mitchell, A. P. (1993b). Molecular characterization <strong>of</strong> the <strong>yeast</strong> meiotic regulatory<br />

gene RIM1. Nucleic Acids Res 21, 3789-3797.<br />

Svetlov, V., and Cooper, T. (1995). Review: compilation and characteristics <strong>of</strong> dedicated<br />

transcription factors <strong>in</strong> Saccharomyces cerevisiae. Yeast 11, 1439-1484.<br />

Sym, M., Engebrecht, J. A., and Roeder, G. S. (1993). ZIP1 is a synaptonemal complex prote<strong>in</strong><br />

required for meiotic chromosome synapsis. Cell 72, 365-378.<br />

55


Szent-Gyorgyi, C. (1995). A bipartite operator <strong>in</strong>teracts with a heat shock element to mediate<br />

early meiotic <strong>in</strong>duction <strong>of</strong> Saccharomyces cerevisiae HSP82. Mol Cell Biol 15, 6754-6769.<br />

Thevele<strong>in</strong>, J. M., and de W<strong>in</strong>de, J. H. (1999). Novel sens<strong>in</strong>g mechanisms and targets for the<br />

cAMP-prote<strong>in</strong> k<strong>in</strong>ase A pathway <strong>in</strong> the <strong>yeast</strong> Saccharomyces cerevisiae. Mol Microbiol 33, 904-<br />

918.<br />

Thompson-Jaeger, S., Francois, J., Gaughran, J. P., and Tatchell, K. (1991). Deletion <strong>of</strong> SNF1<br />

affects the nutrient response <strong>of</strong> <strong>yeast</strong> and resembles mutations which activate the adenylate<br />

cyclase pathway. Genetics 129, 697-706.<br />

Toda, T., Cameron, S., Sass, P., Zoller, M., Scott, J. D., McMullen, B., Hurwitz, M., Krebs, E. G.,<br />

and Wigler, M. (1987a). Clon<strong>in</strong>g and characterization <strong>of</strong> BCY1, a locus encod<strong>in</strong>g a regulatory<br />

subunit <strong>of</strong> the cyclic AMP-dependent prote<strong>in</strong> k<strong>in</strong>ase <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol<br />

7, 1371-1377.<br />

Toda, T., Cameron, S., Sass, P., Zoller, M., and Wigler, M. (1987b). Three different genes <strong>in</strong> S.<br />

cerevisiae encode the catalytic subunits <strong>of</strong> the cAMP-dependent prote<strong>in</strong> k<strong>in</strong>ase. Cell 50, 277-287.<br />

Toda, T., Uno, I., Ishikawa, T., Powers, S., Kataoka, T., Broek, D., Cameron, S., Broach, J.,<br />

Matsumoto, K., and Wigler, M. (1985). In <strong>yeast</strong>, RAS prote<strong>in</strong>s are controll<strong>in</strong>g elements <strong>of</strong><br />

adenylate cyclase. Cell 40, 27-36.<br />

Toone, W. M., Johnson, A. L., Banks, G. R., Toyn, J. H., Stuart, D., Wittenberg, C., and<br />

Johnston, L. H. (1995). Rme1, a negative regulator <strong>of</strong> <strong>meiosis</strong>, is also a positive activator <strong>of</strong> G1<br />

cycl<strong>in</strong> gene expression. Embo J 14, 5824-5832.<br />

Trachtulcova, P., Janatova, I., Kohlwe<strong>in</strong>, S. D., and Hasek, J. (2000). Saccharomyces cerevisiae<br />

gene ISW2 encodes a microtubule-<strong>in</strong>teract<strong>in</strong>g prote<strong>in</strong> required for premeiotic DNA replication.<br />

Yeast 16, 35-47.<br />

Tre<strong>in</strong><strong>in</strong>, M., and Simchen, G. (1993). Mitochondrial activity is required for the expression <strong>of</strong><br />

IME1, a regulator <strong>of</strong> <strong>meiosis</strong> <strong>in</strong> <strong>yeast</strong>. Curr Genet 23, 223-227.<br />

Treitel, M. A., Kuch<strong>in</strong>, S., and Carlson, M. (1998). Snf1 Prote<strong>in</strong> K<strong>in</strong>ase Regulates<br />

Phosphorylation <strong>of</strong> the Mig1 Repressor <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 18, 6273-<br />

6280.<br />

Tsuboi, M. (1983). Tsuboi, M. The isolation and genetic analysis <strong>of</strong> sporulation-deficient mutants<br />

<strong>in</strong> Saccharomyces cerevisiae. Mol Gen Genet 191, 17-21.<br />

Tung, K. S., Hong, E. J., and Roeder, G. S. (2000). The pachytene checkpo<strong>in</strong>t prevents<br />

accumulation and phosphorylation <strong>of</strong> the <strong>meiosis</strong>-specific transcription factor Ndt80. Proc Natl<br />

Acad Sci U S A 97, 12187-12192.<br />

56


Uetz, P., Giot, L., Cagney, G., Mansfield, T. A., Judson, R. S., Knight, J. R., Lockshon, D.,<br />

Narayan, V., Sr<strong>in</strong>ivasan, M., Pochart, P., et al. (2000). A comprehensive analysis <strong>of</strong> prote<strong>in</strong>prote<strong>in</strong><br />

<strong>in</strong>teractions <strong>in</strong> Saccharomyces cerevisiae. Nature 403, 623-627.<br />

