. 'r;.'. _ féy.‘1‘.;. r7 3 -» ’ 9559'? Pu.’ - . ’% I I "c ' - - n .45..-. u g . _ fimvm ‘. .1 It {a} -';‘*;x. 'i «mur‘ V an I flap." "~— .. 3?: . ‘ W I\ “‘ u 1;?” . "IL 11511?” x. L 5-7:... vs. 5r”; .4: v‘ ‘lit 0 £1 ‘ Wt!!! 1% T A". ~ ‘ u , . F" (u "i ‘ .: ’19:“. 9, .‘v 1w! 't?‘ ‘t a . its ‘1 “1...“... . 31::1 . -- “M «In ‘ -r '. ._-~.n- g...“- 2‘1}: ‘ x. 5 . 24" l ”‘4‘ 52.12.1953 '9 . _,y ‘ m} s: ‘ "Ar“? \ 31 ’w ‘5? :i ont‘l-‘ ‘3’ 731" r‘K“ "é 531331 -‘_':.--F-— . -: "1 ~,_"-.. ‘ - ”,an .. - ‘ ...- w...- ‘ . -_ v- ,. <..—..~. .. . v ‘ 'm-rv“. _ - . ’u p6?) llllIllalllllllllllllllillllllllllllliWilli“llllllllllllllll 31293 01820 0224 This is to certify that the dissertation entitled Structural and Functional Dissection of the Activation Domains of the Yeast Transcriptional Activator HAP4 presented by John L. Stebbins has been accepted towards fulfillment of the requirements for Ph.D . degree in Genetics firms {dajor prflssor Date (Meg/a? MS U is an Affirmative Action/Equal Opportunity Institution 0- 12771 LIBRARY Michigan State Unlverslty PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINE return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 1M chlRC/D‘qupfiG—au STRUCTURAL AND FUNCTIONAL DISSECT ION OF THE ACTIVATION DOMAINS OF THE YEAST TRANSCRIPTIONAL ACTIVATOR HAP4 By John L. Stebbins A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Genetics Program & Cell and Molecular Biology Program 1 999 ABSTRACT STRUCTURAL AND FUNCTIONAL DISSECT ION OF THE ACTIVATION DOMAINS OF THE YEAST TRANSCRIPTIONAL ACT IVATOR HAP4 By John L. Stebbins The yeast protein Hap4p activates the transcription of genes required for growth on nonferrnentable carbon sources. Like other transcriptional activators, Hap4p when overexpressed causes toxicity. Unlike other activators, this toxicity is not relieved by mutations in ADA2, ADA3, or GCN5, nor is the transcriptional activity of Hap4p reduced in strains lacking ADA2, ADA3, or GCN5. Moreover, Gal4 fusion proteins bearing the activation domain of Hap4 (but not those of VP16 or GCN4) were able to function in ada2 or 3015 mutant strains. These results suggest that the HAP4 activation domain is intrinsically ADA2, ADA3, and GCNS independent. Our goal is to define the essential features and mechanism of action of Hap4p as a prototypical ADA-independent activator. The activation domain of Hap4p has been previously mapped to amino acids 330- 554 (C-domain). In this work the boundaries of this C-terminal activation domain has been further defined by deletion mutational analysis. The ADA-independent activation function has been mapped to the 359-476 region of Hap4p. Within this minimal C- domain we have identified several clusters of hydrophobic amino acids that are critical for the trans-activation function of this domain. The simultaneous mutation to serine of He390/I‘rp39l/Lys393/Ieu394, Phe456/Tyr458, or Leu466/Met467 resulted in a greater than 10-fold fold reduction in activation potential of the 359-476 region of Hap4. Contrary to previous results, we found that Hap4p 1-330 (N -domain) is able to support growth of yeast on lactate medium. Moreover, when tethered to lexA, the Hap4p N-domain can activate a reporter gene with lexA binding sites upstream. These results imply that some region within amino acids 1-330 of Hap4p also possesses the ability to activate transcription. This function has been mapped by deletion mutational analysis to the 124-329 region of Hap4p. Transcriptional activation by this region is ADA- depcndent and is less efficient than the C-domain (aa 359-476) of Hap4p. The activation potential of the 124-329 region of Hap4 is dependent on Phel48/Leul49/Phe151 as the simultaneous mutation of these residues to serine resulted in a construct that is devoid of trans-activation potential. To identify mechanistic targets of the Hap4C subdomain, we designed genetic strategies based on screening or selecting for loss of lexA—Hap4 (aa330-554) activation function. While we have been unable to identify any such targets as a result of these genetic strategies, our attempts have resulted in the development of a new and promising strategy. We will attempt to suppress point mutations that reduce the trans-activation potential of the 359-476 region of Hap4 with a high copy yeast genomic library. We hypothesize that we will be able to identify a mechanistic target of the 359-476 region in this way. The genetic screens described to this point represent one strategy for identifying putative targets of the Hap4 activation domain. An alternative strategy is to test for genetic interactions between Hap4 and genes already implicated in co-activator mechanisms from other studies. We therefore tested the ability of Hap4 330-554 to activate transcription in yeast strains that lacked SPD, SPT7, SPT8, or SPTZO and found that lexA-Hap4 (aa330-554) is dependent on SPT7 and SP720 for its ability to function as a transcriptional activator. Copyright by .HDPHQIiEYTEIHBHVS 1 999 To my family, and to Kirsten with love. You are the most important people in my life! Meet triumph and disaster And treat those two imposters just the same Paraphrased from Rudyard Kipling’s “If" vi ACKNOWLEDGMENTS I am deeply indebted to Steve Triezenberg, as he has been both patient and stem with me as needed. In large part due to his efforts and examples I have been able to develop and grow as a scientist. I thank the members of my committee: Mike Thomashow, Dennis Thiele, Helmut “Bert” Bertrand, and Lenny Robbins. They have been singularly and collectively great. Their enthusiasm and effort has really helped the evolution of not only this project but my scientific-self as well. I would especially like to thank Dennis as he has always been an eager participant even though it meant that he had to drive to East Lansing from Ann Arbor. I thank members of the Triezenberg Lab past and present (Jeff, Eugene, Rath - the original nut case, Peter — the original grump, The Big Danish (SJ 0), Crazy-Mao, Greeny (DG), Marty, Lee, Susan, Kostas, Yuri, and Fan) for their companionship, friendship, help, and advice. I learned a lot from each one of you. THANKS! I have made many life-long friends during my time here and thank them for their role in helping me through. This list includes among others: Tubby, Tuffy, Tim, Mark the “Dr. Dr.”, Luc, Odette, and especially my local—H: Spencaaaaaahh. I would also like to thank the members of the greater yeast community. Without the generous help of people like Fred Winston and members of the Winston Lab, Shelley Berger, Reyes Candau, Craig Peterson, Dan Gottschling, Kevin Morano, Eric Andrulis, Jeff Smith, Susanne Kleff, and especially Dave McNabb (he has been extremely patient and generous as he has fielded a majority of my inquires) none of this would be possible. I thank my family for their support, encouragement, and cease-fire on the “When are you going to graduate?” questions. And lastly, I thank Kirsten. I owe her a great deal vii for all that she has done to help me attain this goal. She has spent far too much time alone while I have been in the lab. Thanks and I love you! viii TABLE OF CONTENTS List of Figures xi List of Tables xiii List of Abbreviations xiv Chapter One Literature Review 1-29 Overview 1 Promoter Architecture 2 Promoter Context 7 RNA Polymerase II and Associated Factors ll Chromatin Remodeling Complexes 24 Hap2/3/4/5 27 Chapter Two Identification and Characterization of the Hap4 Activation Domains 30-72 Introduction 30 Experimental Methods 34 Deletion mutagenesis of Hap4 aa330-554 35 Deletion mutagenesis of Hap4 aa1-33O 37 B-Galactosidase Assays 37 Immunoblot Assays 37 Point mutant construction 38 Growth curve 39 Results 45 Discussion 64 Chapter Three Identification of Hag activation domain targetm 73-100 Introduction 73 Experimental Methods 77 EMS mutagenesis 77 Mutant isolation using S-Floroorotic Acid 78 Yeast strain production 79 High copy suppression of Hap4 point mutants 79 ix Results 82 L40/AMR70 82 BCYOS 84 JSYO3 86 High Copy Suppression 89 SPT3/7/8/20 94 Discussion 98 Chapter Four Future direction and mmfitives 101-106 Bibliography 107- l 31 LIST OF FIGURES Figure 1. Specific activities of deletion mutants of the Hap4 C-terminal domain (aa-330-554) ............................................................................................................... 47 Figure 2. Immunoblot analysis of yeast cells expressing deletion mutants of HAP4 gene fused to lexA. .......................................................................................... 49 Figure 3. The amino acid sequence of the Hap4 359—476 region .............................. 51 Figure 4. The contribution of hydrophobic residues to the activation potential of the Hap4 359-476 region ........................................................................................ 53 Figure 5. Immunoblot analysis of yeast cells expressing point mutants of the HAP4 gene fused to lexA ......................................................................................... 54 Figure 6. Hap4 (aa1-330) is capable of supporting growth in lactate media in the yeast strain DMYl48 ........................................................................................ 56 Figure 7. Specific activities of deletion mutants of Hap4 N-terminal domain (aal-330) .................................................................................................................... 57 Figure 8. The amino acid sequence of the Hap4 124-160 region .............................. 59 Figure 9. The contribution of hydrophobic residues to the activation potential of the Hap4 124-329 region ........................................................................................ 60 Figure 10. Effect of deletion of gcn5 on the ability of Hap4 derivatives to activate transcription of a lacZ gene ........................................................................... 62 Figure 11. Hap4 (aal-330) fails to support growth in lactate media in gcn5 yeast. .................................................................................................................. 63 Figure 12. Sequence comparison of portions of the transcriptional activation domains of VP16, GCN4, and Hap4 .......................................................................... 70 Figure 13. Schematic representation of the evolutionarily conserved sequences in KlHAP4 and ScHAP4 ......................................................................... 72 Figure 14. The ability of a spontaneously occun‘ing JSYO3 mutant to support trans-activation by lexA-Hap4 (aa330-554) .................................................. 91 Figure 15. The ability of a spontaneously occurring JSYO3 mutant to support trans-activation by lexA-Hap4 or lexA-GCN4. ............................................ 92 xi Figure 16. The effect of deletion on either SPT3, SP77, SPT8, or SP720 on the ability of lexA-Hap4 (aa330-554) to activate transcription ................. 96 Figure 17. Immunoblot analysis of Hap4 (aa330-476) fused to lexA in either wild type, Aspt7 or Aspt20 yeast cells ............................................................. 97 xii LIST OF TABLES Table l. Oligonucleotides used to construct Hap4 deletion mutants ........................ 40 Table 2. Mutagenic Oligonucleotides ........................................................................ 42 Table 3. Table of yeast strain genotypes ................................................................... 44 Table 4. Table of yeast strain genotypes .................................................................... 81 xiii EMS FOA GTF HAP4C HAP4N OD PAGE PBS PCR PIC SDS-PAGE TAF TBP LIST OF ABBREVIATIONS amino acid activation domain 3-amino- 1 ,2,4-triazole base pair C-terminal domain of RNA polymerase H ethyl methane sulfonate S-fluoroorotic acid general transcription factor heme activated protein carboxyl terminal domain of HAP4 amino terminal domain of HAP4 kiloDaltons RNA polymerase optical density polyacrylamide gel electrophoresis phosphate buffered saline polymerase chain reaction pre-initiation complex sodium-dodecyl-sulfate polyacrylamide gel electrophoresis TBP associated facter TATA-binding protein xiv TFII transcription factor, RNA polymerase H UAS upstream activating sequence URS upstream repressing sequence VP16 virion protein 16 of herpes simplex virus XV W W91 OVERVIEW The appropriate and timely response to environmental stimuli and the control and coordination of developmental processes and pathways are examples of how the accurate control of gene expression is required for the survival and growth of an organism. Gene expression is regulated through transcriptional and post-transcriptional mechanisms. While the decision whether or not to transcribe a given gene at any given moment is an obvious level of regulation, many other regulatory mechanisms exist such as mRNA stability, mRNA export, differential splicing, as well as post-translational protein modifications such as phosphorylation. Perturbation of the control of gene expression can lead to various diseases such as diabetes and cancer. Thus the understanding of how organisms maintain accurate control of gene expression is a topic of intense scientific investigation. In eukaryotes, transcription of the nuclear genome is the responsibility of three different RNA polymerases. RNA polymerase I synthesizes the large ribosomal RNAs. RNA polymerase II (RNAPII) is responsible for transcribing protein-coding genes and the production of some small nuclear RNAs. RNA polymerase III synthesizes the remaining small nuclear RNAs, as well as 5S ribosomal RNA and transfer RNAs. All three polymerases depend on auxiliary factors called general transcription factors (GTFs) for core promoter recognition. The GTFs, also referred to as basal factors, that are required for recruitment of RNAPII to its cognate promoters include TFIIA, TFIIB, TFIID, TFIIE, TFIIF, and TFIIH (where TFII stands for transcription factor for RNA polymerase fl). Together the GTFs and RNAPH and promoter DNA make up the pre- initiation complex (PIC). The typical core promoter elements can include a TATA-like sequence, located upstream of the transcription initiation site, and an initiator sequence (Inr) that encompasses the transcriptional start site. The core promoter directs the accurate initiation site, PIC complex formation site, and contributes to the overall level of transcription. Regulatory elements include both upstream elements such as upstream activating sequences (UAS) and upstream repressing sequences (URS) and bind activator or repressor proteins respectively. These regulatory elements direct the spatial and temporal patterns of gene-selective transcription. Since the focus of my work is the control of expression of protein-encoding genes in the budding yeast Saccharomyces cerevisiae, I focus on the details of RN APH function in this organism for the remainder of this review. I will emphasize the role of yeast genetics in establishing current models and in defining the roles of the known factors. PROMOTER ARCHITECTURE RN APII dependent promoters in Saccharomyces cerevisiae are typically composite in nature containing combinations of five elements: upstream activating sequences (UAS), upstream repressing sequences (URS), TATA elements, sites of transcriptional initiation (Inr element), and homopolymeric (dA-dT) elements. UAS UAS elements are the sites for binding gene-specific activator proteins. Activator binding results in promoter stimulation in response to environmental stimuli. UAS elements are analogous to metazoan enhancer elements. However, while enhancers are able to function when located either 5’ or 3’ of the TATA element, yeast UAS elements are functional only when located 5’ of the TATA element in the few cases tested thus far (Guarente, et al., 1984; Struhl, 1984). Once associated with its cognate UAS element, a transcriptional activator protein facilitates an increase in mRN A production by RNAPII. Mechanisms employed by transcriptional activators will be discussed in detail in sections to follow. URS URS elements are the sites of binding for gene-specific transcriptional repressor proteins. Repressor production is a cellular response to environmental stimuli. Association of a repressor with its cognate URS binding site results in a reduction of mRNA production by RNAPH. Several different mechanisms for repression exist. Repressor proteins bound to URS elements can recruit co-repressor complexes such as Ssn6-Tupl. Ssn6Tupl appears to mediate repression by modifying promoter-proximal chromatin structures (Edmondson, et al., 1996; Gavin and Simpson, 1997; Huang, et al., 1997; Roth, 1995). When the repressor Ume6 binds to its cognate URS element it recruits the Sin3-de3 histone deacetylase complex (Kadosh and Struhl, 1997) which in turn removes acetyl groups from histone H4 and lysine 5 and 12 within histone H3 (Kadosh and Struhl, 1998). This alteration of histones H3 and H4 results in a repressive chromatin structure. Other known repression mechanisms include inhibition of activator UAS association, such as the mechanism used by the YB-l protein, which binds to single-stranded regions within the promoter, thus preventing loading and/or function of other DNA-specific transactivating factors (MacDonald, et al., 1995). Direct inhibition, or quenching, of activator function as seen with the Kruppel protein (Licht, et al., 1993). Direct GTF interaction such as Drl-DRAPI inhibition of the association of TBP with TFIIA and TFIIB (Inostroza, et al., 1992; Johnson, 1995; Meisteremst and Roeder, 1991). TATA TATA elements are the sites of binding for the TATA-binding protein (TBP). TBP-TATA association nucleates the assembly of the pre-initiation complex. In S. cerevisiae the TATA element is located within a window 30 to 120 bp 5’ of the site of transcriptional initiation. This is in contrast to most other eukaryotes where the TATA element is located, without exception, 25-30 bp 5’ of the transcriptional initiation site (Struhl, 1995). While many sequence variations of the TATA-element exist, “TATAAA” has been defined as the optimal sequence for the TATA element in yeast through saturation mutagenesis (Chen and Struhl, 1988; Singer, et al., 1990; Wobbe and Struhl, 1990). The importance of TBP-TATA element association is underscored by the discovery of altered specificity TBP molecules, TBP“, that bind to the TATA derivative TGTAAA (Strubin and Struhl, 1992). TGTAAA in the absence of TBP’“3 does not support efficient transcriptional