Van Dyck, E., Foury, F., Stillman, B., and Brill, S. J. (1992). A s<strong>in</strong>gle-stranded DNA b<strong>in</strong>d<strong>in</strong>g<br />

prote<strong>in</strong> required for mitochondrial DNA replication <strong>in</strong> S. cerevisiae is homologous to E. coli SSB.<br />

EMBO J 11, 3421-3430.<br />

Versele, M., Lemaire, K., and Thevele<strong>in</strong>, J. M. (2001). Sex and sugar <strong>in</strong> <strong>yeast</strong>: two dist<strong>in</strong>ct GPCR<br />

systems. EMBO Rep 2, 574-579.<br />

Vershon, A. K., Holl<strong>in</strong>gsworth, N. M., and Johnson, A. D. (1992). Meiotic <strong>in</strong>duction <strong>of</strong> the <strong>yeast</strong><br />

HOP1 gene is controlled by positive and negative regulatory sites. Mol Cell Biol 12, 3706-3714.<br />

Vidal, M., and Gaber, R. F. (1991). RPD3 encodes a second factor required to achieve maximum<br />

positive and negative transcriptional states <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 11, 6317-<br />

6327.<br />

Vidan, S., and Mitchell, A. P. (1997). Stimulation <strong>of</strong> <strong>yeast</strong> meiotic gene expression by the<br />

glucose-repressible prote<strong>in</strong> k<strong>in</strong>ase Rim15p. Mol Cell Biol 17, 2688-2697.<br />

Wang, H., and Stillman, D. J. (1993). <strong>Transcriptional</strong> repression <strong>in</strong> Saccharomyces cerevisiae by<br />

a SIN3-LexA fusion prote<strong>in</strong>. Mol Cell Biol 13, 1805-1814.<br />

Ward, M. P., Gimeno, C. J., F<strong>in</strong>k, G. R., and Garrett, S. (1995). SOK2 may regulate cyclic AMPdependent<br />

prote<strong>in</strong> k<strong>in</strong>ase-stimulated growth and pseudohyphal development by repress<strong>in</strong>g<br />

transcription. Mol Cell Biol 15, 6854-6863.<br />

Washburn, B. K., and Esposito, R. E. (2001). Identification <strong>of</strong> the S<strong>in</strong>3-B<strong>in</strong>d<strong>in</strong>g Site <strong>in</strong> Ume6<br />

Def<strong>in</strong>es a Two-Step Process for Conversion <strong>of</strong> Ume6 from a <strong>Transcriptional</strong> Repressor to an<br />

Activator <strong>in</strong> Yeast. Mol Cell Biol 21, 2057-2069.<br />

Xiao, Y., and Mitchell, A. P. (2000). Shared roles <strong>of</strong> <strong>yeast</strong> glycogen synthase k<strong>in</strong>ase 3 family<br />

members <strong>in</strong> nitrogen-responsive phosphorylation <strong>of</strong> meiotic regulator ume6p [In Process<br />

Citation]. Mol Cell Biol 20, 5447-5453.<br />

Xie, J., Pierce, M., Gailus-Durner, V., Wagner, M., W<strong>in</strong>ter, E., and Vershon, A. K. (1999). Sum1<br />

and Hst1 repress middle sporulation-specific gene expression dur<strong>in</strong>g mitosis <strong>in</strong> Saccharomyces<br />

cerevisiae. Embo J 18, 6448-6454.<br />

Xu, L., Ajimura, M., Padmore, R., Kle<strong>in</strong>, C., and Kleckner, N. (1995). NDT80, a <strong>meiosis</strong>-specific<br />

gene required for exit from pachytene <strong>in</strong> Saccharomyces cerevisiae. Mol Cell Biol 15, 6572-6581.<br />

Xu, W., and Mitchell, A. P. (2001). Yeast PalA/AIP1/Alix Homolog Rim20p Associates with a<br />

PEST-Like Region and Is Required for Its Proteolytic Cleavage. J Bacteriol 183, 6917-6923.<br />

57


Yoshida, M., Kawaguchi, H., Sakata, Y., Kom<strong>in</strong>ami, K., Hirano, M., Shima, H., Akada, R., and<br />

Yamashita, I. (1990). Initiation <strong>of</strong> <strong>meiosis</strong> and sporulation <strong>in</strong> Saccharomyces cerevisiae requires a<br />

novel prote<strong>in</strong> k<strong>in</strong>ase homologue. Mol Gen Genet 221, 176-186.<br />

Yoshimoto, H., Wada, H., and Yamashita, I. (1993). <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> <strong>meiosis</strong><strong>in</strong>duc<strong>in</strong>g<br />

IME1 and IME2 genes by GAM gene products <strong>in</strong> Saccharomyces cerevisiae. Biosci<br />

Biotech Biochem 57, 1784-1787.<br />

Zhan, X. L., Hong, Y., Zhu, T., Mitchell, A. P., Deschenes, R. J., and Guan, K. L. (2000).<br />

Essential functions <strong>of</strong> prote<strong>in</strong> tyros<strong>in</strong>e phosphatases ptp2 and ptp3 and rim11 tyros<strong>in</strong>e<br />

phosphorylation <strong>in</strong> saccharomyces cerevisiae <strong>meiosis</strong> and sporulation [In Process Citation]. Mol<br />

Biol Cell 11, 663-676.<br />

58


Figure legends<br />

Fig. 1. Developmental choices <strong>of</strong> MATa/MATα diploid cells <strong>of</strong> Saccharomyces cerevisiae.<br />