activity. Promoters lacking canonical TATA elements (TATA-less promoters) have also been identified. The name “TATA-less” denotes weak TBP-DN A affinity rather than a qualitatively distinct promoter element. Thus while TATA-less promoters are still dependent on TBP binding, TBP association with the promoter is unlikely to be the rate- limiting step in PIC formation. TATA-less promoters likely depend on other factors, such as TAFns, to recognize the promoter and nucleate PIC formation. INR The Inr element is a transcriptional regulatory sequence that encompasses the transcriptional start site. The Inr was defined as the promoter element present in the terminal-deoxynucleotidetransferase gene, distinct from TATA, that can nucleate PIC assembly (Smale and Baltimore, 1989). The term ‘initiator’ was first used to describe the 60bp region spanning the transcription start site of the TATA containing sea urchin H2A gene (Grosschedl and Bimstiel, 1980). Deletion of this 60bp region resulted in transcription initiation 25bp downstream from the TATA element, but promoter strength was reduced 4-fold. However, the variable distance between the TATA element and the Inr in yeast implies the existence of specific Inr sequences that direct proper transcriptional initiation in addition to stimulating mRN A production. TATA elements apparently establish the window within which initiation can occur but Inr elements are defined by specific sequences within that window (Hahn, et al., 1985; Healy, etal., 1987; Li and Sherman, 1991; Nagawa and Fink, 1985; Rudolph and Hinnen, 1987). Mutational analysis revealed a loose, but consistent Inr sequence of Py Py A N T/A Py Py, where Py = pyrimidine and underline marks the first nucleotide of the transcript (J avahery, et al., 1994). Inr elements are core promoter elements able to nucleate PIC assembly and direct accurate transcription initiation in a TATA independent manner in higher eukaryotes (W eis and Reinberg, 1992). Proteins that bind INR elements in higher eukaryotes include CIF (Kaufmann, et al., 1996), YYl (Usheva and Shenk, 1994), E2F (Means, et al., 1992), TFII-I, and USF (Roy, et al., 1997; Roy, et al., 1993; Sadovsky, et al., 1995), as well as RNAPII (Aso, et al., 1994; Carcamo, etal., 1991). Whether Inr elements are capable of functioning as independent promoter elements in yeast is unclear. However, experiments focusing on the GAL80 promoter have provided some intriguing results. The GAL80 promoter contains both TATA and Inr elements, with these two elements directing initiation at distinct sites in response to different environmental stimuli. Thus two independent pathways lead to GAL80 transcription, one INR dependent and one TATA dependent (Sakurai, et al., 1994). A nuclear protein that binds to the GAL80 INR has also been detected (Sakurai, et al., 1994). Poly (dA-dT) Elements Homopolymeric dA-dT sequences are commonly found in yeast promoters and in several cases have been shown to be necessary for wild-type levels of transcription in viva (Struhl, 1985). Such sequences posses unique structural characteristics which include an unusually short helical repeat (10.0 bp instead of 10.6 bp typical of B-DNA) (Peck and Wang, 1981; Rhodes and King, 1981), a distinctively narrow minor groove (Alexeev, et al., 1987), and structural rigidity (Nelson, et al., 1987). Poly dA-dT sequences do not favor the assembly and stability of nucleosomes in vitro (Kunkel and Martinson, 1981; Prunell, 1982). Thus, poly dA-dT elements function as promoter elements based on their intrinsic structure (Iyer and Struhl, 1995). Poly dA-dT tracts are abundant within the S. cerevisiae genome and occur predominantly at unit nucleosomal length both upstream and downstream of open reading frames (Raghavan, et al., 1997). However, the idea that poly dA-dT elements are simply architectural elements is contradicted by the finding that a related oligonucleotide was able to squelch the positive effect of a poly dA-dT element had on transcription (Lue, et al., 1989). This result argues for a specific polydA-dT binding protein. Indeed, such a protein, datin, has been identified (Winter and Varshavsky, 1989). However, the exact role that datin plays in transcriptional activation is still unclear. PROMOTER CONTEXT The previously described promoter elements exist in the nucleus in the context of chromatin, a structure of varying complexity that plays a regulatory role in gene expression. Chromatin simply defined is a complex with a 2:1 mass ratio of protein to DNA. The structure of chromatin varies from a decondensed and unfolded state to a highly ordered and compacted state. It follows that in the decondensed state the DNA is more accessible to transcription factors and thus transcriptionally active whereas in the highly ordered and compacted state it is inaccessible to transcription factors and is thus transcriptionally silent. In its simplest form chromatin consists of 145-147bp of DNA coiled around a histone octamer that consists of two copies of each histone protein, H2A, H2B, H3, and H4. This octamer structure is also known as a nucleosome (Luger, et al., 1997). The nucleosome structure is repeated in yeast about every 165bp (Thomas and Furber, 1976) leaving short segments of linker DNA. The extended chromatin fiber is a 10nm diameter filament with an appearance that resembles “beads on a string”. In higher eukaryotes the linker histone H1 binds to the linker DNA and facilitates the folding of the 10nm filament into the 30nm chromatin fiber and stabilizes the compacted structure (Butler and Thiele, 1991; Finch and King, 1976; Renz, et al., 1977; Thoma, et al., 1979). No functional homolog of histone H1 exists in yeast, a finding that is not so surprising given the small amount of linker DNA between nucleosomes in yeast, and likewise there is a paucity of 30nm chromatin in yeast. Covalent modifications of the histones affect the packaging of nucleosomal DNA. For example, acetylation of the lysine residues at the N-terminal tails is believed to loosen the nucleosomal structure and expose the DNA sequence, while deacetylation of these residues is believed to tighten the nucleosome and make DNA less accessible (Hampsey, 1997 ; Pazin and Kadonaga, 1997). Recently many different chromatin remodeling complexes have been described - complexes that act at the level of nucleosome rearrangement. These complexes will be discussed in detail below. Mutations in HTAI-HTBI, one of two gene pairs encoding histones H2A and H2B have phenotypes similar to mutations in SPT4, SPT5, and SPT6, genes that encode proteins that affect chromatin structure. Since htaI-htbl mediated suppression occurs by altering chromatin structure (Hirschhom, et al., 1992) and Spt5, Spt6, and Spt7 have been implicated in transcriptional regulation (Winston and Carlson, 1992) a link between chromatin and transcriptional repression is established. Non-histone chromatin proteins that serve as both negative and positive regulators of gene expression exist in yeast. These proteins are often referred to as “architectural” proteins because of their structural role in bending or kinking DNA strands into conformations that either facilitate or hinder interactions of other proteins with DNA. The association of the nonhistone chromosomal proteins with DNA thus results in a higher order of chromatin structure that has both negative and positive effects on gene expression, depending on the locus. The recently described “CP” complex has been shown to be required for global gene activation and/or for chromatin-mediated repression (Brewster, et al., 1998). The CP complex in yeast is an abundant nuclear dimer consisting of Cdc68 and Pob3 proteins. A cdc68(Ts) impairs gene expression at the restrictive temperature as a result of an alteration in Cdc68 protein structure (Rowley, et al., 1991). Other cdc68 mutations result in derepression of SUC2 and GALl when the UAS sequences for SUC2 and GALl are deleted (Lycan, et al., 1994; Malone, et al., 1991; Prelich and Winston, 1993). The CP complex maintains the H0 gene in a transcriptionally inactive state in the absence of the Swi4-Swi6 transcriptional activator (Lycan, ct al., 1994). The high mobility group (HMG) proteins are another set of nonhistone chromosomal proteins. HMG] and HMG2 are related proteins that affect the assembly of nucleosomes and the organization of chromatin structure in vitro (Bonne Andrea, etal., 1984; Nightingale, et al., 1996). Stimulatory effects of HMGl/2 on transcription in eukaryotic cell extracts have been reported (Paull, et al., 1996; Shykind, et al., 1995; Singh and Dixon, 1990; Tremethick and Molloy, 1988). HMGl/2 do not, however, function directly as transcriptional activators (Landsman and Bustin, 1991). In contrast, other reports have shown that HMGl/2 proteins act as repressors of RNAPH mediated transcription in vitro (Ge and Roeder, 1994; Stelzer, et al., 1994). Other HMG proteins with functions less well defined included HMG14/ 17 and HMG I/Y. HMG14/17 are small, highly charged proteins shown to be modestly enriched in an actively transcribing gene relative to inactive genes (Postnikov, et al., 1991). HMG W are small, highly related proteins known to interact with the PRDII promoter element and to stimulate the simultaneous binding of NF-KB to this element (Thanos and Maniatis, 1992). HMG W has a similar affect on NF-KB binding to the nitric-oxide synthase promoter/enhancer (Perrella, et al., 1999). The condensation of chromosomes into heterochromatin and heterochromatin-like structures represents one extreme of the spectrum of chromatin structures. Heterchromatin is refractory to the binding and influence of transcriptional activator proteins. This was elegantly demonstrated by the loss of ADEZ gene expression as a result of local chromatin structures - an effect known as position effect variagation (Gottschling, et al., 1990). Heterochromatin regions appear to be condensed since they prevent access to DNA-altering enzymes, are late replicating in S phase, and are associated in foci that appear to be at the nuclear periphery (Gotta and Gasser, 1996). S. cerevisiae heterochromatin contains histone H4 that is uniquely hypoacetylated at lysines K5, K8, and K16 but not at K12 (references in (Lowell and Pillus, 1998)). In yeast such regions are found at the telomeres as well as at the silent (HM) mating and ribosomal DNA (rDNA) loci. The transcriptional repression witnessed at these loci is mediated by the SIRI-4 (silent information regulator) genes. Repression at the HM loci and telomeres also requires other factors including Raplp (reviewed in (Shore, 1994)), as well as histones H3 and H4 (reviewed in (Loo and Rine, 1995)). The SIR genes are variably required for silencing at different loci. rDNA silencing depends only on SIR2 (Smith and Boeke, 1997). In contrast, all four SIR genes contribute to silencing the HM loci, whereas 10 telomeres only require SIR2, SIR3, and SIR4 (Aparicio, et al., 1991; Haber and George, 1979; Klar, et al., 1979; Rine and Herskowitz, 1987; Rine, et al., 1979). RNA POLYMERASE 11 AND ASSOCIATED FACTORS RNA Polymerase II Yeast RNAPII is a complex enzyme comprising twelve polypeptides encoded by RPBI through RPBIZ (W oychik and Young, 1994). Extensive structural conservation exists among the RNAPII subunits from diverse eukaryotic organisms; six of the yeast RNAPII subunits can be functionally replaced by their human counterparts (McKune, et al., 1995). RNAPII is composed of common as well as class-specific subunits. Five subunits, pr5, pr6, pr8, pr10, and pr12 are essential components of all three eukaryotic polymerases. RNAPII subunits prl, pr2, pr3, and pr11 are homologous to RNA polymerase I and RNAPII] subunits. Only pr4, pr7, and pr9 are unique to RNAPH. With the exception of RPB4 and RPB9, all of the genes that encode the yeast RNAPH subunits are essential. Yeast genetics has provided some insight into the in viva functions of some of the individual RNAPH subunits. RPBI and RPBZ mutations have been isolated that influence the accuracy of transcriptional initiation, implicating these two subunits in start site selection (Amdt, et al., 1989; Berroteran, et al., 1994; Hekmatpanah and Young, 1991). 6-azauracil (6-AU) sensitivity is a phenotype of other RPBI and RPBZ mutations, a phenotype associated with transcriptional elongation defects. This finding suggests that both pr1 and pr2 play a role in transcriptional elongation (Archambault, etal., 1992; ll Powell and Reines, 1996). Sequence similarity of RPBI and RPBZ to Band B’ of the prokaryotic RNA polymerase also indicates that RPBI and RPBZ likely encode the catalytic subunits of RNAPII (Markovtsov, et al., 1996; Mustaev, et al., 1997; Severinov, et al., 1996; Zaychikov, et al., 1996). pr3 is involved in assembly of the RNAPII complex (Kolodziej and Young, 1989; Kolodziej, et al., 1990; Kolodziej and Young, 1991). pr9 has a zinc finger motif, mutations in which affect start site selection (Furter Graves, etal., 1994; Gadbois, etal., 1997; Hull, et al., 1995; Sun and Hampsey, 1996). An rpb9 allele that suppresses a TFIIB mutation that affects start site selection has been identified (Sun, et al., 1996). This finding leads to the supposition that pr9 affects start site selection in concert with TFIIB. A form of yeast RNAPII lacking the pr4 and pr7 subunits has been identified. Indeed biochemical analysis revealed that pr4 and pr7 are loosely associated with RNAPH. While yeast RNAPII lacking pr4 and pr7 is capable of elongation in vitro it is deficient in accurate initiation in vitro (Edwards, et al., 1991). pr4 has been implicated as playing a role in tolerance of RNAPII to stress. rpb4 mutants display substantially impaired growth rates at elevated temperature or under conditions of nutritional deprivation (Choder and Young, 1993). The largest subunit of the RNAPII complex, pr1, has a unique tandemly repeated heptapeptide sequence at its carboxy-terminus. This carboxy-terminal domain (CTD) is highly conserved among species and has a consensus sequence Tyr-Ser-Pro- Thr-Ser-Pro-Ser. The number of repeats appears to increase with the complexity of the organism, ranging from 26-27 in yeast to 52 in human cells (Young, 1991). The CTD is essential for viability, yet its function is not yet clear. Yeast that contains RNAPH lacking all but 8 to 10 repeats is viable (Nonet and Young, 1989; West and Corden, l2 1995). However, strains with only 8 to 10 CTD repeats display a cold-sensitive growth defect, a phenotype that was exploited to isolate extragenic suppressors of CTD truncations and resulted in the discovery of the SRB genes (Nonet and Young, 1989). RNAPII with a truncated CTD is also unable to initiate transcription at promoters that lack a TATA element (Akoulitchev, et al., 1995; Buermeyer, et al., 1995). Two forms of RNAPH exist in vivo, RNA polymerase 110, which is hyperphosporylated at the CT D, and RNA polymerase IIA, which is not phosphorylated. RNA polymerase IIA preferentially enters into the PIC, whereas RNA polymerase 110 is found in the elongating complex (Dahmus, 1996). Conversion of RNA polymerase 11A to 110 correlates with the transition from initiation to elongation (Lu, et al., 1991; O’Brien, et al., 1994). These results implicate the CTD as having a central role in the conversion of RNAPII from a molecule intent on promoter recognition to one that is elongation competent. CTD kinase candidates are numerous and include TFIIH (Feaver, et al., 1991; Lu, et al., 1992; Serizawa, et al., 1992), P-TEFb (Marshall, et al., 1996), SrblO/ll (Liao, et al., 1995), Cdc2 (Cisek and Corden, 1989), and Ctkl (Lee and Greenleaf, 1991). Whether these CTD kinases are gene specific or if they affect different steps during the transition from promoter recognition to elongation is still unclear. Just as RNAPII is converted from the HA form to the 110 form by a kinase it is recycled back to the HA by a phosphatase (Chambers and Dahmus, 1994). CTD phosphatase activity is regulated by TFIIB and TFIIF (Chambers, et al., 1995). It has been proposed that since RNA polymerase IIA preferentially enters the PIC that TFIIB and TFIIF are responsible for RNAPH recycling (Lei, et al., 1998). General Transcription Factors l3 As mentioned previously the general transcription factors (GTFs) include TFIIA, TFHB, TFIID, TFIIB, TFIIF, and TFIIH. The GTFs were identified in order-of-addition experiments as the protein fractions necessary for accurate transcription initiation by RNAPH from double stranded templates in vitro (Sayre, et al., 1992). These factors plus RNAPII comprise the PIC. TFIID TFIID is a multisubunit complex that consists of TATA binding protein (TBP) and TBP-associated factors (TAFus) (reviewed in (Burley and Roeder, 1996; Hernandez, 1993). In vitro results indicate that, depending on the promoter, either TBP or TFIID is involved in promoter recognition and thus serves to nucleate PIC formation. At promoters that contain a TATA element TBP is sufficient whereas TAFns are required at TATA-less promoters (Aso, et al., 1994; Kaufmann and Smale, 1994; Martinez, et al., 1994; Smale, et al., 1990; Verrijzer, et al., 1995). Immunodepletion of yTFIID by TAP-specific antibodies blocks in vitro transcription by RNAPII and not RNAPI or RNAPIII, this indicates that TFIID is a RNAPII specific complex (Klebanow, et al., 1997). TBP is a universal transcription factor, required for initiation by all three eukaryotic RNA polymerases (Hernandez, 1993). TBP is a subunit of the 750kDa TFIID complex composed of TBP and TBP-associated factors (T AFs). Whereas TBP functions in basal level transcription, TFIID is required for response to transcriptional activators in metazoan in vitro transcription systems (Burley and Roeder, 1996; Pugh and Tjian, 1992; Verrijzer and Tjian, 1996). 14 Yeast TBP is a monomer of 27 kDa and can be divided into two structural domains, the less well conserved N-terminal domain and the highly conserved C-terminal domain. The conserved C-terminal domain of TBP consists of two direct repeats, suggesting that TBP may have been the result of a gene fusion event of two monomers. The C-terminal region of TBP is both necessary and sufficient for TATA binding as well as transcription initiation (Zhou, et al., 1993). The crystal structure of the C-terminal TBP domain from yeast has been solved (Kim, et al., 1993) and reveals a “molecular saddle” type structure made up of two roughly symmetrical halves. The direct repeats of the TBP molecule display the expected twofold symmetry. Each repeat folds into a twisted beta-sheet, topped by a long C-terminal alpha helix that is available for interaction with other factors. TBP has a convex upper surface as well as a concave lower surface capable of DNA minor groove binding via a curved eight-stranded antiparallel beta-sheet on the inner surface of the saddle. TBP-DNA interaction results in a kink in the DNA that bends the DNA about 80° toward the major groove. The DNA bending brings proteins bound on either side of TBP, such as TFHA and TFIIB, into close physical proximity. The notion that GTF function is conserved among species is bolstered by the fact that yeast TBP is functionally interchangeable with mammalian TBP in in vitro transcription experiments (Buratowski, et al., 1988; Cavallini, et al., 1989; Hahn, et al., 1989). The yeast TBP gene is known as SPTIS. SPT15 was identified in a genetic selection for suppressors of a Ty insertion in the HIS4 promoter (his4-9I 7d) (Eisenmann, et al., 1989). In the his4-9l 75 mutant, transcription initiates within the Ty (Selement resulting in non- functional transcripts. Mutations within SPT15 suppressed his4-9175 by shifting initiation from within the 5element to the his4 promoter, resulting in a His+ phenotype. SPT15 is an 15 essential gene and spt15 mutants have pleiotropic effects, further demonstrating the essential nature of this transcription factor (Eisenmann, et al., 1989). TAFIIS Two observations lead to the speculation that yeast TBP existed in a complex with other proteins. First, mutations in several SPT genes resulted in a similar set of pleiotropic phenotypes including SPTI5 (TBP). Secondly, the SPT3 gene product was shown to physically interact with TBP (Edwards, et al., 1991). Indeed, 12 TAF subunits have now been identified in yeast ranging in size from 17 to 155 kDa (Klebanow, et al., 1997; Klebanow, et al., 1996; Moqtaderi, et al., 1996; Poon, et al., 1995; Reese, 1994; Walker, et al., 1996). All of the genes encoding the yTAFus are essential except the gene that encodes yTAFn30. Furthermore, significant similarity exists between TAFus from various species. The only metazoan TAFu that does not have a homolog in yeast is TAFn110, which is required for activation by the glutamine-rich activator Spl. It is interesting to note that glutarnine-rich activators are not known exist in yeast and that metazoan glutamine-rich activators function poorly, if at all, in yeast. The lack of a TAFu110 homolog may be the reason that glutamine-rich activators do not exist in yeast (Moqtaderi, et al., 1996). Yeast TAFnl45, also known as y TAFnl30, binds to TBP and is believed to serve as a scaffold for the assembly of the other TAFns into TFIID. Biochemical and crystallographic analysis has indicated that a histone octamer-like structure may exist within TFIID. TAFns that possess domain structures reminiscent of those found in histone H3-, H4-, and H2B have been identified in human, Drosophila, and yeast TFIID complexes (Burley and Roeder, 1996; Nakatani, et al., 1996; Xie, et al., 1996). In the case of yeast TFIID, a putative yTAFn histone octamer-like structure would contain a heterodimer of yTAFnl7 (H3-1ike) and 16 yTAFn60 (H4-like) sandwiched between two yTAFn61 (H2B-like) homodimers. No H2A- like TAFn has been identified in any species. This has lead to the proposal that this octamer like structure may wrap DNA on the surface of TFIID. However, the recent nucleosome crystal structure indicates that arginine side-chains in histones that interact with DNA are not conserved in histone-like TAFus (Luger, et al., 1997). Thus if the histone-like TAFus do bind DNA, they likely do so differently than the histones that comprise a nucleosome. The essential nature of the yTAFn genes clearly points to a critical role for them in viva. However, their exact role remains undefined. Yeast TAFus were demonstrated to function as co-activators of transcription in vitro (Sauer, et al., 1995; Sauer, et al., 1996; Verrijzer and Tjian, 1996) which lead to the presumption that this was a general phenomenon. However, this presumption was undermined when depletion or inactivation of several yTAFns, including yTAFul45/yTAFnl30, did not compromise transcriptional activation in viva (Moqtaderi, et al., 1996; Walker, et al., 1996). In the case where the yTAFus were required for response to activators, the requirement mapped to the core promoter element instead of the elements bound by activators — the promoter lacked a TATA-element and thus the requirement for the yTAFns (Moqtaderi, et al., 1996). This result does however implicate yTAFnl45/ yTAFul 30 as a promoter selectivity factor. Indeed, yTAFul45/ yTAFul30 mediates the process of activation at promoters lacking a TATA element (Moqtaderi, et al., 1996). These conclusions were later tempered when it was demonstrated that the histone-like yTAFns were indeed required for the expression of a subset of yeast genes (Apone, et al., 1998; Michel, et al., 1998; Moqtaderi, et al., 1998). It is possible that yTAFus are activator cofactors, but only for a subset of genes - specifically for genes involved in cell cycle control. A taf90(Ts) mutant fails to progress 17 through the Gle phase of the cell cycle at the restrictive temperature (Apone, et al., 1996). Another example is a tafl45(Ts) mutation that blocks transcription of G1/S cyclin genes at the restrictive temperature (Walker, et al., 1997). TFIIB Yeast TFIIB is a monomer of 38kDa encoded by the S UA7 gene. S UA7 was identified in a genetic screen where suppressors of a translational defect at the CYCI locus were sought. A recessive mutation at the sua7 locus that shifted the transcription start site downstream of the normal start site such that the translational impediment was eliminated from the cycI transcript was isolated (Pinto, et al., 1992). This result implicates TFIIB as being involved in start site selection in yeast (Li, et al., 1994; Pinto, et al., 1992). TFIIB enters the PIC after TBP binding to the TATA element and recruits RNAPII. TFIIB also interacts with other GTFs, including the TFIIF subunits RAP30 and RAP74 (Buratowski, et al., 1989). However, TFIIB associated with RAP74 is unable to associate with RAP30, perhaps because of some dynamic interaction between TFIIF and TFIIB during PIC assembly (Fang and Burton, 1996). TFIIB has been implicated as the direct target of many gene-specific activators, leading to the proposal that certain activators stimulate transcription by TFIIB recruitment (Lin and Green, 1991; Roberts, et al., 1993). TFIIB has been shown to interact with TAFn40 as well (Goodrich, et al., 1993). TFIIB consists of two domains, an N-terminal cysteine rich region that includes a zinc binding motif and a C—terminal protease resistant region that contains two imperfect 75 amino acid repeats (Ha, et al., 1991; Malik, et al., 1991; Pinto, et al., 1992). The C-terminal 18 region (TFIIBc) is capable of binding TBP but not RNAPII or TFIIF and is thus unable to support transcription (Barberis, et al., 1993; Buratowski and Zhou, 1993; Ha, et al., 1991; Yamashita, et al., 1993). In the absence of transcriptional activators the N- and C-terminal domains of TFIIB are involved in an intramolecular interaction that preclude interaction with other PIC members. In the presence of a transcriptional activator TFIIB undergoes a conformational change that allows for PIC assembly (Roberts, 1994). This has lead to the proposal that TFIIB is a target of transcriptional activators. The three dimensional structure of a TFIIBc-TBP-DNA complex has been solved (Nikolov, et al., 1995). In this structure TFIIB binds directly to TBP, and also contacts DNA immediately upstream and downstream of the TATA element. These TFIleDNA contacts are made possible by the bend put into the DNA as a result of TBP binding. The binding of TFIIB thus stabilizes the TBP-TATA interaction. Like TBP, TFIIB recognizes specific sequences found in certain promoters (Lagrange, et al., 1998). The cysteine rich N-terminal zinc binding domain is absent in the crystal structure, but it is thought to bind DNA in the vicinity of the transcription start site and to stabilize melting of the promoter. TFIIF Yeast TFIIF is composed of three subunits with apparent molecular masses of 105, 54, and 30 kDa encoded by the genes TFGI, TFGZ, and TF G3 respectively (Henry, et al., 1992). TF GI and TFGZ are both about 50% similar to RAP30 and RAP74, the two subunits that comprise human TFIIF, respectively (Henry, et al., 1994). TFIIF binds tightly to RNAPII and suppresses nonspecific binding of RNAPH to DNA (Conaway and Conaway, l9 1993). Stabilization of the PIC is another function of TFHF (Conaway and Conaway, 1993; Greenblatt, 1991). Photo-crosslinking studies have shown that the two largest subunits of TFIIF bind promoter DNA between the TATA element and start site (Coulombe, et al., 1994; Kim, et al., 1997; Robert, et al., 1996). The largest subunit of TFIIF also binds DNA upstream of the TATA element and induces a conformational change that affects the position of RNAPII relative to the DNA template (Forget, et al., 1997). TFIIF also plays a role in transcriptional elongation by suppressing the transient pausing of RNAPH (Bengal, et al., 1991; Bradsher, et al., 1993; Flores, et al., 1989; Izban and Luse, 1992; Price, et al., 1989). TF GI is the only gene of any of the TFIIF components to be identified in a genetic selection for transcription factors. SSU 71 was identified as a suppressor of a TFIIB defect that resulted in spurious start site selection (Sun and Hampsey, 1995). Sequence analysis revealed that Ssu71 is identical to ng1 (Henry, et al., 1994; Sun and Hampsey, 1995). Thus, TFIIB functionally interacts with TFIIF. TFIIE Yeast TFIIE is composed of a 56 kDa subunit (TFIIB-0t) and a 34 kDa subunit (TFIIEB). The OLB heterodimer structure of the yeast TFIIB is in contrast with the 012152 heterotetramer form of TFIIE found in human cells (Leuther, et al., 1996). Order of addition experiments show that TFIIE enters the PIC after RNAPII and TFIIF and that TFIIB recruits TFIIH (Buratowski, et al., 1989; Flores, et al., 1992). TFIIB interacts with the unphosphorylated form of RNAPH, with both subunits of TFIIF, and with TFIIH (Flores, et al., 1989; Maxon and Tjian, 1994). The functional importance of the TFIIB-TFIIH 20 interaction was demonstrated by the inability of S. cerevisiae TFIIB to functionally replace the S. pambe TFIIE unless TFIIH from S. cerevisiae replaced its counterpart from S. pambe as well (Li, et al., 1994). TFIIB influences TFIIH in the following ways: TFIIE recruits TFIIH to the PIC, TFIIB stimulates TFIIH dependent phosphorylation of RNAPII, and TFIIB stimulates TFIIH-dependent ATP hydrolysis (Lu, et al., 1992; Ohkuma, et al., 1995; Ohkuma and Roeder, 1994). Another possible role for TFIIE is in the formation or stabilization of the melted promoter region based on the fact yTFIIE binds single-stranded DNA (Kuldell and Buratowski, 1997). TFHE has also been implicated as the direct target of some transcriptional activators (Sauer, et al., 1995; Zhu and Kuziora, 1996). TFIIH Upon TFIIH entry the formation of the PIC is complete (Flores, et al., 1992). Yeast TFIIH is a nine subunit complex with a total mass of approximately 500 kDa. TFIIH is the only GTF that has known enzymatic properties, which include DNA-dependent ATPase activity (Conaway and Conaway, 1989; Roy, et al., 1994), ATP-dependent DNA helicase activity (Schaeffer, et al., 1993; Serizawa, et al., 1993), and CTD kinase activity (Feaver, et al., 1991; Lu, et al., 1992; Serizawa, et al., 1992). TFIIH plays a central role in forming an open transcription complex. The TFIIH component Rad25, an ATP-dependent DNA helicase, is believed to be necessary to form the open promoter complex (J iang, et al., 1994; Wang, et al., 1992). If the DNA template is premelted or supercoiled, a structure that inherently favors unwinding, the requirement for TFIIH, TFIIB, and ATP can be bypassed (Holstege, et al., 1995; Pan and Greenblatt, 1994; Parvin and Sharp, 1993; Parvin, et al., 1994; Tantin and Carey, 1994; Tyree, et al., 1993). 21 The CTD phosphorylation function of TFIIH is implicated in the transition of the initiation competent form of RNAPII to the elongation competent form of RNAPII. Presumably, phosphorylation of the CTD causes a conformational change in the PIC, disrupting the interaction between CTD and TBP thus leading to promoter clearance (Dahmus, 1994; O’Brien, et al., 1994). TFIIH also promotes the transition from very early elongation complexes to stable elongation complexes (Dvir, et al., 1997). TFIIA Yeast TFIIA is composed of two polypeptides with apparent molecular masses of 32 and 13.5 kDa encoded by the essential genes TOAI and TOA2 respectively (Radhakrishnan, et al., 1997; Ranish and Hahn, 1991; Ranish, et al., 1992). TFIIA associates with the PIC through interaction with TBP and stabilizes TBP-TATA element binding (Buratowski, et al., 1989; Irnbalzano, et al., 1994). The fact that TFIIA is dispensable for accurate initiation in a purified system direct by TBP calls its classification as a GTF into question (Cortes, et al., 1992; DeJong, et al., 1995; Hansen and Tjian, 1995; Ozer, et al., 1994; Sayre, et al., 1992; Sun, et al., 1994; Yokomori, etal., 1993). TFIIA does, however, counteract negative factors that associate with the TFHD complex such as Drl-DRAPl/NCZ, PC3/Dr2, HMG], and Motl , and thus TFIIA is necessary for TFIID directed transcription (Auble, 1994; Ge and Roeder, 1994; Inostroza, et al., 1992; Meisteremst, et al., 1991; Merino, et al., 1993). Co-overexpression of TOAI and TOA2 suppresses the phenotypes of TBP mutants identified as having reduced TATA element affinity. This finding suggests that TFIIA is important for transcriptional activation, and the degree to which different promoters require 22 TFIIA may vary (K. Amdt, personal communication). Yeast TFIIA is known to interact with specific transcriptional activators such as Zta (Ozer, et al., 1994; Yokomori, et al., 1994), the coactivators PC4 and HMG2 (Ge, et al., 1994; Shykind, et al., 1995), and VP16 (Kobayashi, etaL,1998) Holoenzymes The fact that several GTFs were known to associate with RNAPII in the absence of DNA implied the existence of a holoenzyme complex (Conaway and Conaway, 1993). Indeed such complexes have been identified in yeast and in mammalian cells. Essentially the same yeast holoenzyme was independently identified in both the Young and Kornberg labs. Using antibodies directed against the SRB proteins, Young and coworkers isolated and characterized a version of the holoenzyme (Koleske and Young, 1994). Simultaneously Kornberg and coworkers identified a high molecular wieght complex associated with the CTD that they called “mediator” or MED (Kim, et al., 1994). The MED complex contains the SRB proteins (reviewed (Bjorklund and Kim, 1996)). The yeast RNAPII holoenzyme includes all twelve pr molecules, SRB/MED proteins, TFIIB, TFIIF, and TFIIB (Koleske and Young, 1994). A distinct form of yeast holoenzyme was discovered using a strategy based on an RNAPII affinity column immobilized through the CTD (Shi, et al., 1997). Proteins common to both versions of holoenzyme include TFIIB, TFIIS, TFIIF, and Gall 1, but not SRB/MED proteins (Wade, et al., 1996). Novel holoenzyme proteins include Paf 1, Cdc73, Ccr4, and Hprl (Chang and Jaehning, 1997; Shi, et al., 1997; Wade, et al., 1996). The second holoenzyme affects the expression of a different spectrum of genes than the SRB/MED containing holoenzyme and is thus unique (Shi, et al., 1996). 23 Quantitation by comparative Western blotting of RNAPII and SRB proteins has determined that only a portion of RNAPH is present in SRB/MED holoenzyme complexes (Koleske and Young, 1994). The percentage of total RNAPH present in the Pafl-Cdc73- Ccr4-Hpr1 holoenzyme has yet to be determined, although crude analysis indicates that this holoenzyme is less abundant than the SRB/MED holoenzyme (Hampsey, 1998). The identification of holoenzyme has challenged the assumption that RNAPII and the GTFs assemble on the promoter in a stepwise fashion. It is possible that a subset of RNAPII is present in a preassembled complex lacking only TFIID and TFIIA. Thus assembly of the PIC at a promoter may occur in only a two step process with TFIII) promoter recognition followed by holoenzyme association with THE). CHROMATIN REMODELING COMPLEXES SWIISNF The SWI genes were originally identified in a genetic screen for yeast mutants defective in mating type switching (Stern, et al., 1984), whereas the SNF genes were identified based on an inability to express the S UCZ gene and therefore unable to ferment sucrose (Neigeborn and Carlson, 1984). SNFZ and SW12 were later found to be identical, and thus the connection between the two systems was made. Suppressors of snf and swi mutations were found in the ssn and sin genes, respectively. SSN20 and SIN2 turned out to be identical to SPT6 (Clark Adams and Winston, 1987; Neigebom, et al., 1987) and HHTI (Kruger, et al., 1995), respectively. As previously described, SPT6 is a member of a class of SPT genes that either encode histones or are linked to chromatin function. HHTI encodes 24 histone H3. Thus the link between the S WI/SNF gene products and chromatin was established. The mOdel that evolved maintains that the SPT/SIN gene products are involved in the establishment of inactive chromatin whereas the SWIISNF genes are involved in overcoming the repressive effects of chromatin (Winston and Carlson, 1992). Indeed, this model received support from the finding that the chromatin structure at the SUC2 promoter is altered in swi2/snf2 and snf5 mutants, and this chromatin defect is suppressed by A(htal- htbl) (Hirschhom, et al., 1992). The SWI/SNF complex in yeast is an ll-subunit complex (Cairns, et al., 1994; Cote, et al., 1994). SWIISNF subunits include Swil, SwiZ/Snf2, Swi3, Snf5, Snf6, Snfl l, SWp29, Swp56, SWp61, Swp73, and Swp82 (Burns and Peterson, 1997). The only enzymatic activity known to reside in the complex is the DNA-dependent ATPase activity of Swi2/Snf2. A functional link to the PIC is made through SWp29, which is identical to both the ng3 subunit of TFIIF and the TAFn30 subunit of TFIID (Cairns, et al., 1996). SWIISNF was reported to be a component of the SRB/mediator holoenzyme, but some methods of purifying holoenzyme do not yield SWIISNF (Cairns, et al., 1996; Li, et al., 1996; Myers, et al., 1998; Wilson, et al., 1996). The SWIISNF complex binds a DNA structure that mimics the structure of DNA as it enters and exits the nucleosome (Quinn, et al., 1996). This property raises the possibility that SWI/SNF binds specifically to nucleosomal DNA and subsequently is able to disrupt nucleosome structure by modulating the DNA structure and topology. The SWI/SNF complex is only required for the expression of 6% of yeast genes when tested in glucose rich medium, including [-10, SUC2, Ty, ADHI, ADHZ, 1N0], STA] as determined by the relative abundance of mRNAs in yeast strains with and without Swi2/Snf2 (Holstege, et al., 1998). 