In the presence <strong>of</strong> carbon and nitrogen sources cells adopt the <strong>yeast</strong> form morphology; upon<br />

nitrogen limitation and <strong>in</strong> the presence <strong>of</strong> high levels <strong>of</strong> glucose, a dimorphic transition to a<br />

filamentous growth rem<strong>in</strong>iscent <strong>of</strong> hyphae takes places. In the absence <strong>of</strong> nitrogen and glucose<br />

and the presence <strong>of</strong> acetate as the sole carbon source cell enter the meiotic cycle, form<strong>in</strong>g four<br />

haploid spores engulfed <strong>in</strong> a sac, the ascus.<br />

Fig. 2. A transcriptional cascade governs <strong>in</strong>itiation <strong>of</strong> <strong>meiosis</strong>.<br />

The meiotic signals, i.e. the presence <strong>of</strong> Mata1 and Matα2, the absence <strong>of</strong> glucose and nitrogen<br />

and the presence <strong>of</strong> acetate leads to G1 arrest as well as to the expression and activity <strong>of</strong> Ime1.<br />

IME1 encodes a transcriptional activator required for the transcription <strong>of</strong> early <strong>meiosis</strong>-specific<br />

genes (EMG). EMG are <strong>in</strong>volved with premeiotic DNA replication and meiotic recomb<strong>in</strong>ation.<br />

Ndt80 and Ime2, a transcriptional activator and a prote<strong>in</strong> k<strong>in</strong>ase, whose transcription depends on<br />

Ime1, are required for the transcription <strong>of</strong> middle <strong>meiosis</strong>-specific genes (MMG). MMG are<br />

<strong>in</strong>volved with nuclear division and spore formation. The transcription <strong>of</strong> the late <strong>meiosis</strong>-specific<br />

genes (LMG) depends on Ime1 and Ime2, and these genes are required for spore maturation. The<br />

nutrient signal has a direct affect also on the activity <strong>of</strong> Ime2 and the transcription <strong>of</strong> LMG.<br />

An arrow represents a positive role, a l<strong>in</strong>e a negative role. A scissors symbolizes the negative<br />

feedback role <strong>of</strong> Ime2 <strong>in</strong> regulat<strong>in</strong>g degradation <strong>of</strong> Ime1.<br />

Fig. 3. Schematic structure <strong>of</strong> IME1 5’ untranslated region.<br />

The regulated transcription <strong>of</strong> IME1 is mediated by the comb<strong>in</strong>atorial effect <strong>of</strong> dist<strong>in</strong>ct elements.<br />

The MAT signal mediates repression activity <strong>of</strong> 2 elements UCS3 and UCS4. The carbon source<br />

signal is transmitted to four elements. UCS1, UASru and IREu function as repression elements <strong>in</strong><br />

the presence <strong>of</strong> glucose. In addition, UASru IREu, as well as UASrm function as activation<br />

elements <strong>in</strong> the absence <strong>of</strong> glucose and the presence <strong>of</strong> acetate as the sole carbon source. UCS1<br />

functions as a negative element <strong>in</strong> the presence <strong>of</strong> nitrogen.<br />

Filled boxs - elements required for transcriptional activation. Open boxs – elements whose<br />

function is only to repress transcription. A positive role is marked with an arrow, a negative role<br />

by a l<strong>in</strong>e. Larger effects are denoted by heavier l<strong>in</strong>es while lesser effects are denoted by slender<br />

l<strong>in</strong>es.<br />

59


Fig. 4. The function <strong>of</strong> positive and negative elements regulat<strong>in</strong>g the transcription <strong>of</strong> IME1.<br />

IME1 with various portions <strong>of</strong> its 5’ untranslated region were fused at the sixth am<strong>in</strong>o acid to the<br />

E.coli lacZ gene. The chimeric genes were <strong>in</strong>tegrated at the LEU2 loci <strong>of</strong> a MATa/MATα diploid<br />

(stra<strong>in</strong> Y4122, S288C background). Samples were taken from 1x10 7 cells grown <strong>in</strong> either SD<br />

(synthetic glucose) or PSP2 (SA – synthetic acetate (Kassir and Simchen, 1991). In addition, cells<br />

grown <strong>in</strong> PSP2 to 1x10 7 were washed once <strong>in</strong> water and resuspended <strong>in</strong> SPM (sporulation media<br />

(Kassir and Simchen, 1991). Samples were taken to extract prote<strong>in</strong>s and measure lacZ levels after<br />

3 and 6 h <strong>in</strong> SPM. The level <strong>of</strong> β-Galactosidase is given <strong>in</strong> Miller units. The results are the<br />

averages <strong>of</strong> three to five <strong>in</strong>dependent transformants. Standard deviations were less than 10%. The<br />

studied elements are marked as filled boxes. A nested deletion is marked by a l<strong>in</strong>e. NT – not<br />

tested.<br />

Fig. 5. Cdc25 transmits the nitrogen signal that modulates the repression activity <strong>of</strong> UCS1, a<br />

URS element <strong>in</strong> IME1 5’ region.<br />

UCS1 was <strong>in</strong>serted between HIS4 UAS and the TATA box <strong>in</strong> a his4-lacZ chimera. The level <strong>of</strong><br />

β-Galactosidase was determ<strong>in</strong>ed <strong>in</strong> three to five <strong>in</strong>dependent transformants. Standard deviations<br />

were less than 10%. The results are given as relative levels <strong>in</strong> comparison (<strong>in</strong> each case) to the<br />

level <strong>of</strong> expression <strong>of</strong> UASHIS4-his4-lacZ (control).<br />

Insertion <strong>of</strong> UCS1 <strong>in</strong> the heterologous UASHIS4-his4-lacZ chimera leads to a substantial reduction<br />