25 RSC The RSC (remodels the structure of chromatin) complex was isolated from yeast based on homology to components of the SWIISNF complex (Cairns, et al., 1996). RSC is a 15-subunit complex that includes several SWIISNF related subunits as well as a DNA- dependent ATPase (Cairns, et al., 1996; Cao, et al., 1997). However, RSC is approximately 10-fold more abundant that SWIISNF and unlike any genes encoding SWIISNF subunits certain genes encoding RSC subunits are essential. No current evidence links RSC to transcription. RSC does not play a role in either SUC2 or GALIO expression nor have any histone mutants been found that suppress RSC defects (Hampsey, 1998). RSC mutants cause cell cycle arrest in the Gle phase of the cell cycle (Cao, et al., 1997). It thus appears that SWIISNF and RSC are functionally distinct. I-Iistone Acetyltransferases When the transcriptional activation domain (AD) of VP16 was fused to the DNA binding domain of Gal4 and overexpressed in viva in yeast, cell growth was inhibited (Berger, et al., 1992; Gill and Ptashne, 1988). Isolation of yeast mutants that relieved this toxicity led to the identification of ADA I, ADA2, ADA3, ADAS/SPTZO, and GCN5, genes that encode putative adaptor proteins (Berger, et al., 1992; Horiuchi, et al., 1997; Marcus, et al., 1996; Marcus, et al., 1994; Pina, etal., 1993; Roberts and Winston, 1996). ADA2, ADA3, and GCN5 were necessary for maximal activation by VP16 and Gcn4 (Berger, et al., 1992; Georgakopoulos, et al., 1995; Georgakopoulos and Thireos, 1992; Pina, et al., 1993). Ada2 binds to the activation domains of both of these activators as well (Barlev, et al., 1995; 26 Silverrnan, et al., 1994). The discovery that GCN5 encoded a histone acetyltransferase (HAT) allowed a specific function to be assigned to this protein and provided a direct link between histone acetylation and transcriptional activation (Brownell and Allis, 1996; Brownell, et al., 1996) - a link that had been presumed for 30 years (Allfrey, et al., 1964). Acetylation of lysine residues at the N-terminal tails of histones neutralizes their positive charge and presumably reduces the interaction of the core histone tails with DNA, which is postulated to result in the activation of gene expression. However, it remained troubling that while Gcn5 was able to acetylate free histones it was unable to use histones assembled into nucleosomes as a substrate (Kuo, et al., 1996; Yang, et al., 1996). Recently, the issue has been settled as a result of the identification of a complex that includes Gcn5 and is able to use histones assembled into nucleosomes as substrates. The 1.8 MDa SAGA complex (Spt-Ada- Gcn5-Acetyltransferase) contains Gcn5, Adal , Ada2, Ada3, Spt3, Spt7, Spt8, Spt20, Tral , and several yTAFu subunits (Grant, et al., 1997; Grant, et al., 1998). Other nucleosomal HAT complexes have been identified, some that contain Gcn5 and are distinct from SAGA (RuizGarcia, et al., 1997; Saleh, et al., 1997). It was also determined that under optimal solution conditions GCN5 alone is indeed able to acetylate nucleosomes (Tse, et al., 1998). HAP fl3l4l5 The Hap2/3/4/5 protein complex positively regulates many of the genes involved in the TCA cycle and oxidative phosphorylation in the budding yeast Saccharamyces cerevisiae (Forsburg and Guarente, 1989; Hahn and Guarente, 1988; Pinkham and Guarente, 1985; Pinkham, et al., 1987). The Hap2/3/5 polypeptides form the DNA binding component of the complex (McNabb, et al., 1995 ; Olesen, et al., 1987; Xing, et 27 al., 1993) specific for the CCAAT-related cis element of UASZ found upstream of the CYCI gene and other Hap2/3/4/5 regulated genes. Hap2/3/5 are quite similar to proteins known to contain a histone fold and likely employ this structure to associate with DNA (Coustry, etal., 1996). Hap4 is responsible for the transcriptional activation capability of the Hap2/3/4/5 complex (Forsburg and Guarente, 1989). HAP4 expression is repressed in the presence of glucose, presumably by Migl which is known to bind upstream of HAP4 (Ronne, 1995), and stimulated by an undefined factor (D. McNabb personal communication) in the presence of non-fermentable carbon sources (Guarente, et al., 1984). Hap4 is not required for the DNA-binding activity of the Hap2/3/4/5 complex (Olesen and Guarente, 1990; Xing, et al., 1993). In this respect Hap4 is analogous to the VP16 trans-activator of herpes simplex virus, which also contains a potent transcriptional activation domain but associates with DNA cis regulatory elements only through a complex containing two cellular proteins, Oct-l (Abate, et al., 1990; apRhys, et al., 1989; Kristie and Roizman, 1988; O’Hare and Goding, 1988; Preston, et al., 1988) and HCF (Wilson, et al., 1993). In some respects, the mechanisms employed for transcriptional activation by VP16 and Hap4 appear to be fundamentally different. Whereas mutations in ADA2, ADA3, or GCN5 significantly reduce activation by VP16 and Gcn4, they have a much smaller effect on activation by Hap4 (Berger, et al., 1992; Pina, et al., 1993) (personal observation). Furthermore, Wang et a1. (Wang, et al., 1995) demonstrated that VP16 and HAP4 have different co-factor requirements. However, they were unable to clone a HAP4 specific adaptor. 28 CBF-A, CBF-B, and CBF-C appear to be the human homologues of Hap2, Hap3, and Hap5 respectively. CBF-A/B/C are the subunits of CBF (also known as NF-Y and CH) the heterotrimeric ubiquitous DNA binding protein that binds to CCAAT motifs in promoters of numerous eukaryotic genes (Chodosh, et al., 1988; Dom, et al., 1987; Maire, et al., 1989; Maity, et al., 1988). In the case of CBF, CBF-B and CBF-C provide the trans-activation potential (Coustry, et al., 1996), and no human homolg of HAP4 has been identified. Recently, a functional homologue of Hap4 was identified in the respiratory yeast Kluyveramyces lactis (Bourgarel and Nguyen, 1999) that exhibits two small domains (11 and 16 amino acids respectively) highly homologous to corresponding domains of Saccharamyces cerevisiae Hap4, while overall similarity is rather weak. The K. lactis homolog of Hap4 was identified by heterologous complementation of a Ahap4 Saccharamyces cerevisiae mutant. This is the first example of a Hap4 homologue in any organism other than Saccharamyces cerevisiae. Hap2 and Hap3 homologues have also been identified in K. lactis (Mulder, et al., 1994; Nguyen, et al., 1995). 29 W Identification and Characterization of the Hag Activation Domainj s) INTRODUCTION The Hap2/3/4/5 protein complex positively regulates many of the genes involved in the TCA cycle and oxidative phosphorylation in the budding yeast Saccharamyces cerevisiae (Forsburg and Guarente, 1989; Hahn and Guarente, 1988; Pinkham and Guarente, 1985; Pinkham, et al., 1987). Hap4 is responsible for the transcriptional activation capability of the Hap2/3/4/5 complex (Forsburg and Guarente, 1989), and is not required for the DNA-binding activity (McNabb, et al., 1995; Olesen, et al., 1987; Xing, et al., 1993). In this respect, Hap4 is analogous to the VP16 trans-activator of herpes simplex virus, which also bears a potent transcriptional activation domain but associates with DNA cis regulatory elements only through a complex containing two cellular proteins, Oct-1 (apRhys, et al., 1989; Kristie and Roizman, 1988; O’Hare and Goding, 1988; Preston, et al., 1988) and HCF (Wilson, et al., 1993). Hap2/3/5 are ’ constitutively expressed whereas Hap4 expression is strongly induced when any non- ferrnentable sugar is the sole carbon source (Forsburg and Guarente, 1989; Hahn and Guarente, 1988; McNabb, et al., 1995; Pinkham and Guarente, 1985). Therefore, the activity of this complex transcriptional activator is regulated by production of the activation subunit. Transcriptional activation proteins are typically bipartite in structure, with separate domains for DNA binding and trans-activation (Mitchell and Tjian, 1989; 30 Triezenberg, 1995). Activation domains have traditionally been classified according to amino acid composition. Thus activators have been classified as acidic, glutamine-rich, proline-rich, serine/threonine-rich, isolucine-rich, or basic (Attardi and Tjian, 1993; Estruch, et al., 1994; Johnson, et al., 1993). However, mutational analyses of Gal4 (Leuther, et al., 1993), RelA (Blair, et al., 1994), p53 (Lin, et al., 1994), VP16 (Cress and Triezenberg, 1991; Regier, et al., 1993; Sullivan, etal., 1998), Gcn4 (Drysdale, et al., 1995), the glucocorticoid receptor (Almlof, et al., 1997; Iniguez-Lluhi, et al., 1997), C 1 (Sainz, et al., 1997), NRF—l and NRF—2 (Gugneja, et al., 1996), and Spl (Gill, et al., 1994) have demonstrated that specific patterns of bulky hydrophobic and aromatic amino acids are more critical for activation domain function than the residues that are most abundant. As shown in the cases of VP16 (Triezenberg, et al., 1988), p53 (Horikoshi, et al., 1995) and Gcn4 (Drysdale, et al., 1995), activation domains can often be further divided into regions each capable of activating transcription independently and all of which are dependent on hydrophobic residues. The VP16 activation domain can be divided into small definable regions, VP16 N (aa410-546) and VP16 C (33456-490), each of which relies on a central cluster of hydrophobic residues (Cress and Triezenberg, 1991; Regier, et al., 1993; Sullivan, et al., 1998). In contrast, Gcn4 relies on multiple overlapping regions containing seven different clusters of hydrophobic residues (Jackson, et al., 1996). Collectively, these results suggest that activation domains are composite in nature, containing multiple hydrophobic motifs. One reason that transcriptional activators often possess multiple activation domains may be that different segments of the activation domain interact with different 31 proteins in the PIC. The ability to interact with numerous targets is required because transcriptional activators act at a variety of promoters that may have different rate limiting steps. For instance, VP16 is known to interact in vitro with TBP (Ingles, et al., 1991; Stringer, et al., 1990), TFIIB (Labow, et al., 1990; Lin and Green, 1991; Lin, et al., 1991; Roberts, et al., 1993), TFIIA (Kobayashi, et al., 1998), TFIIH (Xiao, et al., 1994), and TAFn40 (Goodrich, et al., 1993). Transcriptional activators have evolved able to interact with numerous targets and therefore able to act at a variety of promoters no matter what the local constraints to activation. Despite the similar requirement for hydrophobic residues, no consensus motif characteristic of activation domains has been defined. This has lead to the proposal that the critical structural features for activation domains are somewhat flexible. The activation domains of VP16, Gal4, Gcn4, RelA, and glucocorticoid receptor have been demonstrated to be unstructured in aqueous solution under neutral pH (Dahlman-Wright, et al., 1995; Donaldson and Capone, 1992; O’Hare and Williams, 1992; Schmitz, et al., 1994; Van Hoy, et al., 1993). However, activation domains can adopt specific structures in different solvents (Dahlman-Wright, et al., 1995; Donaldson and Capone, 1992; O’Hare and Williams, 1992; Schmitz, et al., 1994; Van Hoy, et al., 1993). This lead to the speculation that activation domains are able to adopt higher-order structure upon contacting their target molecules by an “induced fit” mechanism. Indeed, fluorescence studies later demonstrated the existence of an induced conformation for the VP16 activation domain upon interaction with TBP, and to a lesser extent, TFIIB (Shen, et al., 1996). Other examples of induced fits include p53/MDM2 (Blommers, et al., 1997) and VP16/hTAF1132 (Goodrich, et al., 1993). 32 Hap4 and VP16 are both members of the acidic class of transcriptional activators. However, a fundamental difference in the two activation domains is that in yeast, Gal4- VP16 is unable to stimulate the expression of a lacZ reporter gene in the absence of Ada2, Ada3, or Gcn5 while Gal4-Hap4 is able to stimulate expression of the same lacZ reporter under such conditions (Horiuchi, et al., 1995). This suggests that a qualitative difference exist between the VP16 activation domain and the Hap4 activation domain. Hap4 was also shown to be capable of activating transcription in the absence of Ada2 when acting in concert with Hap2/3/5 at a natural promoter (Berger, et al., 1992). The observation that Hap4 (aa330-554) is able to function independent of Ada2, Ada3, and Gcn5 while VP16 tested in the same context is dependent on Ada2, Ada3, and Gcn5 indicates that the Hap4 activation domain is intrinsically Ada2, Ada3, and Gcn5 independent. To define such an ADA independent activation domain, identification and characterization of the Hap4 minimal activation domain(s) was undertaken. This work thus represents the only known dissection of an ADA-independent activator. I constructed an extensive set of deletion mutants of the Hap4 330-554 region, fused to the lexA DNA binding domain present within amino acids 1-202, and tested their ability to activate transcription of a lacZ reporter gene. This approach resulted in the identification of the Hap4 359-476 region as a potent Ada2, Ada3, and Gcn5 independent activation domain. Resident within the 359-476 region of Hap4 are seven clusters of hydrophobic amino acids with 2-4 hydrophobic amino acids present per cluster. I constructed Hap4 activation mutants in which serine was substituted for all of the hydrophobic residues within any given cluster to serines and tested the effect on trans-activation potential. I found that L36SS/L366S and I37OS/L37IS mutations had no 33 effect on the activation potential of Hap4 359-476 and I4508/L452S had only a moderate negative effect. Hap4 359-476 Y427S/L428S/F429S/L43OS resulted in a 5-fold reduction in activation potential. In contrast, I39OS/W39lS/Y393S/L394S, F456S/Y4588, and L466S/M467S resulted in a greater than 10-fold reduction in activation potential. The original work on Hap4 was carried out in the yeast strain BWGl-7a (Forsburg and Guarente, 1989). In BWGl-7a the 1-330 region of Hap4 was unable to support growth on a non-fermentable carbon source. However, when tested in the more robust yeast strain DMYl48, Hap4 1-330 was able to support growth on a non- fermentable carbon source implying the existence of an activation domain within this region. The sequences needed for efficient transcriptional activation when tethered to lexA extended from amino acid 124 to amino acid 270 and those sequences needed for maximal activation potential extended from amino acid 124 to amino acid 329. However, both segments require GCN5 for efficient activation. The 124-329 region depends on hydrophobic residues for it trans-activation function. An F148S/L149S/F1518 mutant was unable to direct the activated transcription on an integrated lacZ reporter gene. EXPERIMENTAL METHODS Yeast Strains The yeast strains used for the experiments detailed in this chapter are described in Table 3. 34 Deletion Mutagenesis of Hap4 aa330-554 Deletion mutants of the HAP4 gene that encodes the domain of Hap4 from amino acid 330-554 were constructed to identify a minimal activation domain. The initial constructs were fused to the Gal4 DNA binding domain (aa1-147). Hap4 deletion mutants were also studied as fusion proteins joined to lexA. The plasmid pJLSSS is an ARS/CEN recipient plasmid for mutants of the HAP4 gene fragment encoding amino acids 330-554. pJLSSS provides the Gal4 DNA-binding domain (aa1-147). This plasmid was constructed by subcloning the SacI-BamHI fragment from pJLS34, which contains the Gal4-Hap4 fragment with the ADHI promoter upstream, into the SacI and BamHI sites of pRS315. 1. Amino terminal deletions of Hap4 (aa330-554) were obtained by linearizing pJLSSS with XhaI followed by ExaIII/Sl treatment for various amounts of time. Following the EonII/Sl treatment the DNA was cut with SacI and then subcloned into a vector obtained by digesting pJLSSS with Sacl and SmaI. Deletion endpoints and reading frames were confirmed by dideoxy sequencing. Three HAP4 deletion mutants were selected from this set, with amino terminal truncations at codons 359, 400, and 424. Carboxy terminal deletions were obtained by employing unique XbaI and Bng sites present within HAP4 and combining them with the above described amino terminus deletion endpoints. HAP4 DNA that already encoded the amino terminal truncations previously described was digested with either XbaI or Bng followed by a fill-in reaction using the Klenaw fragment of E. cali DNA polymerase I followed by digestion with XhaI. The resulting fragment was subcloned into XhaI-Smal treated pJLS114. 