<strong>in</strong> its expression <strong>in</strong> vegetative growth conditions. Partial relief <strong>of</strong> repression is observed upon<br />

nitrogen depletion, suggest<strong>in</strong>g that the activity <strong>of</strong> UCS1 as a URS element is mediated by<br />

nitrogen. Depletion <strong>of</strong> Cdc25 by shift<strong>in</strong>g cdc25-5 cells to the non-permissive temperature leads to<br />

relief <strong>of</strong> repression <strong>in</strong> vegetative growth media, and has no effect upon nitrogen depletion (SPM),<br />

suggest<strong>in</strong>g that Cdc25 transmits the nitrogen signal to UCS1.<br />

Fig. 6: The UAS activity <strong>of</strong> the IME1 elements UASru, IREu and UASrm is modulated by<br />

the carbon source.<br />

His4-lacZ fusions carry<strong>in</strong>g various elements from IME1 5’ untranslated region were <strong>in</strong>tegrated at<br />

the LEU2 loci <strong>of</strong> a MATa/MATα diploid (stra<strong>in</strong> Y4122, S288C background). Cells were grown <strong>in</strong><br />

either SD (synthetic glucose) or PSP2 (SA – synthetic acetate (Kassir and Simchen, 1991). In<br />

addition, cells grown <strong>in</strong> PSP2 to 1x10 7 were washed once <strong>in</strong> water and resuspended <strong>in</strong> SPM<br />

[sporulation media (Kassir and Simchen, 1991)] and <strong>in</strong>cubated for 6 hours. Samples were taken<br />

from 1x10 7 cells/ml to extract prote<strong>in</strong>s and measure lacZ levels. The level <strong>of</strong> β-Galactosidase is<br />

60


given <strong>in</strong> Miller units. The results are the averages <strong>of</strong> three to five <strong>in</strong>dependent transformants.<br />

Standard deviations were less than 10%.<br />

UASru, IREu and UASrm support the expression <strong>of</strong> his4-lacZ, suggest<strong>in</strong>g that all three elements<br />

possess a UAS activity. The UAS activity is low <strong>in</strong> SD, and <strong>in</strong>creased activity is observed <strong>in</strong> SA<br />

and SPM media, suggest<strong>in</strong>g that their activity is repressed by glucose and/or depends on acetate.<br />

The IREd element that is homologus to IREu (see Fig. 8) shows very low UAS activity.<br />

Fig. 7. cAMP reduces sporulation and the expression <strong>of</strong> IME1 through the IREu element.<br />

MATa/MATα diploid cells (stra<strong>in</strong> Y4122, S288C background) carry<strong>in</strong>g ime1-lacZ or his4-lacZ<br />

fusions with various elements from IME1 5’ untranslated region <strong>in</strong>tegrated at the LEU2 loci.<br />

Cells were grown <strong>in</strong> PSP2 [synthetic acetate (Kassir and Simchen, 1991)] with or without 10mM<br />

cAMP to 1x10 7 cells/ml, washed once <strong>in</strong> water and resuspended <strong>in</strong> SPM [sporulation media<br />

(Kassir and Simchen, 1991)] with or without 10mM cAMP, respectively. Samples were taken at 6<br />

hours <strong>in</strong> SPM. The level <strong>of</strong> β-Galactosidase is given <strong>in</strong> Miller units. The results are the averages<br />

<strong>of</strong> three to five <strong>in</strong>dependent transformants. Standard deviations were less than 10%. At 24 hours<br />

<strong>in</strong> SPM samples were taken to count the percentage <strong>of</strong> asci. Results are given as relative levels.<br />

Addition <strong>of</strong> cAMP leads to a reduction <strong>in</strong> sporulation (72.6% and 49.8% <strong>in</strong> the absence and<br />

presence <strong>of</strong> cAMP, respectively) as well as a reduction <strong>in</strong> the expression <strong>of</strong> ime1-lacZ. We<br />

assume that the m<strong>in</strong>or effect is due to the function <strong>of</strong> the cAMP specific phosphodiesterases.<br />

Addition <strong>of</strong> cAMP reduces the UAS activity <strong>of</strong> IREu while it has no effect on the activity <strong>of</strong><br />

either UASru or UASrm.<br />

Fig. 8. IREu and IREd are almost identical repeats <strong>in</strong> IME1 5’ region carry<strong>in</strong>g the known<br />

UAS sequences STRE and SCB.<br />

Sequence alignments <strong>of</strong> IREu (upstream) and IREd (downstream) to the stress response element<br />

(STRE) and the Swi4/6 Cycle box (SCB). We suggest that Msn2 b<strong>in</strong>ds the STRE element <strong>in</strong><br />

IREu s<strong>in</strong>ce Msn2 b<strong>in</strong>ds IREu as well as STRE elements <strong>in</strong> stress response genes. We further<br />

suggest that Sok2 b<strong>in</strong>ds the SCB element s<strong>in</strong>ce it is highly homologous to the DNA-b<strong>in</strong>d<strong>in</strong>g<br />

doma<strong>in</strong> <strong>of</strong> Swi4 that b<strong>in</strong>ds such elements, and s<strong>in</strong>ce genetic analysis reveals that it b<strong>in</strong>ds IREu<br />

(see text). The physical association between Sok2 and Msn2 suggests that they b<strong>in</strong>d the DNA as a<br />

heterodimer.<br />

61


Fig. 9. A schematic structure <strong>of</strong> IME1 5’ untranslated region show<strong>in</strong>g the positive and<br />

negative transcription factors which regulate its expression.<br />

The regulated transcription <strong>of</strong> IME1 is mediated by the comb<strong>in</strong>atorial effect <strong>of</strong> dist<strong>in</strong>ct elements.<br />