35 An internal deletion mutant removing codons 424 to 476 was constructed by introducing a Bng site after the codon for aa423 of Hap4 that would exist in addition to the Bng site present after codon 476. Subsequent digestion of this construct with Bng followed by ligation resulted in the desired deletion. The additional BglII site was introduced by PCR using primers ST65 and ST66. The PCR product, spanning codons 330 to 424 (including the desired BglII site) was cloned first into the TA cloning vector from Invitrogen. The resulting plasmid was then digested with HpaI and BglII to release the HAP4 fragment encoding amino acids 330-424. This fragment was subcloned into SmaI-BglII treated pJLSSS, which provided the Gal4 DNA-binding domain and also HAP4 codons 476-554. Plasmids expressing lexA-Hap4 fusion proteins were constructed using various combinations of the PCR oligonucleotides shown in Table 1. Two plasmids were central to the subcloning strategy. The plasmid pDB20.lexA is a 211 origin, URA3 marked plasmid that contains the ADH I promoter and terminator flanking lexA. A polylinker with the sequence CTG-GCG-GCC-GCG-CGC-AAG-CTT is present immediately downstream of lexA (the CT G in bold type is the final lexA codon that immediately precedes the polylinker). The Not! site in the polylinker, highlighted in italic type, was used in all subclones. The plasmid pRS4l4 (Sikorski and Hieter, 1989) is a low copy TRPI marked plasmid. PCR fragments spanning Hap4 codons 330-424, 330- 476, 424- 476, 424-554, 476-554, or 13424-476 were cloned into the Invitrogen TA vector, excised with Bag] and ligated into pDB20.lexA. For all other constructs discussed, PCR products were digested with Bag] and subcloned directly into pDB20.lexA. Once in pDB20.lexA, the DNA fragments encoding the various lexA-Hap4 fusion proteins flanked by the 36 ADHI promoter and terminator were subcloned into pRS414 either as XmaI-Sall, Sac]- SalI, or BamHI fragments. All sequences were verified by dideoxy sequencing. Deletion mutagenesis of Hap4 aal-330 To define a minimal activating region within Hap4 amino acid 1-330, I constructed deletion mutants of the HAP4 gene fragment encoding amino acids 1-330 and fused them to lexA. These deletion mutants were constructed using various combinations of PCR oligonucleotides shown in Table l. The PCR products were digested with EagI and subcloned directly into pDB20.lexA. Once in pDB20.lexA, the DNA fragments encoding the various lexA-Hap4 fusion proteins flanked by the ADHI promoter and terminator were subcloned into pRS414 either as XmaI-SalI, SacI-Sall, or BamHI fragments. All sequences were verified by dideoxy sequencing. B-Galactosidase Assays Plasmids expressing lexA-Hap4 fusions were transformed into yeast strains JSY03, JSY05, or L40, all of which have an integrated lacZ gene with eight lexA binding sites upstream. Eight ml cultures were grown to OD595=0.5. B-Gal assays were performed as described (Guarente, 1983; Miller, 1972). Three independent assays were performed for each mutant. Immunoblot Assays To assess the steady-state level of each lexA-Hap4 fusion protein it was necessary to use a plasmid with a 2n origin transformed into yeast strain J YSO3. 10 ml cultures 37 were grown to an approximate OD595=0.5, harvested by centrifugation, and resuspended in 0.25 ml of buffer containing 100 mM Tris-Cl pH 8.0, 20% glycerol, lmM DTT. Extracts were prepared by vortexing six times at top speed in 15-second bursts, waiting on ice 15-seconds between bursts, in the presence of glass beads (425-600 microns from Sigma catalog # G-8772) added to the level just below the meniscus. Extracts were transferred to clean microfuge tubes and centrifuged at 4° C 14000 rpm for 5 min. Relative protein concentrations were determined for each extract using a dye-binding assay (Bradford, 1976). Equal amounts of protein were loaded onto 12.5% SDS-PAGE gels. Following electrophoresis, proteins were electrophoretically transferred onto nitrocellulose. Membranes were blocked in 7% powdered milk, 20 mM Tris-Cl pH 7.5, 137 mM NaCl, 0.0038M hydrochloric acid, 0.01 % Tween 20 and then incubated with a mixture of three different mouse monoclonal antibodies (mABs) raised against lexA produced by Yuri Nedialkov: 2 jig/ml YN-lexA-2-l2, 4.3 [lg/ml YN-lexA-6-10, and 2 ug/ml YN-lexA-l6-7. Incubation with primary mABs was followed by incubation with a secondary antibody directed against mouse IgG, conjugated to horseradish peroxidase (Sigma). Milk (1.5%) in 20 mM Tris-Cl pH 7.5, 137 mM NaCl, 1M hydrochloric acid, and 0.01% Tween 20 was included as a blocker in all antibody incubations. Blots were developed using an enhanced chemiluminesence system (ECL, Amersham or Renaissance, DuPont NEN Life Science Products). Point Mutant Construction Clustered point mutations within the HAP4 region from codons 359 to 476 were constructed using the QuickChange mutagenesis kit (Stratagene, La Jolla, CA). The 38 oligonucleotides used are described in Table 2. The mutations were introduced directly into a low copy TRPl-marked plasmid from which lexA-Hap4 fusion protein was expressed from the ADHI promoter. The presence of the correct base-pair changes and the absence of secondary mutations were confirmed by DNA sequence analysis. Growth Curve To compare the ability of Hap4 aal -330 to support growth on a non-fermentable carbon source with that of full length Hap4(aa1-554) 1 performed a growth curve experiment. The yeast strain DMYl48 was transformed with plasmids encoding either wild type or mutant Hap4 protein. DMYl48 transforrnants were initially grown to saturation (OD595 >15) in selective glucose-containing media. A portion of the saturated culture, 0.2 ml, was used to inoculate a 100 ml culture of YP-lactate. Growth was monitored by OD595. 39 8:22: 8E0<0E080EOQO>nO>OOO>OQ>> QOQOQOQOQOQOOQOOOQA4>OO HQQHQHH>QQQOO>>O>E>> 00880088000808?» HQ>OOOO>>>Q>§QH>OH QOQOQOQOQOQOOQQOOQA4>AA HOOHQOQHQQHQOOHQEG>A4 QCOOQOQOQOQOQOOQOOOQE >>>HQSQQOOO0>>O>QO>QQ OOQOQOQOQOQOQOOQOOQOHO Qjagmngngnn>an>a an QCQOQOQOQOROOQOOQfiwga O>QQQ>>HQ>>QO§>HQ>OQ 0.5.0.3. QOQOOOQOOOQOOQOOOO>Q>> >>Q>OQ>>>>H>OO>>Q>O>H> .000.» ”932: 533.030: C002 0182. :08 m9. 0: 832.:on 53 comm: m" 8.3m 0m :00? 20: a8 man—:93. 822. 0152 58 8 858:8 0 $00 con: 0:2 8:8 om $00? 08% a8 man—=08. C002 0182. :80 no: E. 83588 90: cam? m. 8: 8. C00? 20: a8 30:58. roads 0132 58 8 50.0980 0 «80 none: 8.8.. 003A 0». C008: mmmm a8 man—:93. C003 0132. .53 mo: 0: 83588 9”: comm: 0" 000m 0». 300A. 20: £8 man—anon. C003 0132. 53 mo: 0: 83588 :5: comm: 0: 83¢ 0m $009 20: a8 man—:08. 882 0:32. 53 8 530908 0 £00 none: 03% Ego 0m :00? THE a8 man—sang. C009. 0132 53 mom 0: 83388 50: cam? m: 85¢ 0m $009 20: 38 32:03. 208” >= :002 01:53 :88 :88. £5. :5 880030: 0». mag 25 mafia. 08 985:8 8 co 3-355 2:: Ex.» cmmam =5 20: man 5 3o 02%:38: o." 0250. .25 3:935 80:38 man—:98 5o Em: 0000: «0195 man So 00.0—3on 5 0wao :59 90 20: 28 5 85 012” OHQ-QOQ.QOO.GOQ-OQO->>Q-OHH. 41 .934 28 dam? .mamk’ .moamH E 838m 652% 9 EcoE=mEoo BEE“ .833 .334 ES .mmam> dam? .moofl E 833% .gmhm 2 Ecofimifioo BEE Saab .8022 98 m3: E 9.36m .Emhm 8 Suaofizmfiou ESE .833 .mme Ea £59 .mmmE Eu: 5 938m .momhm 8 Ecogafioo BET—m 533 .89: Ba 33: XE a £23m .momhm 2 Econ—macho“. BEEQ $33 .meA 93 wok: 3.5 E 3.30% .ommhm 8 Emanciafioo 5th $33 .mSS 23 803 is: 5 338m .wmmbm 8 abacoEm—mEoo “255 533 .mcov—Z Ea m3: 5 838m don km 9 3358.388 BEE .253 .3va Ba fin; .mmmi Eu: 5 938m .ocmhm 2 5582988 5.55 Saab .8? Ba m3: an: a 338m .95 Pm 9 Ecofizmfioo Baum 83D .mgg 93 wet: van: 5 8.3.3— 602% 2 Ecofizmfioo BEE“ 3&3 .mbcmq Ea mowmq v93.— E 338% .momhm 8 idEoE=QEoo BEE 5&3 cormsbefi Ego—om UUEUUCOH UH>QO>>>O>>OQ>HQ>>>OQ§§QAQHO>QO>QO>QQ> $69 @159 coat—mamas 8 madam. HQEQOQQO Nam—:8 m: 4A3? :Nmm. mhwom. 2E swam. O>QO§O>HOQQ>QOOHQ>Q>>Q>>Q>>Q>QQ§O>HOQH 532 @152 ooBEmBoamQ 8 mqum. €300.38 W035 m: 4&3? Ewmm. 950m. E5 :25. 43 Strain Genotype Reference DMYl48 ura3-52, leuZDl, his3DZOO, trpID63, D. McNabb, personal hap4::hisG hap5::hisG communication MAT-a JSY03 ura3::lexA80p-lacZ, lysZ::lexA8op-URA3, This work leu2-3,112, trpIA99, adeZ, his34200 MAT—a JSY05 ura3::lexA8op-lacZ, lysZ::lexA8op-URA3, This work leu2-3,112, trpl A99, ade2, his3A200, gcn5::hisG MAT-a L40 his3A200, trp1-901, leu2-3,112, ade2, lysZ- (Hollenberg, et al., I 995) 801 am, URA3::(lexA0p)8-lacZ, LYS2::(lexA0p)4-HIS3 MAT-a TABLE 3. Description of yeast strains. 44 RESULTS The observation that HAP4 (aa330-554) is able to function independent of ADA2, ADA3, and GCN5 while VP16 tested in the same context is dependent on ADA2, ADA3, and GCN5 indicates that the HAP4 activation domain is intrinsically ADA2, ADA3, and GCN5 independent. I wanted to know what defined such an ADA independent activation domain and thus I set out to identify and characterize the HAP4 activation domain. I constructed an extensive set of deletion mutants of the HAP4 330—554 region, fused to the lexA DNA binding domain, and tested their ability to activate transcription of a lacZ reporter gene. The transcriptional activity of these mutants was intially tested using a lacZ reporter gene on a high copy (211.) plasmid, pLGSDS, in the yeast strain PSY316. The results obtained using this system were fraught with significant intra- and inter-assay fluctuations, making analysis difficult. Hoping to eliminate such fluctuations, I constructed the yeast strain J SY03 (genotype shown in Table 3) that had an integrated lacZ reporter gene with eight lexA binding sites upstream. Quantitation of transcriptional potential of the deletion mutants using J SY03 is shown in Figure 1. Neither N-terminal deletion to residue 359 that resulted in the 359- 476 construct, nor C-terminal deletion to residue 476 that resulted in the 330-476 construct, caused in a significant loss of activation potential. Further N-terminal deletion to residue 400 or C-terminal deletion to residue 464 resulted in a reduced effect on activation. The positive contribution made by the 464-476 region is obvious (compare 330-476 with 330—464, 359-476 with 359-464, and 400-476 with 400-464). The 464-476 region is essential, however, only when aa400 is the N-terminal boundary. Otherwise, 45 the 464-476 region is necessary only to achieve maximal activation (compare 330-476 with 330—464 and 359-476 with 359-464). 46 Hap4 Segment Fused to lexA B-Galactosidase Units U 390:110 359 464 160:10 803140 533 230140 540:40 3:2 020 17301260 FIGURE 1. Specific activities of deletion mutants of the Hap4 C-terminal domain (aa 330-554). The indicated Hap4 regions were fused to lexA and expressed from the ADHI promoter on TRPI marked ARS/CEN plasmids. The lexA-Hap4 expression plasmids were transformed into J SY03 which contains an integrated lacZ reporter gene with eight upstream lexA binding sites. B-galactosidase units were measured in extracts of cells grown to mid-log phase, and are reported as nmoles of o-nitrophenol/minute/mg protein. lexA alone has no activation potential. 47 Curiously, the deletion mutant lacking the 424-476 region is active although neither half has significant activity. It is possible that by joining amino acid 423 with amino acid 477 an artificial activation domain was created. Alternatively, this may be a case of potent synergy between two very weak activation domains, as neither the 330-424 nor the 476-554 region possessed any significant independent activation potential. However, this result does seem to indicate that while the 359-476 region is sufficient to activate transcription, it is not necessary for such. Immunoblot analysis shown in Figure 2 indicate that all mutant lexA-Hap4 fusion proteins are present in detectable, although different, amounts except 330-554. Thus differences in trans-activation potential between deletion mutants cannot be attributed to differences in expression or stability of the deletion mutant proteins. It should be noted that a low copy ARS/CEN plasmid was used to express the lexA-HAP4 fusion proteins in the B-galactosidase assay whereas a high copy 2}; plasmid was used to express the same constructs for the immunoblot. This is necessary because the ARS/CEN plasmids did not produce detectable amounts of protein. The results shown in Figure] demonstrate that the boundaries of the activation domain lie between amino acids 359-400 and amino acids 464-476. As shown in the cases of VP16 (Cress and Triezenberg, 1991; Regier, et al., 1993; Sullivan, et al., 1998), p53 (Horikoshi, et al., 1995) and GCN4 (Drysdale, et al., 1995) activation domains are dependent on hydrophobic residues. The 359—476 region of Hap4 possesses seven different clusters of hydrophobic amino acids: L365/L366, 1370/L371, 1390/W391/Y393/L394, Y427/L428/F429/L430, I4SOIIA52, F456/Y 458, and L466/M467 (also seen in Figure3). We hypothesized that some of these clusters would 48 FIGURE 2. A. Immunoblot analysis of yeast cells expressing deletion mutants of the HAP4 gene fused to lexA. B. Immunoblot analysis of yeast cells expressing deletion mutants and point mutants of the HAP4 gene fused to lexA. The indicated HAP4 deletion mutants were fused to lexA and expressed from the ADHI promoter on a URA3 marked 21.1 plasmid. Aliquots of whole cell extracts were electrophoresed in SDS polyacrylamide gels and blotted to nitrocellulose. The filters were incubated with a mixture of three different polyclonal serums directed against the lexA protein and developed using enhanced chemiluminescence. 49 359-476 124-329 50 365 370 NDDNMSL LNLPI LiEETVSSG 390 I WNY LJPSSSSQQ DDKVEEN QDSSRALKKNTNSEKAN' Q 427 AKNDETYL F LQDQDESADS 450 456 HHHDELGSEITLAQDKFSY LI 466 PPTLEEL EEQDCNNGRS FIGURE 3. The amino acid sequence of the Hap4 359—476 region. Hydrophobic residues are in larger type. Boxed areas highlight clusters of hydrophobic amino acids. Each hydrophobic residue within a given cluster was changed to serine. The resulting mutant was expressed as a lexA fusion from the ADHI promoter on aTRPI marked ARS/CEN plasmid. The lexA-Hap4 expression plasmids were transformed into JSY03 and tested for the ability to activate transcription of a lacZ reporter gene. 51 be essential for this region of Hap4 to active transcription. Therefore mutants were generated that encoded a change of each hydrophobic amino acid within any given cluster to serine. The resulting mutants were tested for their ability to activate transcription of an integrated lacZ reporter gene with eight lexA binding sites upstream. The results shown in Figure 4 reveal that L36SS/L366S, 137OS/L37IS had no effect on the activation potential of Hap4 359-476 and I4SOS/L4528 had only a moderate negative effect. Hap4 359-476 Y427S/L428S/F429S/L43OS resulted in a S-fold reduction in activation potential. In contrast, 139OS/W39IS/Y393S/L394S, F456S/Y4588, and L466S/M467S resulted in a greater than lO-fold reduction in activation potential. These results are consistent with the previous analysis of the deletion mutants. The region between amino acid 359 and 400 of Hap4 was necessary only for maximal activation in the absence of GCN5 (see Figure 10) and neither L36SS/L3668 nor I37OS/L37 18 had any effect on the activation potential of Hap4 359- 476. The region between amino acid 464 and 476 of Hap4 is essential for the activation potential of the 359-476 region and L466S/M467S resulted in a greater than 10—fold reduction in activation potential. The immunoblot seen in Figure 5 reveals that all of the mutations that resulted in a decrease in trans-activation potential had no effect on protein stability. Thus differences in trans-activation potential between mutants cannot be attributed to differences in expression or stability of the mutant proteins. To determine whether or not the 359-476 region was sufficient to carry out all known transcriptional activation functions of Hap4, a fusion of the 359-476 region to Hap4 1-330 was planned. The possibility that Hap4 possesses multiple activation domains was raised when in the course of this pursuit Hap4 1-330 was unexpectedly 52 .ESEQ wE§:EE\_o=ofiobE-o do 338: 3 .3332: 033 3:5 emaEona—awé .o_£w=wo= 0.33 «as: bee wEEeEOo 32238 E 3238 o3236£awé .3523 .0:an 025 Soc 335% =3 “30» E emaEona—am -n we c.2338 Ewe—5m 53c >33qu 53.: 836.: mam d: 58% use» 05 Go 088% 05 BE 3223:: Sofia: 3% weaves 58— Emfi 55 23w MS: a mo gimp—axe 83:8 9 3:3“ 05 8a. 338a 93 $8— 2 come 33 v Bswfi E noncomov 3532: as: 2: no comm damp. thonm 3mm 05 .3 33:83 :03?on 05 2 3:23.. 0325063 .3 5333.80 2E. .v .5503 $09.. 9mm.— mmmVJ mmNQn— mam»; whee: mag; mum‘— mmNVJ w an; m En.— momma ._.>> ”00%.. womfiu. mom: whNV> mama. was”. mac”.— cam cccw comw suun esepgsotoews-etea cooN 53 WT 124-329 L139S/Il40S F456S/Y 4SSS/L459S/L46GS/M467 S L466S/M467 S F456S/Y4588/L459S I4SOS/L4SZS Y427S/L4288/F429S/L4308 139OS/W39IS/Y393S/L394S I37OS/L37 lS L3658/L3668 WT 359-476 FIGURE 5. Immunoblot analysis of yeast cells expressing point mutants of the HAP4 gene fused to lexA. The indicated HAP4 point mutants were fused to lexA and expressed from the ADHI promoter on URA3 marked Zu plasmid. Aliquots of whole cell extracts were electrophoresed in SDS polyacrylamide gels and blotted to nitrocellulose. The filters were incubated with a mixture of three different monoclonal serums directed against the lexA protein and developed using enhanced chemiluminescence. 54 discovered to be able to support growth on a non-ferrnentable carbon source. Inferred from this result was that the 1-330 region of Hap4 possesses transcriptional activation potential, despite a previous report to the contrary (Forsburg and Guarente, 1989). The original Hap4 work was carried out in the yeast strain BWGl-7a (Forsburg and Guarente, 1989). In this context the 1-330 region of Hap4 was unable to support growth on a non- ferrnentable carbon source. However, when tested in DMYl48, a more robust yeast strain, Hap4 1-330 was able to support growth on a non-fermentable carbon source (Figure 6). The less potent activation domain resident within 1-330 is unable to support growth of some yeast strains on a non-fermentable carbon source but does nonetheless possess transcriptional activation potential. When tested directly, the 1-330 region of Hap4 when tethered to lexA was able to activate transcription of a lacZ reporter gene, however with significantly less vigor than 330-554 (see Figure 7). The boundaries of the activation domain resident within amino acids 1-330 of Hap4 were thus sought. Deletion mutants of Hap4 (1-330) fused to the lexA DNA binding domain were constructed as shown in Figure 7 and were tested for their activation potential using the integrated lacZ reporter strain JSY03. Deletion of the N-terminus up to residue 123 resulted in no loss of activation potential; however, further deletion to amino acid 160 resulted in an almost complete loss of activation potential. Thus the N-terminal boundary of the activation domain present within 1-330 must lie between amino acid 124 and amino acid 160. Deletion of C-terminal Hap4 sequences to aa224 resulted in a total loss of activation potential (see result for 1-224 in Figure 7) whereas Hap4 (1-270) maintained activation potential. The C-terminal boundary of sequences needed for efficient 55 00595 + CONTROL + HAP41-554 1 .5 1 0.5 0 0 20 40 60 80 100 TIME (HOURS) FIGURE 6. Hap4 (aal-330) is capable of supporting growth of the yeast strain DMYl48 in lactate medium. Growth assesement was carried out in liquid medium minus tryptophan containing 2% lactate as the sole carbon source. DMYl48 yeast were transformed with TRPI marked ARS/CEN plasmids expressing either Hap4 1-330, Hap4 1-554, or the empty parental vector. Single colonies were picked and grown to saturation in media lacking tryptophan and containing 2% glucose. These cultures were used ti inoculate media lacking tryptophan and containing 2% lactate as the sole carbon source. OD595 readings were recorded at the indicated times. 56 Hap4 Segment Fused to lexA B-Galactosidase Units 1#0 40° 1 3° 1 103 2 i 1 1_224 2° 1 3 96—224 70 11° ”—270 28° 3 6° 96—329 74° 315° ”#29 928 3- 333 1H0 6 i 2 1H” 2° 1110 217 329 48 .1: 12 FIGURE 7. Specific activities of deletion mutants of Hap4 N-terminal domain (aa 1-330). The indicated Hap4 regions were fused to lexA and expressed from the ADH I promoter on TRPI marked ARS/CEN plasmids. The lexA-Hap4 expression plasmids were transformed into JSY03 which contains an integrated lacZ reporter gene with eight upstream lexA binding sites, B-galactosidase units were measured in extracts of cells grown to mid-log phase, and are reported as nmoles of o-nitrophenol/minute/mg protein. lexA alone has no activation potential. 57 transcriptional activation therefore lies between amino acid 224 and amino acid 270. Immunoblot analysis indicated that the deletion mutant proteins are expressed in detectable, although not similar, amounts with the exception of 96-329 and 124-329. For the reasons stated above, high copy plasmids were used for the immunoblot analysis. We hypothesized that the 124-329 region of Hap4 would also be dependent on hydrophobic amino acids for the trans-activation function that it possessed. Since the 124-160 region of Hap4 was essential for the trans-activation capability of the 124-329 region we looked in this region of the protein for clusters of hydrophobic amino acids. Two hydrophobic clusters within the 124-160 region were identified and point mutants where serine was substituted for most of the residues within the clusters were constructed (see Figure 8). The mutants constructed were Ll39S/Il4OS and Fl488/L149S/F1518 and their trans-activation potential was quantified. The results shown in Figure 9 reveal that Ll39S/Il4OS mutations had little to no effect on the trans-activation potential of the 124- 329 region of Hap4 whereas the Fl488/Ll49S/F lSlS mutation resulted in a complete loss of activation potential of this region. The determination of the relative abundance of the Ll39S/Il4OS mutant protein is shown in Figure 5. The determination of the relative abundance of the F1488/Ll49S/F1515 mutant protein is underway. The ability of Hap4 to activate transcription is less dependent on Ada2, Ada3, and Gcn5 than other transcriptional activators (e.g. Gcn4 and VP16) (Berger, et al., 1992; Pina, et al., 1993). Thus the ability to activate transcription in the absence of any of these factors should be a reasonable “quality control” test for any activation domain identified. The ability of 359—476 and 124-329 to function as transcriptional activators in the Agcn5 strain JSY05 was therefore tested. Both 124-329 and 359-476 are dependent on Gcn5 to 58 EKRKIGRHIQTIDE KLINDSNYLAFLKFD DLENEKFR FIGURE 8. The amino acid sequence of the Hap4 124-160 region. Hydropho- bic residues are in larger type. Boxed areas highlight clusters of hydrophobic amino acids. Each hydrophobic residue within a given cluster was changed to serine. The resulting mutant was expressed as a lexA fusion from the ADHI promoter on aTRPI marked ARS/CEN plasmid. The lexA-Hap4 expression plasmids were transformed into JSY03 and tested for the ability to activate transcription of a lacZ reporter gene. 59 1250T .3 5 1000- G 0 fl 2 750- n a 500- 3? 8 250- o. WT L139$ F1488 I14OS L149$ F1513 FIGURE 9. The contribution of hydrophobic residues to the activation potential of the Hap4 124-329 region. Each of the Hap4 mutants described in Figure 4 was fused to lexA and assayed for the ability to activate expression of a lacZ gene with eight lexA binding sites upstream integrated into the genome of the yeast strain L40. Bars indicate mean activity (with standard deviation) of B-galactosidase in yeast cell abstracts from three parallel cultures. B-galactosidase activities in extracts containing only lexA were negligible. B-galactosidase units were measured as nmoles of o-nitrophenol/minute/mg protein. 60 attain maximal transcriptional activation potential. However, the absence of Gcn5 more significantly affects the activation potential of the 124-329 region (Figure 10). The activation potential of the 124-329 region is reduced 4.2 fold in the absence of Gcn5 whereas the activation potential of the 359-476 region is only reduced 1.6 fold in the absence of Gcn5. A corallary of this finding is that the 1-330 region of Hap4 might not be able to support growth on a non-fermentable carbon source in the absence of Gcn5. When growth of DMY45 harboring a plasmid that expressed either Hap4 1-330 or Hap4 1-554 on a non-fermentable carbon source was monitored, Hap4 1—330 was unable to support growth in this media whereas Hap4 1-554 was able to support growth (Figure l 1). Thus it appears that the 124-329 region is a Gcn5 dependent activation domain whereas the 359—476 region is a Gcn5 independent activation domain. 61 EEEBOQ =ou~>som o: 35 0:2“ w~ 356 9E c2583.“: 033 mEEmEa =Ew3Exo Emméxxnz 2F .mEEmEm ZmU\mm< nod—SE FDR no 38an ~EQ< 05 Set “.3353 98 «£2 9 32¢ 225 waEme 3mm BEEvE 2E. .ocom N63 a .«o 5:933“: 0338 8 3533c van: Go ban“ 2: so 08% mo 53.8 mo 30km .2 EDGE vvvcszsclr 6 mmmmmmmmmwss- 9897.“... Z muwuwuwuwaeum is... — a W -3» 1 O S .cfim e S a -8an 93B m. fl £66. -82: 62 .8332: 3:5 05 an 8288 803 $538 nomQO 68:8 conga 28 05 mm 8802 §N 9:58:00 28 523999 was—om. ~62: 82:85 9 wow: 803 35:3 82: .8833 §~ 9:59:00 23 5:83.05 wig—02 «:58 5 5:83am 8 550% can now—0E 203 moEo—oo 2mEm .883 3:23 595 05 .5 6mm; was: .984 353 558 wcfimoaxo mEEmmE ZmU\mm< v8.38 ~$§ 5:5 watchman: 203 “89» $99. .0958 E558 20m 05 mm 8802 §N mama—3:8 Swag“? 35E «62: Raw: 5 :5 63.58 33 32:88.8 530.5 .58» flaw E «:58 898— E 53on comma 9 23 8mm; 33 33A .2 ”$505 €85 we: 3:.- M2 . 33:2: _ 8222: In... :20 «28> IT . 2 63 Discussion The use of deletion mutagenesis as a tool for the study of activation domain structure and function has yielded a great deal of information. This strategy has resulted in the demonstration that multiple activation regions can reside within a single protein, as in the cases of GCN4 (Drysdale, et al., 1995; Jackson, et al., 1996), VP16 (Cress and Triezenberg, 1991; Regier, et al., 1993; Sullivan, et al., 1998; Walker, et al., 1993), p53 (Lin, et al., 1994), and RelA (Blair, et al., 1994). Detailed study of each motif independently has revealed a functional dependence of each of these on the hydrophobic character of specific amino acids. These studies imply that activation domains are constructed from repeated motifs that depend on hydrophobic residues. VP16 is an example of a transcriptional activator that contains multiple regions, each dependent on one or two particular residues, capable of transcriptional activation. The activity of the VP16N region is dependent on F442 (Cress and Triezenberg, 1991; Regier, et al., 1993), and the activity of the VP16C region is dependent on F473/F475 for the ability to activate transcription (Sullivan, et al., 1998). In contrast, GCN4 contains seven hydrophobic clusters that make redundant contributions to transcriptional activation (Jackson, et al., 1996). These examples demonstrate the existence of regions of varying complexity. Despite a similar dependence on key amino acids with hydrophobic sidechains, transcriptional activators function via distinct mechanisms. The ADA-dependent versus ADA-independent phenomenon is an example of how some activators (e. g. VP16 and Gcn4) show a requirement for certain co-factors whereas others (e. g. Hap4 and Gal4) are able to function independently of the same co-factors. A similar theme can be seen for regions within a given protein. For instance, VP16C and VP16N have been shown to function via distinct mechanisms. The VP16N region efficiently de-represses an integrated reporter in mammalian cells, whereas the VP16C region is unable to do so (Horn, 1999). In contrast, VP16C stimulates TFIID recruitment via TFIIA in vitro whereas VP16N is unable to do so (Kobayashi, et al., 1998). Tethering regions together that employ different mechanisms of action results in synergy with respect to trans- activation potential. This synergistic affect can be seen when comparing the ability of VP16N with that of VP16N+C to activate transcription. A similar result has been demonstrated for Spl and VP16N+C (He and Weintraub, 1998). It is purported that the multimerization of activation domains of distinct functions (i.e. Spl and VP16 N+C) results in synergistic increases in activator potency. However, evidence for “domain synergy” also exists whereby multimerization of two or more identical subunits also results in synergistic increases in trans-activation potential (Emami and Carey, 1992). Thus synergy does not require the presence of two distinct regions. The yeast activator Hap4 is an example of a prototypical ADA-independent transcriptional activator. To better understand the essential characteristics of an ADA- independent activator, we set out to define and characterize the region(s) within Hap4. Deletion mutant analysis revealed that two activation regions were resident within Hap4. The N-terminal region resides within the 124-329 region of Hap4 and the C- temlinal region resides within the 359-476 region of Hap4. Identification of the N- terminal region represents a novel finding, as previous results indicated that the 1-330 region of Hap4 was devoid of trans-activation potential (Forsburg and Guarente, 1989). 65 Hap4N represents an ADA-dependent region whereas Hap4C represents an ADA- independent region and thus the implication is that these two regions activate transcription through unique pathways. The relative activities of the deletion mutant proteins shown in Figures 1 and 7 may be misleading because all deletion mutant proteins were not present at equivalent levels. Equal amounts of total protein were loaded into each lane shown in Figure 2, however, the immunoblot indicates that the amount of lexA-Hap4 present varies. Therefore, for instance, it may not be true that lexA-Hap4 (aa330-476) is 2-fold more potent a transcriptional activator than lexA-Hap4 (aa330-554) because, as seen in Figure 2, there is a great deal more lexA-Hap4 (a330-476) present than lexA-Hap4 (aa330-554). Thus the specific activity of lexA-Hap4 (aa330-554) is actually likely to be much higher than that of lexA-Hap4 (aa330-476). We used a low copy vector in the activity assay and a high copy vector in the immunoblot because the amount of protein produced from the low c0py vector was undetectable by immunoblot. We are unable to use the high copy vector in the activity assay because over-expression of Hap4 (aa330-554) is toxic. However, while we cannot confidently assess the specific activities of the various constructs, we are able to accurately identify minimal Hap4N and Hap4C activation regions. Both Hap4N and Hap4C depend on hydrophobic residues for the ability to activate transcription. The effect of mutation of clusters of hydrophobic amino acids to serine within Hap4 359-476 and 124-329 can be grouped into three distinct classes: mutations that have no effect on trans-activation potential of Hap4 359-476 (L139S/Il4OS, L36SS/L366S, and I37OS/L37IS), mutations that had a moderate negative 66 effect on the activation potential (I4508/L4528), and mutations that had a significant negative effect (greater than S-fold) on the trans-activation potential (F 1488/Ll49S/F ISIS, 139OSIW39lS/Y393S/L394S, F456S/Y458S, and L4668/M467S). I will consider possible explanations for these results below. The C-region of Hap4 contains two clusters, L365/L366 and I370/L371, and the N-region one cluster, Ll39/Il40, that when all of the hydrophobic residues present within them are changed to serine, have no effect on trans-activation potential. The simplest interpretation of this result is that these residues play no role in the activation of transcription. Of course, it is also possible that these clusters play redundant roles. In the C-region, L365/L366 and 1370/L37l may be redundant to themselves or to other unidentified elements within the 359-476 region of Hap4. To determine whether they are redundant to each other, simultaneous mutation of L365, L366, 1370, and L371 to serine would be necessary. In the N-region other as yet unidentified hydrophobic residues that contribute to the activation potential of the 124-329 region may exist. A systematic mutagenesis scheme would answer this question. The I4508/L4528 mutation has a moderately negative effect on trans-activation potential of Hap4 359-476. Thus 1450 and L452 might contribute to the overall trans- activation potential of Hap 359~476 but are not the sole determinants of this. For example, these residues may partially contribute to direct interactions with target proteins but mutation of these two residues to serine does not totally abrogate the interaction. Alternatively, they may make structural contributions that are essential for target interaction. 67 The F148S/Ll49S/F1518 mutations within Hap4N and the I39OS/W39IS/Y393S/L394S, F456S/Y4588, and L466S/M467S mutations within Hap4C have a significant negative effect on the trans-activation potential of each sub- domain. These data indicate that these residues make critical contributions to target interaction, either directly or indirectly. By nature, hydrophobic amino acid side-chains have a propensity to be buried in protein tertiary structures. The finding that Hap4 is dependent on hydrophobic amino acids for the ability to activate transcription may indicate that the role of hydrophobic amino acids within the activation domain of Hap4 is to facilitate formation of a structure that is essential for Hap4 to activate transcription. However, this model becomes less attractive because other activation domains are typically unstructured unless accompanied by their target molecule. An alternative model proposes that critical hydrophobic residues remain exposed on the surface of the activation domain and directly interact with its target. The interaction of activation domain with target thus results in hydrophobic amino acid side-chains buried in protein quaternary structure. The central role of hydrophobic residues in directing protein-protein interactions has been demonstrated in the cases of the p53 transcriptional activation domain interacting with its repressor MDM2, glucocorticoid steroid receptor interacting with GRIP-l, and VP16 with TBP. Crystallographic analysis revealed that the essential hydrophobic amino acid side-chains of GRIP-l, p53, and SRC-l fits into a hydrophobic groove within the glucocorticoid receptor, MDM2, and the peroxisome proliferator- activated receptor-y respectively (Feng, et al., 1998; Kussie, et al., 1996; Nolte, et al., 1998). Fluorescence spectroscopy techniques indicated that regions surrounding both 68 F442 and F473 within VP16 are more ordered in structure in the presence of TBP as opposed to TFIIB (Shen, et al., 1996). We hypothesize that a similar hydrophobic surface of Hap4 may interact with target proteins at key steps along the pathway to transcriptional activation. The high copy suppression strategy described in the following chapter represents an attempt determine if the hydrophobic residues identified are involved in target protein interaction. The amino acid sequence surrounding L466 and M467 of Hap4 is similar to regions of GCN4 (L123 and F124) and VP16-C (M478 and F479) known to be important for their ability to function as transcriptional activators GDrysdale, etal., 1995; Sullivan, et al., 1998). In the sequence alignment displayed in Figure 12, L466 and M467 of Hap4 correspond to the residues M478, F479 of VP16 and L123, F124 of GCN4. However, both VP16 and GCN4 are Ada2, Ada3, and GCN5 dependent activators and thus the 466-467 region of Hap4 is not likely to be the sole determinant of the Ada2, Ada3, and GCN5 independent nature of the 359-476 region of Hap4. More likely candidates for the determination of Ada2, Ada3, and GCN5 independence are I390/W391/Y393/L394. Simultaneous mutation of each of these residues to serine results in an ablation of trans-activation potential. Deletion of amino acids 359-399 results in a significant loss of activation potential and more importantly a loss of GCN5 independence (see Figure l and Figure 9). Taken together, these results indicate that the determinants for Ada2, Ada3, and GCN5 independent activation reside within the 359—399 region of Hap4 and are likely to include I390/W391/Y393/L394 either totally or 69 €35 98 6200 65> .8 3388 c2338 Quasar—853 05 mo 303.8.“ .«o 53888 cocoawom .2 5505 Ezooa m m 92>; _ a z a §_o._m.._v_zn§ 35.. m ._.BmxmznmgzmTEVzoo o m ammo§ or; 70 in part. To test this hypothesis, we could compare the ability of the I39OS/W 391S/Y 393S/L394S and Y427S/14288/F429S/L43OS Hap4 mutants to activate transcription of the lacZ reporter gene in a gcn5 yeast strain, expecting that activation by the Y427S/L4288/F429S/L43OS mutant to be affected and the I39OS/W39lS/Y39BS/L394S mutant to be unaffected by the deletion of gcn5. The recent discovery of a Hap4 homologue in K. lactis (Bourgarel and Nguyen, 1999) is very interesting. The existence of a Hap4 homologue allows for cross species comparisons of the homologues and indeed the process has already begun. Sequence alignment shows only two well-conserved regions of the protein exist: part of the Hap2/3/5 interaction region and the region that corresponds to aa462-472 of Hap4 from S. cerevisiae (see Figure 13). Not only does the 462-472 region have a motif of amino acids that is reminiscent of GNC4 and VP16 activating regions as mentioned above, but my work has also shown the importance of the 462-472 region of S. cerevisiae Hap4, as the L4668/M467S double point mutant has the most severe effect on the activation potential of the 359-476 region of Hap4. I plan to focus my attention on this region and determine whether or not it is possible to identify its mechanistic target. If it is possible to identify the mechanistic target of this part of Hap4, it will be possible to look for conserved homologues in K. lactis as well. 71 an: 83 ~63?»ng .m can SEE 2v: MES .M E moocoacom 333:8 bra—5:226 05 mo confinemocme 23828 .3 mmDDE Vn—(Iow «hvacmmfiqmmdhmwov mhmlummmnm4>31¥whmg ton—Gmwfidmmqhn—oom mmmflmxmmna4>>>zmwhmcv 321.! 