The MAT signal is transmitted to UCS4 by Rme1. The activity <strong>of</strong> Rme1 as a repressor requires<br />

Rgr1 and S<strong>in</strong>4. The cAMP/PKA signal transduction pathway transmits the glucose signal to IREu<br />

through two transcription factors, Sok2 and Msn2. In glucose growth media phosphorylation <strong>of</strong><br />

Sok2 by PKA promotes its repression activity, whereas phosphorylation <strong>of</strong> Msn2 sequesters the<br />

prote<strong>in</strong> <strong>in</strong> the cytoplasm. In acetate growth media Sok2 repression is relieved by a decrease <strong>in</strong> the<br />

activity <strong>of</strong> PKA, as well as by the function <strong>of</strong> Ime1. The <strong>in</strong>creased b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> Msn2 to IREu leads<br />

to transcriptional activation. Yhp1 b<strong>in</strong>ds to UASv, however, its activity is not required for the<br />

transcription <strong>of</strong> IME1. Nitrogen, via Cdc25, regulates the URS activity <strong>of</strong> UCS1. It is not known<br />

whether the effect <strong>of</strong> Cdc25 is mediated through the cAMP/PKA and/or MAPK pathways.<br />

Filled box - elements required for transcriptional activation. Open box – elements that whose<br />

function is only to repress transcription. A positive role is marked with an arrow, a negative role<br />

<strong>in</strong> l<strong>in</strong>e. Dashed l<strong>in</strong>es and a question mark – prelim<strong>in</strong>ary data.<br />

Fig. 10. The expression and activity <strong>of</strong> Ime1 are regulated by nutrients.<br />

Schematic representation on how glucose and nitrogen regulates the availability and activity <strong>of</strong><br />

Ime1. The nucleus is depicted between the two circular l<strong>in</strong>es, and cell membrane by a heavy l<strong>in</strong>e.<br />

A positive role is marked with an arrow, a negative role by a straight l<strong>in</strong>e. The transcription <strong>of</strong><br />

IME1 is <strong>in</strong>hibited <strong>in</strong> the presence <strong>of</strong> glucose and nitrogen. Translation <strong>of</strong> IME1 mRNA is<br />

<strong>in</strong>hibited <strong>in</strong> the presence <strong>of</strong> a nitrogen source. In the presence <strong>of</strong> nitrogen, Ime1is sequestered<br />

from the nuclei due to phosphorylation by Cdc28/Cln. Interaction <strong>of</strong> Ime1 with Ume6, and<br />

consequently transcriptional activation <strong>of</strong> early <strong>meiosis</strong>-specific genes (EMG) is <strong>in</strong>hibited by<br />

glucose and nitrogen.<br />

Fig. 11. Conditions lead<strong>in</strong>g to the choice between silenc<strong>in</strong>g and expression <strong>of</strong> early <strong>meiosis</strong>specific<br />

genes.<br />

A. Vegetative growth conditions B. Meiotic conditions.<br />

Ume6 b<strong>in</strong>ds to the URS1 element <strong>in</strong> early-<strong>meiosis</strong> specific genes (EMG) under all growth<br />

conditions. In vegetative growth conditions Ume6 recruits to repression complexes, S<strong>in</strong>3/Rpd3<br />

and Isw2, lead<strong>in</strong>g to histone H4 deacetylase and chromat<strong>in</strong> remodel<strong>in</strong>g, respectively, and<br />

silenc<strong>in</strong>g. Under these conditions Rim15 and most probably Rim11 are non-active. The activity<br />

<strong>of</strong> Rim15 is repressed by PKA (Tpk1,2,3) phosphorylation. Under meiotic conditions (acetate<br />

62


media and nitrogen depletion) phosphorylation <strong>of</strong> Ime1and Ume6 by Rim11 and Rim15 promotes<br />

the <strong>in</strong>teraction <strong>of</strong> Ime1 with Ume6. Ime1 function is required to relieve repression by S<strong>in</strong>3 and to<br />

activate transcription. In addition, transcriptional activation depends on histone acetylation by<br />

Gcn5.<br />

Fig. 12. Ime2 is highly homologous to hCDK2.<br />

A. Sequence alignment <strong>of</strong> Ime2 to hCDK2. B, Ribbon depiction <strong>of</strong> the crystal structure <strong>of</strong><br />

hCDK2 (Prote<strong>in</strong> Data Bank structure code 1HCL). C. Ribbon depiction <strong>of</strong> the homology based<br />

model <strong>of</strong> Ime2 based on the CDK structure. The follow<strong>in</strong>g stretches <strong>of</strong> residues are colored for<br />

clarity: the ATP b<strong>in</strong>d<strong>in</strong>g site (yellow), The PSTAIRE site (orange), the Catalytic site (red) and the<br />

T-loop (blue). The Ime2 model was prepared us<strong>in</strong>g the 3D-PSSM model<strong>in</strong>g algorithm (Kelley et<br />

al., 2000). The figures were prepared us<strong>in</strong>g InsightII (Accelrys).<br />

Fig. 13. <strong>Transcriptional</strong> <strong>regulation</strong> <strong>of</strong> middle <strong>meiosis</strong>-specific genes.<br />