72 CHAPTER THREE Identification of Ha Activation Domain Tar et 8 INTRODUCTION An abundance of data indicates that transcriptional activators are able to physically interact with numerous factors. The relevance of these interactions is not always as clear. These factors fall into two different classes: general transcription factors (GTFs) and coactivators. Coactivators are distinct from GTFs in that they are dispensable for basal-level transcription in vitro but are required for activated transcription in vivo or in vitro. Our initial interest in Hap4 stemmed from the fact that while other activators required the coactivators Ada2, Ada3, and GCN5 for their ability to activate transcription in viva, Hap4 did not. If Hap4 does not target Ada2, Ada3, and GCN5, then what piece of the transcription apparatus does it target? It is possible that Hap4 targets other coactivators, GTFs, or both. Transcriptional activation requires accessory factors, variously termed co- activators, adaptors, and mediators, in addition to the GTF proteins needed for basal transcription. In some cases co-activators appear to bridge the interaction between gene- specific activator proteins and GTFs, whereas in other cases coactivators facilitate chromatin remodeling. Several functionally distinct classes of coactivators have been described in detail in the introductory chapter. These include TAF components of TFIID, the SRB/mediator complex that associates with RNAPH to form holoenzymes, SAGA and related complexes that catalyze nucleosomal histone acetylation, and SWIISNF and related chromatin remodeling complexes. The GTFs TFIIA, TFIIB, TBP, and TFIIH also 73 have been implicated as transcriptional activator targets. A more detailed description of their role in activated transcription follows. TFIIA is required for transcriptional activator response in viva (Ma, et al., 1993; Yokomori, et al., 1994). Transcriptional activators such as Zta, VP16, and T antigen bind directly to TFIIA, and this binding correlates with the ability to stimulate TFIIA-TFIID- TATA (DA) complex formation (Damania, et al., 1998; Kobayashi, et al., 1995; Kobayashi, et al., 1998; Ozer, et al., 1994; Sun, et al., 1994; Yokomori, et al., 1993). A related hypothesis is that TFIIA affects activator response by inducing an isomerization of the DA complex. Accumulation of this isomerized complex correlates with transactivation, although an inability to directly trap this intermediate complex precludes any direct observation of its activity relative to unisomerized PIC. Isomerization of the DA complex into a form capable of binding TFIIB is necessary for activation of transcription (Chi and Carey, 1996). Recently, a mutation in the small subunit of yeast TFIIA (T 0A2), W76A, that results in a loss of activation potential by Gal4-Hap4 (330- 554) was described (Ozer, et al., 1998). Thus TOA2 may be a target of the Hap4 activation domain resident within the 330—554 region of the protein. TFIIB has been implicated as the direct target of many gene-specific activators, leading to the proposal that certain activators stimulate transcription by TFIIB recruitment (Lin and Green, 1991; Roberts, et al., 1993). However, the model of activator recruitment of TFIIB to the promoter is controversial. Recruitment of TFIIB to the promoter was orginally hailed as the “rate limiting step” in PIC formation and thus a worthy target of transcriptional activators (Lin, etal., 1991). However, a study of VP16 indicated a requirement for more than interaction with TFIIB to obtain full activation 74 because mutations in VP16 that result in a decrease in activation potential do not show a lower affinity for TFIIB (Walker, et al., 1993). Furthermore, mutations in TFIIB that disrupt the stability of the TBP-TATA-TFIIB complex do not affect the response of these complexes to activation by VP16 (Chou and Struhl, 1997). Recent evidence indicates that TFIIB undergoes a conformational change in the presence of TBP and DNA (Neish, et al., 1998). It had earlier been noted that TFIIB undergoes a conformational change in response to VP16 (Roberts, 1994). Thus the role of VP16, and transcriptional activators in general, has been hypothesized as inducing a conformational change in TFIIB to facilitate the binding of TFIIB to the TBP-DNA complex. A species-specific region of yeast TFIIB, which extends from amino acid 144 to amino acid 157, has been identified (Shaw, et al., 1996). Deletion of this region of yTFIIB severely compromised the ability of Hap4 to activate transcription while having little affect on the activation potential of Hapl , GCN4, or Gal4. Thus TFIIB may be a target of the Hap4 activation domain, although it is formally possible that this TFIIB mutant is unable to maintain a critical protein-protein contact necessary at the promoter tested with some protein other than Hap4. TBP binding to transcriptional activators such as VP16, Spl, and Gal4 has been demonstrated in vitro (Emili, et al., 1994; Ingles, et al., 1991; Melcher and Johnston, 1995). Moreover, mutational analysis of TBP has identified TBP residues that disrupt activator function in viva (Amdt, et al., 1995; Kim, et al., 1994; Kim and Roeder, 1997; Lee and Struhl, 1995; Liljelund, et al., 1993). These findings have lead to several hypotheses as to the relevance of this interaction. Primary among these hypotheses is that TBP recruitment to the promoter is assisted by transcriptional activators. The fact 75 that TBP tethered to the promoter via a DNA-binding domain relieves the requirement for an activation domain seems to support the notion that TBP recruitment could be a mechanism employed by transcriptional activators (Chatterjee, 1995; Klages, 1995; Majello, et al., 1998; Xiao, et al., 1995). This result is consistent with the findings that TBP-TATA recognition may be the rate limiting step in transcription in vivo that can be enhanced by activators (Klein, 1994), and that overexpression of TBP in viva stimulated transcription from TATA containing reporter genes (Colgan and Manley, 1992). TBP can exist as a dimer in soultion but binds DNA as a monomer, raising the possibility that the function of certain transcriptional activators is to dissociate the TBP dimer, allowing TBP to interact with the TATA-element (Coleman and Pugh, 1997; Coleman, et al., 1995; Taggart and Pugh, 1996). A mutation that disrupts the transcriptional activation by VP16 was found on the DNA binding surface of TBP (Kim, et al., 1994). Four different clusters of hydrophobic residues within TBP that are essential for activator function were identified through the systematic mutation of residues predicted by the crystal structure to be surface exposed in human TBP (Bryant, et al., 1996). Lastly, it has been proposed that TBP undergoes a conformational change that affects its binding to TATA in response to transcriptional activators such as VP16 (Liljelund, et al., 1993). TFIIH interacts with the activation domains of both VP16 and p53 (Xiao, et al., 1994). Recently, a connection between several activation domains and the kinase activity of TFIH has been demonstrated. The RARa activation domain AF-l can bind to TFIIH and be phosphorylated by TFIH-I both in vitro and in viva (Rochette-Egly, et al., 1997). This phosphorylation is required for in viva activation by RARa. Likewise, p53 can be phosphorylated by the TFIIH kinase in vitro and this enhances the DNA binding activity 76 of p53 (Lu, et al., 1997). The HIV-1 transactivator Tat binds TFIIH both in vitro and in viva and stimulates the phosphorylation of the RNAPII CTD by TFIIH (Garcia Martinez, et al., 1997). Overall, these results indicate that the link between TFIIH and transcriptional activation domains may involve either protein modifying the action of the other. This chapter describes several genetic strategies designed to identify the target(s) of the Hap4 activation domain present within the 330—554 region of the protein. EXPERIMENTAL METHODS Yeast Strains The yeast strains used for the experiments detailed in this chapter are described in Table 4. Ethyl Methane Sulfonate Mutagenesis Strategies that employed ethyl methane sulfonate (EMS) mutagenesis were carried out on BCYOS, L40, AMR70, and JSY03 yeast strains. In all cases, yeast containing a lexA-Hap4 expressing plasmid were grown to a density of 2 x 107 cells/ml in four 10 ml cultures of appropriate drop-out media. Cells were spun down and washed twice with about 5 ml of sterile water each time. Cells were resuspended in 3 ml of 100 mM KHPO4 pH7.0 (Phosphate buffer). Phosphate buffer was made fresh by mixing 4.62 ml of 1M KH2P04 and 7.38 ml of 1M K2HP04 in a final volume of 120 ml. EMS was added to the four tubes in the following amounts: 0, 10, 15, and 20 ul. Cells were 77 incubated at 30°C for one hour with shaking. After one hour, 10 ml of 5% sodium thiosulfate (Na28203) freshly made and filter sterilized was added to neutralize the EMS. Cells were spun down and waste was disposed as hazardous waste. Cells were washed twice with about 5 ml of phosphate buffer and then cells were resuspended in 4 ml of phosphate buffer. Each culture was diluted by a factor of 105 and 200 pl of each was plated onto YPD. The amount of EMS that resulted in 50% mortality was determined. That amount of EMS was then used to treat a fresh 10 ml culture. Upon final resuspension in phosphate buffer, yeast were split up into as many aliquots of ~2 x 106 cells as possible and incubated in 2 ml of appropriate glucose containing media at 30°C for 4-5 hours with shaking. Cells were plated onto appropriate media containing FOA at a concentration of 2 g/liter. (NOTE: This mutagenesis protocol should this be performed in several replicates to avoid the “jackpot eflect”, which is the repeated isolation of a particular mutant farmed early in the process that expanded throughout the population. ) Mutant Isolation Using S-Fluoroorotic Acid Colonies that survived on S-fluoroorotic acid (FOA) plates (Guthrie and Fink, 1991) were replica plated onto X-gal plates (Rose, et al., 1990) using a nitrocellulose filter. Blue colonies were indicative of false positives and white colonies were harvested for further analysis. 78 Yeast Strain Production To obtain a yeast strain that contained the trpI A99 allele, as well as integrated URA3 and lacZ reporter genes with lexA binding sites upstream, BCYOS (obtained from Barak Cohen in Roger Brent’s lab) was mated with TAT7 (obtained from Eric Andrulis in Rolf Stemglanz’s lab) and sporulated (Rose, et al., 1990). BCYOS was the source of the integrated URA3 gene with upsdtream lexA binding sites whereas TAT7 was the source of the integrated lacZ gene with upstream lexA binding sites. J SYO2 resulted from one of the segregants of that mating. J SYO2 contained integrated URA3 and lacZ reporter genes with lexA binding sites upstream but not the trpI A99 allele. The trpI A99 allele was introduced using the plasmid pJL228. The plasmid pJL228 was first cut with XhoI in order to linearize it. The resulting linear piece of DNA that contained a URA3 gene was transformed into J SY02. Yeast that grew in the absence of uracil were assumed to be the result of integration of the linear DNA into the genome of J SY02. Yeast was then grown in the presence of FOA to promote the excision of the URA3 gene. The resulting strains, ahving the trpl A99 allele was named JSY03. High Copy Suppression of Hap4 Point Mutants L40 cells that contain mutant lexA-Hap4 expressing plasmids are sensitive to 3- AT as a result of an inability to promote high levels of expression of the HIS3 gene. The HIS3 gene product is inhibited by 3-AT, and the level of resistance to 3-AT is an indicator of the trans-activation potential of a given lexA-Hap4 359-476 mutant. We hypothesized that suppression of 3-AT sensitivity may result as a consequence of over- expression of a mechanistic target of Hap4 359—476. To identify yeast genes which, 79 when highly expressed, might suppress 3-AT sensitivity, these yeast strains were transformed with a LEU2 marked high copy S. cerevisiae genomic library (YEP13 based) obtained from American Type Culture Collection (ATCC #37323). The library has a titer of 2.0 x 1010 colony forming units per milliliter with l x 104 clones needed for complete coverage of the yeast genome. 150,000 colonies were harvested from plates and plasmid DNA was recovered using the Wizard® Plus Maxiprep DNA Purification System from Promega. L40 cells harboring mutant versions of lexA-Hap4 359-476 were transformed with the library DNA (Agatep, et al., 1998). L40 contains a HIS3 gene with lexA binding sites located upstream. Transformants were plated directly onto plates that contained amounts of 3-aminotriazole (3-AT) known to be toxic. Upon isolation of suppressors I will verify the linkage of the suppression phenotype to the high copy yeast library plasmid. Once this has been accomplished I will sequence the ends of the insert to identify which fragment of genomic DNA is present. Database searches will allow for the identification of the genes included on the fragment. The genes will be PCR amplified and sub-cloned into high copy vectors separately and tested for their ability to suppress the mutant phenotype. If there is a large number of genes to process, genes with known roles in gene expression will be tested first. 80 Strain Genotype Reference AMR70 his341200, lysZ-801 am trpI-901, leu2- (Hallenberg, et al., 1995) 3,112, URA3::(leanp)8 -lacZ MAT-at BC Y05 ura3, trpI, his3, leu2, lysZ B. Cohen, personal leu2::lexA( 6ap)-LE U2 communication lys2::lexA(8ap)-URA3 MAT-a DEY1053 met13, rho', MAT-a D. McNabb, personal communication F W294 his4-917}; lysZ-I 73R2, trpI A63, (Gansherafi‘, et al., 1995) leuZAI, ura3-52, spt3A202 MAT-a F W463 his4-91 7}; lys2-I 73R2, trpI A63, (Eisenmann, et al., 1994) leuZAI, ura3-52, spt8A302::LEU2 MAT-a F Y833 his3A200, lys2A202 trpIA63, leuZAI, (Winston, et al., 1995) ura3-52, MAT- a FY1093 his4-917}: lys2-1 73R2, "pl/.163, (Roberts and Winston, leuZAI, ura3-52, ade8, 1996) spt7A402::LE U2 MAT- a F Y I 106 his4-9I 7;; lys2-I 73R2, trpIA63, (Roberts and Winston, leuZAI, ura3-52, spt20A100::URA3, 1996) MAT-0t JSY02 trpI A63, ura3::(leanp)8-lacz, This work lys2::lexA(8ap)-URA3, leu2::lexA( 60p )-LE U2, his3delt0200, MAT-a: JSY03 ura3::lexA8ap-lacZ, lys2::lexA8ap- This work URA3, leu2-3,112, trp1A99, ade2, his3A200, MAT-a TAT7 his3A200, trp1-901, leu2-3,112, ade2, E. Andrulis, personal lys2-801 am, ura3::(leanp)8-lacz, communication LYS2::(leanp)4-HIS3, MAT- a L40 his3A200, trpI-901, leu2-3,112, ade2, (Hallenberg, et al., 1995) lys2-801am, URA3:.°(lem0p)8-lacZ, LYS2::(lexA0pfl-HIS3, MAT- a TABLE 4. Description of yeast strains. 81 RESULTS Hap4 Activation Domain Target Identification The basic strategy employed to identify the target(s) of the Hap4 activation domain resident within the 330-554 region of the protein was to monitor the activation potential of a lexA-Hap4 fusion in strains that contained reporter genes with lexA binding sites upstream. Yeast strains that harbored a mutation in a gene encoding a required target of the Hap4 activation domain would be unable to support activation by lexA-Hap4 and thus would produce a different phenotype than those yeast that were able to support activation by lexA-Hap4. Four different strategies were used with the goal of identifying targets of the Hap4 activation domain. I will describe the logic of each strategy. In the cases where the strategy failed, I will describe what we learned from the experience and how the knowledge was applied to the subsequent strategy. L40/AMR70 A screen was designed that took advantage of the integrated lacZ gene with lexA binding sites upstream, present in the isogenic yeast strains L40 and AMR70. Yeast unable to support activation by lexA-Hap4 (330-554) will stay white when grown in the presence of the chromogenic substrate X-gal. Such yeast were considered as potentially harboring mutations in some Hap4 specific target, collected, and further characterized. Carrying out the screen simultaneously in the two isogenic yeast strains allowed for the expedient arrangement of mutants into complementation groups. 82 Yeast mutants were obtained by treating 2x 108 cells of the yeast strains L40 and AMR70 with ethyl methane sulfonate (EMS) so that 50% mortality was achieved. Following mutagenesis and recovery, equal amounts of both mutant strains were transformed with low copy TRPI marked vectors encoding a fusion of the DNA-binding domain of the bacterial protein lexA to either Hap4(S), made up of the 330-476 region of Hap4, or Hap4(L), made up of the 330-554 region of Hap4. (The rationale behind using Hap4(S) is that it is known to be a more potent trans-activator than Hap4(L). Hap4(L) may therefore be a more sensitive indicator of mutation.) Transformants were plated on plates that lacked tryptophan but included the chromogenic substrate X-gal. Failure of the lexA-Hap4 fusions to drive an integrated lacZ reporter gene should result in white colonies when the yeast are grown in the presence of X-gal. Therefore white colonies among blue colonies were considered mutants. White colonies were picked and plated on plates lacking tryptophan. The phenotype was confirmed by replating on X-gal plates. Because the L40 yeast strain also contained an integrated HIS3 with upstream lexA binding sites I was able to test any putative mutants for their resistance to 3-aminotriazole (3-AT). The putative mutants were subjected to increasing amounts of 3-AT (OmM, 2.5mM, SmM, 25mM, and 50mM). Sixty-nine putative mutants were processed, thirty-five of which showed no sensitivity to 3-AT and were thus discarded. The remaining thirty—four putative mutants were processed, but were difficult to handle. All of the remaining putative mutants showed dramatically reduced transformation efficiencies (greater that 100 fold), and manipulations that involved plasmid transformation were therefore difficult. However, the biggest problem faced was that the mutant phenotypes would disappear. Yeast taken 83 from frozen stocks of the supposed mutant would not display the mutant phenotype. Thus, it was the instability of the mutant phenotypes that resulted in abandonment in favor of a different strategy. BCYOS The previous strategy was a screen, I reasoned that a strategy based on selection would allow for many more yeast to be processed. Thus a new strategy was developed based on the activity of the URA3 product which converts 5-fluoroorotic acid (FOA) into fluorouracil. Fluorouracil is a competitive inhibitor of thymidilate synthase and thus toxic to yeast. The strain used in this strategy was BCY05 (from Barak Cohen) which has an integrated copy of URA3 with lexA binding sites upstream. Yeast will be able to grow in the presence of FOA if activation by lexA-Hap4 is disrupted. Yeast capable of growing in the presence of FOA were considered as potentially harboring mutations in some Hap4 specific target, collected, and further characterized. 