Ndt80 is the transcriptional activator b<strong>in</strong>d<strong>in</strong>g to the MSE element present <strong>in</strong> the 5’ region <strong>of</strong> all<br />

middle <strong>meiosis</strong>-specific genes (MMG). A subset <strong>of</strong> MMG carries an MSE* site (SMK1 <strong>in</strong><br />

comparison to CLB1) that can be bound by Sum1. Sum1 associates with Hst1 that functions as<br />

histone deacetylase, thus lead<strong>in</strong>g to silenc<strong>in</strong>g under vegetative growth conditions and at early<br />

meiotic times. Under meiotic conditions relief <strong>of</strong> repression is due to degradation <strong>of</strong> Sum1 by<br />

Ime2. Sum1 also represses the transcription <strong>of</strong> NDT80, s<strong>in</strong>ce it carries both an MSE and MSE*<br />

elements. In addition, NDT80 carries two URS1 elements that are bound by Ume6. In vegetative<br />

growth media Ume6 recruits the S<strong>in</strong>3/Rpd3 histone deacetylase complex that represses<br />

transcription. Under meiotic conditions Ume6 recruits Ime1 that can activate transcription, but<br />

only <strong>in</strong> the absence <strong>of</strong> Sum1. Ime2 phosphorylates Ndt80, and this phosphorylation may<br />

contribute to the complete transcriptional activity <strong>of</strong> Ndt80. In the absence <strong>of</strong> meiotic<br />

recomb<strong>in</strong>ation the pachytene checkpo<strong>in</strong>t <strong>in</strong>hibits degradation <strong>of</strong> Sum1 and phosphorylation <strong>of</strong><br />

Ndt80. Consequently, the level <strong>of</strong> Ndt80 is reduced, and the unphosphorylated Ndt80 is impaired<br />

<strong>in</strong> activat<strong>in</strong>g the transcription <strong>of</strong> MMG.<br />

Fig. 14. Positive and negative feedback loops control <strong>meiosis</strong>.<br />

Ime1 shows positive auto<strong>regulation</strong> that is required to relieve repression mediated by Sok2 and<br />

thus for its high-level transcription. Expression <strong>of</strong> early <strong>meiosis</strong>-specific genes (EMG) requires<br />

relief <strong>of</strong> repression <strong>of</strong> S<strong>in</strong>3 and transcriptional activation, two functions that are promoted by<br />

Ime1. The transcription <strong>of</strong> middle genes (MMG) depends on Ime2 and Ndt80. <strong>Transcriptional</strong><br />

63


activation by Ime2 is due to phosphorylation <strong>of</strong> the transcriptional activator, Ndt80, and<br />

degradation <strong>of</strong> the transcriptional repressor, Sum1. Ime2 exhibits a negative feedback role, it is<br />

required to shut down the transcription <strong>of</strong> IME1. S<strong>in</strong>ce Ime2 directly phosphorylates Ime1, and<br />

s<strong>in</strong>ce this phosphorylation is required for degradation <strong>of</strong> Ime1 by the 26S proteasome, it was<br />

suggested that the reduction <strong>in</strong> the transcription <strong>of</strong> IME1 as well as EMG is due to the absence <strong>of</strong><br />

Ime1 and reestablishment <strong>of</strong> repression by Sok2 and S<strong>in</strong>3, respectively. The transcription <strong>of</strong><br />

MMG is repressed by the pachytene checkpo<strong>in</strong>t that <strong>in</strong>hibits phosphorylation <strong>of</strong> Ndt80 and<br />

degradation <strong>of</strong> Sum1.<br />

64


Table 1. List <strong>of</strong> positive and negative regulators <strong>of</strong> transcription <strong>of</strong> <strong>meiosis</strong>-specific genes.<br />

Name Brief description <strong>of</strong> function<br />

ABF1 An essential DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> required for DNA replication and the transcription <strong>of</strong> several<br />

<strong>meiosis</strong>-specific genes carry<strong>in</strong>g its b<strong>in</strong>d<strong>in</strong>g site (UASH).<br />

AMA1 A <strong>meiosis</strong>-specific subunit <strong>of</strong> APC/C that is required for degradation <strong>of</strong> CLB1 <strong>in</strong> the meiotic cycle.<br />

Required for the transcription <strong>of</strong> LMG.<br />

BCY1 The regulatory subunit <strong>of</strong> PKA. Positive regulator for IME1 transcription and sporulation.<br />

CDC25 Ras GDP/GTP exchange factor. Positive activator <strong>of</strong> PKA, negative regulator <strong>of</strong> IME1<br />

transcription. Transmits the nitrogen signal to UCS1 and the glucose signal to IREu elements <strong>in</strong><br />

IME1 promoter.<br />

CDC28 Cycl<strong>in</strong> dependent k<strong>in</strong>ase that regulates <strong>in</strong>itiation and progression <strong>in</strong> the mitotic cell cycle. In a<br />

complex with Cln phosphorylates Ime1 and sequesters it out <strong>of</strong> the nuclei. Required to activate<br />

Ime2 at the time cells enter the meiotic division.<br />

CDC35 Adenylate cyclase. Positive activator <strong>of</strong> PKA, negative regulator <strong>of</strong> IME1 transcription and<br />

sporulation<br />

GCN5 Histone acetylase required for the transcription <strong>of</strong> EMG.<br />

GPA2 A Gα prote<strong>in</strong> that associates, when GTP bound, with Ime2. An <strong>in</strong>hibitor <strong>of</strong> Ime2 k<strong>in</strong>ase activity.<br />

HST1 Histone deacetylase. Represses the transcription <strong>of</strong> MMG follow<strong>in</strong>g recruitment by Sum1.<br />

IDS2 A positive regulator <strong>of</strong> Ime2 activity.<br />

IME1 The master regulator gene <strong>of</strong> <strong>meiosis</strong>. A transcriptional activator; recruited by Ume6 to the URS1<br />

element <strong>in</strong> EMG, and is required for their transcription. The transcription <strong>of</strong> IME1 depends on the<br />

presence <strong>of</strong> Mata1 and Matα2 as well as on glucose and nitrogen depletion. Translation and<br />

activity <strong>of</strong> Ime1 are regulated by nutrients.<br />

IME2 An early <strong>meiosis</strong>-specific gene encod<strong>in</strong>g ser<strong>in</strong>e/threon<strong>in</strong>e k<strong>in</strong>ase, homologous to the human CDK2.<br />

Required for timely and high level transcription <strong>of</strong> EMG, and for the transcription <strong>of</strong> MMG and<br />

LMG. Negative regulator <strong>of</strong> Ime1 stability.<br />

IME4 A positive transcriptional regulator <strong>of</strong> IME1 whose transcription depends on the presence <strong>of</strong> Mata1<br />

and Matα2 and nitrogen depletion.<br />

ISW2 In a complex with Itc1 functions as an ATP dependent chromat<strong>in</strong>-remodel<strong>in</strong>g factor. Required for<br />

the repression <strong>of</strong> EMG under vegetative growth conditions. Physically associate with Ume6 that<br />

recruits it to URS1 elements.<br />

ITC1 In a complex with Isw2 functions as an ATP dependent chromat<strong>in</strong>-remodel<strong>in</strong>g factor. Required for<br />

the repression <strong>of</strong> EMG under vegetative growth conditions.<br />

65


MCK1 Prote<strong>in</strong> k<strong>in</strong>ase required for efficient transcription <strong>of</strong> IME1, expression <strong>of</strong> EMG, MMG and spore<br />

maturation. Homologous to mammalian glycogen synthase k<strong>in</strong>ase 3 and S. cerevisiae, RIM11,<br />

YOL128c, and MRK1. Required for phosphorylation <strong>of</strong> Ume6 and sporulation <strong>in</strong> cells deleted for<br />

RIM11 and MRK1.<br />

MDS3 Negative regulator <strong>of</strong> IME1 transcription. Redundant function with its homolog, PMD1<br />

MSN2/4 Z<strong>in</strong>c f<strong>in</strong>ger transcriptional activators, which share redundant function and b<strong>in</strong>d STRE elements.<br />

Negative regulators <strong>of</strong> mitosis. Positive regulators <strong>of</strong> <strong>meiosis</strong> and IME1 transcription through the<br />

IREu element <strong>in</strong> IME1 promoter. Targets <strong>of</strong> the cAMP/PKA pathway.<br />

MRK1 Prote<strong>in</strong> k<strong>in</strong>ase homologous to mammalian glycogen synthase k<strong>in</strong>ase 3 and S. cerevisiae, RIM11,<br />

YOL128c, and MCK1.Required for phosphorylation <strong>of</strong> Ume6 and sporulation <strong>in</strong> cells deleted for<br />

RIM11 and MCK1.<br />

NDT80 A <strong>meiosis</strong>-specific transcriptional activator that b<strong>in</strong>ds the MSE element <strong>in</strong> MMG. An early MMG.<br />

PMD1 Negative regulator <strong>of</strong> IME1 transcription. Redundant function with its homolog, MDS3<br />

PTP2,3 Tyros<strong>in</strong>e phosphatases, required for Mck1 dephosphorylation as well as for high-level transcription<br />

<strong>of</strong> IME1 and IME2, and for expression <strong>of</strong> MMG and LMG.<br />

RAS2 Small G prote<strong>in</strong> that activates the cAMP/PKA and the MAPK signal pathways. Negative regulator<br />

<strong>of</strong> IME1 transcription.<br />

RES1 A dom<strong>in</strong>ant mutation <strong>in</strong> this gene bypasses MAT <strong>regulation</strong> for the transcription <strong>of</strong> IME1.<br />

RGR1 A component <strong>of</strong> the RNA polymerase SRB mediator complex. Required for the repression activity<br />

<strong>of</strong> Rme1.<br />

RIM1 A C2H2 z<strong>in</strong>c f<strong>in</strong>ger prote<strong>in</strong> required for high-level expression <strong>of</strong> IME1, and efficient sporulation.<br />

Required for mitochondria DNA replication and ma<strong>in</strong>tenance. Activity is regulated by cleavage<br />

that depends on Rim8, Rim9, Rim13 and Rim20.<br />

RIM4 An early <strong>meiosis</strong>-specific gene required for high-level expression <strong>of</strong> EMG. Sequence identifies two<br />

RNA recognition motifs that are essential for function.<br />

RIM8 Positive regulator <strong>of</strong> IME1 transcription. Required for activat<strong>in</strong>g Rim1.<br />

RIM9 Positive regulator <strong>of</strong> IME1 transcription. Required for activat<strong>in</strong>g Rim1.<br />

RIM11 Tyros<strong>in</strong>e/ser<strong>in</strong>e/threon<strong>in</strong>e k<strong>in</strong>ase that is homologous to mammalian glycogen synthase k<strong>in</strong>ase 3 and<br />

S. cerevisiae, MCK1, YOL128c, and MRK1.Required for the <strong>in</strong>teraction between Ime1 and Ume6,<br />

and consequently for the transcription <strong>of</strong> EMG. Associates and phosphorylates both Ime1 and<br />

Ume6.<br />

RIM13 Positive regulator <strong>of</strong> IME1 transcription. Required for activat<strong>in</strong>g Rim1.<br />

RIM20 Positive regulator <strong>of</strong> IME1 transcription. Required for activat<strong>in</strong>g Rim1.<br />

66


RIM15 A ser<strong>in</strong>e/threon<strong>in</strong>e k<strong>in</strong>ase whose activity is <strong>in</strong>hibited by PKA phosphorylation. Positive regulator<br />

for the transcription <strong>of</strong> IME1 and for the <strong>in</strong>teraction between Ime1 and Ume6. Phosphorylates<br />

Ume6.<br />

RME1 A Zn f<strong>in</strong>ger DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> that represses the transcription <strong>of</strong> IME1 through the UCS4<br />

element <strong>in</strong> IME1 promoter. The transcription <strong>of</strong> RME1 is negatively regulated by the Mata1/Matα2<br />

complex.<br />

ROX3 A component <strong>of</strong> the RNA polymerase SRB mediator complex. Required for repress<strong>in</strong>g the<br />

transcription <strong>of</strong> mid-late <strong>meiosis</strong>-specific genes.<br />

RPD3 Histone deacetylase. Required for repression <strong>of</strong> EMG <strong>in</strong> media support<strong>in</strong>g vegetative growth.<br />

Positive regulator <strong>of</strong> MMG.<br />

SIN3 Negative regulator that is required for repression <strong>of</strong> EMG <strong>in</strong> media support<strong>in</strong>g vegetative growth.<br />

Recruits Rpd3 to EMG due to its association with both Ume6 and Rpd3. Positive regulator <strong>of</strong><br />

MMG.<br />

SIN4 A component <strong>of</strong> the RNA polymerase SRB mediator complex. Required for the repression activity<br />

<strong>of</strong> Rme1, and for repress<strong>in</strong>g the transcription <strong>of</strong> mid-late <strong>meiosis</strong>-specific genes.<br />

SNF1 Ser<strong>in</strong>e/threon<strong>in</strong>e prote<strong>in</strong> k<strong>in</strong>ase whose activity is <strong>in</strong>hibited by high levels <strong>of</strong> glucose. Required for<br />

the transcription <strong>of</strong> both IME1 and IME2, and for spore formation.<br />

SNF2 Component <strong>of</strong> the Swi/Snf transcriptional activation complex. Required for high-level expression<br />

<strong>of</strong> IME1.<br />

SOK2 <strong>Transcriptional</strong> repressor that b<strong>in</strong>ds SCB like sequences. Positive regulator <strong>in</strong> the mitotic cell cycle<br />

Negative regulator <strong>of</strong> pseudohyphal growth, <strong>meiosis</strong> and IME1 transcription. Sok2 function as a<br />

repressor depends on phosphorylation by PKA.<br />

SSN6 A general transcriptional repressor that regulates the transcription <strong>of</strong> IME1. Forms a repression<br />

complex with Tup1.<br />

SUM1 A transcriptional repressor <strong>of</strong> MMG that b<strong>in</strong>ds the MSE* element. Recruits histone deacetylase<br />

activity by association with Hst1.<br />

SWI1 Component <strong>of</strong> the Swi/Snf transcriptional activation complex. Required for high-level expression<br />

<strong>of</strong> IME1.<br />

TPK1,2,3 Three genes encod<strong>in</strong>g the catalytic subunit <strong>of</strong> PKA.<br />

TUP1 A general transcriptional repressor that regulates the transcription <strong>of</strong> IME1. Forms a repression<br />

complex with Ssn6.<br />

UME2 A component <strong>of</strong> the SRB subcomplex <strong>of</strong> RNA polymerase II holoenzyme. Negative regulator <strong>of</strong><br />

EMG transcription.<br />

67


UME3 A cycl<strong>in</strong> C homolog that associates with the cycl<strong>in</strong> dependent k<strong>in</strong>ase, Ume5. An <strong>in</strong>tegral<br />

component <strong>of</strong> the SRB subcomplex <strong>of</strong> RNA polymerase II holoenzyme. Required to repress<br />

transcription <strong>of</strong> EMG <strong>in</strong> vegetative growth media. Degraded <strong>in</strong> meiotic conditions.<br />

UME5 A cycl<strong>in</strong> dependent k<strong>in</strong>ase that associates with the cycl<strong>in</strong> C homolog, Ume3. An <strong>in</strong>tegral<br />

component <strong>of</strong> the SRB subcomplex <strong>of</strong> RNA polymerase II holoenzyme. Required to repress<br />

transcription <strong>of</strong> EMG <strong>in</strong> vegetative growth media.<br />

UME6 A C6Zn2 DNA-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>. B<strong>in</strong>ds to URS1 element. Functions as a negative regulator by<br />

recruit<strong>in</strong>g the S<strong>in</strong>3/Rpd3 and Isw2 complexes, and as a positive regulator by recruit<strong>in</strong>g Ime1.<br />

YHP1 Homeodoma<strong>in</strong> prote<strong>in</strong> that b<strong>in</strong>ds to UASv, a specific region <strong>in</strong> IME1 5’ region. It is not required<br />

for the transcription <strong>of</strong> IME1.<br />

WTM1,2,3 <strong>Transcriptional</strong> repressors that repress the transcription <strong>of</strong> IME2 synergistically.<br />

YVH1 Tyros<strong>in</strong>e phosphatase whose transcription is <strong>in</strong>duced upon nitrogen depletions. Has redundant<br />

function with Ptp2 <strong>in</strong> regulat<strong>in</strong>g sporulation.<br />

68

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!