2x 108 cells of BCY05 were treated with EMS so that 50% mortality was achieved. The surviving yeast were transformed with a low copy TRPI marked plasmid that carried a lexA-Hap4 (330-554) fusion. Transformants were plated on media that lacked tryptophan and included 0.1% FOA. Survivors were collected and plated on lactate media and tested for their ability to grow at 30°C. (I reasoned that if the putative mutants were indeed defective for Hap4 activation they should not be able to survive on lactate plates.) Those that did grow at 30°C were tested for their ability to grow at either 16°C or 37°C. Petite mutants, which failed to grow on lactate, were identified by crossing to DEY1053 and not considered further. (DEY1053 is a mo strain and thus is 84 unable to compliment the inability of other rho' strains to grow on lactate.) BCY05 also contains a LE U2 gene with upstream lexA binding sites. Putative mutants were also tested for their ability to grow on media lacking leucine as a further criterion. Of the 167 putative mutants that survived FOA, 16 failed to grow on lactate medium at 30°C. Of the remaining putative mutants, 6 failed to grow on lactate medium at 16°C and were considered cold sensitive (cs) while 17 failed to grow on lactate medium at 37°C and were considered temperature sensitive (ts). I chose to first concentrate on those putative mutants that were unable to grow on lactate media at 30°C because they displayed the most severe phenotype and I thought that these contained mutations that would ultimately be the easiest to characterize. I attempted to clear these mutants of the TRPI marked plasmid carrying lexA-Hap4 by passaging them 5-10 times in rich media (YPD). In many cases I was unable to revert to tryptophan auxatrophy by clearing the plasmid because the plasmid was already cleared. The explanation for this finding became apparent when I realized that the trpI allele present in BCY05 is a point mutant and thus prone to reversion. The selection pressure against activation by lexA-Hap4 in the presence of FOA lead to a loss of the plasmid expressing it. It would follow that the yeast in which this happened would now be trp’ because the plasmid contained the only functional TRPI gene in the cell. However, some yeast reverted from up to TRP“. These yeast were able to survive in the presence of FOA because they did not harbor the plasmid that expressed lexA-Hap4, and they were able to survive in the absence of tryptophan because they now contained a functional TRPI gene. In those few cases where it was possible to clear the TRPI marked plasmid carrying lexA-Hap4 from the putative mutants, upon retransfomation of the TRPI marked 85 plasmid carrying lexA-Hap4 no mutant phenotype existed. I reasoned that plasmid based mutations were the cause of the previously observed mutant phenotypes, even though the plasmids in question were not subjected to EMS. JSY03 The critical flaw in the previous strategy was the fact that BCY05 is a tryptophan auxatroph as a result of a single point mutation in the TRPI gene, making this an unstable phenotype. I thus developed a new strategy in the hopes of identifying Hap4 activation domain targets that could provide an insight into mechanistic properties of the Hap4 activation domain. The new strategy remained based on the activity of the URA3 gene product, which converts 5-fluoroorotic acid (FOA) into fluorouracil. The strain used in this strategy, J SY03, has an integrated copy of URA3 with lexA binding sites upstream and, importantly also containing the trplA99 allele. Yeast will be able to grow in the presence of FOA if activation of the URA3 gene by lexA-Hap4 is disrupted. Yeast capable of growing in the presence of FOA will be considered as possibly harboring mutations in some Hap4 specific target, collected, and further characterized. There are two significant differences between the previous failed strategy and the current strategy. The first difference is that J SY03 is a tryptophan auxotroph as a result of the total deletion of the TRPI gene (i.e. trp1A99) instead of a point mutation in the TRPI gene as BCY05 is. The number of false positives that were simply TRPI revertants was too high in the previous strategy. The second change is that instead of transforming the lex-Hap4 expressing plasmid after EMS treatment, I mutagenized JSY03 cells that had already had the plasmid present. The resulting increase in the background of false 86 positives due to plasmid based mutations is balanced by the ability to increase the total number of yeast screened and ultimately selected. A minor difference is that J SY03 contains an integrated lacZ reporter gene with lexA binding sites upstream. This permits easy verification and quantification of the mutant phenotype. J SY03 cells were transformed with a low copy TRPI marked plasmid expressing the lexA-Hap4 (330-554) fusion protein. These cells were treated with EMS so that 50% mortality was achieved. Cells were split into aliquots of ~2x10° cells and revitalized in trp'lglucose medium for 5 hours. The aliquots were then plated onto 61 trp’l 0.2% FOA plates and grown at 30°C. 8000 colonies grew on the 61 plates. When plated on X-gal plates (to test for activation of the lacZ reporter gene), ~10% turned blue indicating that the ability to grow in the presence of FOA was likely the result of a mutation within the URA3 gene. Therefore these putative mutants were no longer considered. To further screen for mutations affecting activation by Hap4 we reasoned that if lexA-Hap4 (330-554) is unable to activate transcription then endogenous Hap4 should be unable to activate transcription. If endogenous HAP4 is unable to function as a transcriptional activator the yeast will be unable to grow on a non-fermentable carbon source such as lactate. Thus failure to grow on lactate is an expected mutant phenotype. I tested 6 colonies from each initial selection plate for their ability to grow on lactate. Of the 366 colonies tested 42 were unable to grow on lactate. These 42 were then cleared of the TRPI marked plasmid and retransformed with a freshly prepared version of the same plasmid to determine if the FOA resistant phenotype and the white on X-gal phenotype was the result of a defect in the plasmid. Freshly transformed cells were grown on up plates, colonies replica plated onto X—gal plates, and the blue/white phenotype assayed in 87 a qualitative manner. 13 of the 42 putative mutants seemed to turn blue slowly and 2 remained white. The ability of these 15 putative mutants to activate the transcription of a lexA-lacZ reporter gene was quantified. This test yielded 4 “mutants” defective for trancriptional activation by lexA-Hap4 (330—554). Tetrad analysis revealed that the lactate phenotype and the inability to activate transcription of the integrated lacZ reporter were not linked. Upon making this discovery I realized that I may have been tracking mutations in the reporter genes themselves and not mutations in targets of Hap4. In order to answer this question I transformed with a plasmid borne lacZ reporter and tested whether the putative mutant strains would support activation by lexA-Hap4. Indeed, the strains were able to support activation of the plasmid borne lacZ reporter. Thus I infer that the mutations I followed were in the reporter genes themselves. This could be determined experimentally by introducing a new integrated reporter gene and testing for its function. If the new reporter gene is functional then we can be confident that the original reporter gene was indeed mutant. If the new reporter gene is not functional, we would presume that the strain in question possesses a mutatibn that influences integrated reporter gene function in trans while having no effect on plasmid based reporter gene function. We reasoned that the cause for the elevated mutation rate may be the use of FOA. Growth on FOA results in nucleotide starvation, specifically thymidylate, which in turn would result in an increased mutation rate. I therefore repeated the process in the absence of EMS treatment and selected for spontaneously occurring FOA-resistant yeast with the hope that they contained mutations in mechanistic targets of the Hap4 330-554 region. J SY03 cells were transformed with a low copy TRPI marked plasmid expressing the 88 lexA-Hap4 (330-554) fusion protein. I plated ~2x107 of these cells onto a total of 20 trp’l 0.2% FOA plates and grew them at 30°C. After three days at 30°C 20 FOA resistant colonies appeared. I transformed the FOA-resistant cells with a plasmid-bome lacZ reporter and tested for the putative mutant strains to support activation by lexA-Hap4. All but one of the putative mutant strains was able to support activation of the plasmid borne lacZ reporter and therefore were discarded. The ability of the single remaining mutant to support trans-activation by lexA—Hap4 is shown in Figure 14. The ability of lexA-Hap4 to activate the expression of a lacZ reporter gene was reduced lO-fold in the spontaneously occurring mutant. (The inflated numbers seen in this assay are likely due to the high copy reporter plasmid used.) To determine if the defect in trans-activation was specific to Hap4 or more general in nature, the ability of the putative mutant to support trans-activation by lexA- GNC4 was also assessed. Figure 15 shows that neither lexA-Hap4 nor lexA-GCN4 was able to activate the expression of the integrated lacZ reporter gene. (The ~10 fold difference between the data seen in Figure 14 and that seen in Figure 15 with regard to the trans-activation potential of lexA-Hap4 is likely due to the increased number of lacZ reporter genes present in Figure 14.) While it may be premature to say that the defect is general in nature, this preliminary evidence indicates that the mutation is not in a factor specific to Hap4. Further characterization of this mutant yeast strain is underway. High Copy Suppression 89 High copy suppression is a strategy whereby the mutant phenotypes that are the result of an enfeebled activator protein are suppressed as a result of an increased amount 90 15000 .4. S 8 12000 H E 9000 - i 2 6000 - ti 3 3000 - o . STRAIN: WT Mutant ACTIVATOR: lexA-HAP4 lexA-HAP4 FIGURE 14. The ability of a spontaneously occurring JSY03 mutant to sup- port trans-activation by lexA-Hap4 (aa330-554). Hap4(aa330—554) was fused to lexA and assayed for the ability to activate expression of a lacZ gene inte- grated into the genome of the yeast strain that is the derivative of JSY03 as well as a lacZ gene present on a high copy LE U2 marked plasmid. Both lacZ genes had eight lexA binding sites upstream. Bars indicate mean activity (with standard deviation) of B-galactosidase in yeast cell abstracts from three parallel cultures. B—galactosidase activities in extracts containing only lexA were negligible. B-galactosidase units were measured as nmoles of c " l L ' ’...L....t..’...g protein. 91 2000 1500 - 1000 - Beta-Galactosidase Units 500 - o I I STRAIN: Wl' Mutant WT Mutant ACTIVATOR: lexA-Hap4 lexA-Hap4 lexA-GCN4 lexA-GCN4 FIGURE 15. The ability of a spontaneously occurring JSY03 mutant to sup- port trans-activation by lexA-Hap4 or lexA-GCN4. Either Hap4(aa330-554) or GCN4 (aa9-l72) was fused to lexA and assayed for the ability to activate expression of a lacZ gene with eight lexA binding sites upstream integrated into the genome of the yeast strain that is a derivative of JSY03. Bars indi- cate mean activity (with standard deviation) of B-galactosidase in yeast cell abstracts from three parallel cultures. [3- -galactosidase activities in extracts containing only lexA were negligible. B- -galactosidase units were measured as nmoles of n " I " ....... t-.’...g protein. 92 of target(s). The Hap4 point mutations that result in the loss of activation potential may diminish the ability of Hap4 to maintain the protein-protein contacts needed to activate transcription. The binding equilibrium may be shifted to the point where the mutant protein is not engaged with its target long enough to result in trans-activation. The presence of an increased amount of target protein may shift the equilibrium back to a range that is sufficient for trans-activation. Thus taking advantage of the point mutants described in Chapter 2, I have developed, but not completed, a new strategy that is based on high copy suppression with the goal of identifying Hap4 activation domain target(s) that could provide an insight into mechanistic properties of the Hap4 activation domain. One advantage of this strategy is that it allows for quick identification of the gene responsible for the suppression. In this strategy I will employ the L40 yeast strain described earlier. Taking advantage of the integrated HIS3 gene with upstream lexA binding sites, I will titrate the levels of 3-AT to which the various Hap4 point mutants are resistant. I will then transform a high copy yeast library and select for suppression of the 3-AT sensitivity previously demonstrated. Upon isolation of suppressors I will verify the linkage of the suppression phenotype to the high copy yeast library plasmid. Once this has been accomplished I will sequence the ends of the insert to identify which fragment of genomic DNA is present. Database searches will allow for the identification of the genes included on the fragment. The genes will be PCR amplified and sub-cloned into high copy vectors separately and tested for their ability to suppress the mutant phenotype. If there is a large number of genes to process, genes with known roles in gene expression 93 will be tested first. For a discussion of experimental plans for establishing the existence of specific protein-protein interactions see the Discussion section of Chapter 2. SPT3I'7/8/20 The genetic screens described to this point represent one strategy for identifying putative targets of the Hap4 activation domain. An alternative strategy is to test for genetic interactions between Hap4 and genes already implicated in co-activator mechanisms from other studies. We therefore tested the ability of Hap4 330-554 to activate transcription in yeast strains that lacked SPT3, SPT7, SPT8, or SP720. These SPT genes are part of the SAGA complex and are known transcriptional co-factors. Hap4 (aa330-554) is able to function independently of many of the SAGA complex members, most notably Ada2, Ada3, and Gcn5. However, the ability of Hap4 (aa330- 554) to function independently of some of the SAGA complex members that does not mean that Hap4 (aa330-554) is not dependent on other members of the complex and thus on the integrity of the complex as a whole. I tested the ability of lexA-Hap4 330-554 to activate the transcription in spt3, spt7, spt8, or spt20 deletion mutant strains. The lacZ reporter gene with lexA binding sites positioned upstream was located on a LE U2 marked plasmid with a 2n origin. The results of this test can be seen in Figure 16. The deletion of SPT3 or SPT8 had no negative effect on the trans-activation potential of Hap4 330-554. In contrast, deletion of SP17 or SP720 resulted in a total ablation of the trans-activation potential of lexA-Hap4 330-554. Irnmunoblots seen in Figure 17 show that deletion of either SPT? or SPT20 did 94 not result in a decreased amount of lexA-Hap4 protein. Therefore, the SPT7 and SPT20 gene products represent putative mechanistic targets of Hap4 (aa330-554). 95 10000 - Beta-Galactosidase Units spt3 spt7 spt8 spt20 WT FIGURE 16. The effect of deletion of either SPT3, SP77, SPT8, or SPT20 on the ability lexA-Hap4(aa330-554) to activate transcription. Hap4(aa330-554) was fused to lexA and assayed for the ability to activate expression of a lacZ gene present on a high copy LEU2 marked plasmid both with eight lexA binding sites upstream in yeast strains lacking either .SPT3, SP77, SPT8, or SPT20 Bars indicate mean activity (with standard deviation) of B—galactosidase in yeast cell abstracts from three parallel cultures. B-galactosidase activities in extracts containing only lexA were negligible. B-galactosidase units were measured as nmoles of o-nitrophenol/minute/mg protein. 96 Wild Type Aspt7 Aspt20 FIGURE 17. Immunoblot analysis of HAP4 (aa330-476) fused to lexA in either wild type, Aspt7, or Aspt20 yeast cells. HAP4 (aa330-476) fused to lexA was expressed from the ADHl promoter on a URA3 marked 21.1 plasmid. Aliquots of whole cell extracts were electrophoresed in SDS polyacrylamide gels and blotted to nitrocellulose. The filters were incubated with a mixture of three different monoclonal antibodies directed against the lexA protein and developed using enhanced chemiluminescence. 97 DISCUSSION While many lessons were learned along the way, the fatal flaw in all of the selection strategies, with the possible exception of the spontaneously occurring mutant strategy, was that I was using a potent activator and an integrated URA3 with 8 binding sites for that activator upstream. The combination of FOA and 8 lexA-Hap4 molecules bound upstream of URA3 proved too toxic to the yeast - not likely to be reversed by a simple point mutation in a Hap4 activation domain target. Perhaps one lexA-binding site upstream of the URA3 gene would have been better because it would have reduced the synergistic effect of having 16 activation domains present upstream. In any case, for such a strategy to be successful one needs a system where a modest change in activity results in a discemable phenotypic change. The use of growth on lactate as a secondary screen may not have been appropriate. The logic held that if the strain in question harbored a mutation in a target of the Hap4 330-554 activation domain that rendered it non-functional, then this mutation should cause the endogenous Hap4 to also be non-functional and thus unable to grow on lactate containing medium. However, the selection was set up using only one of the two activation domains resident within Hap4. It is possible that the target of the 330-554 region of Hap4 is different from the target of the 1-330 region of Hap4. If this is the case then it is possible that yeast harboring a mutation in a target of the 330-554 region of Hap4 would still be able to grow on lactate as a result of a functional activation domain present within the 1-330 region of Hap4. Lastly, all of the strategies used, with the exception of the high copy suppression strategy, were based on trying to break something; namely the link between the Hap4 98 activation domain and its mechanistic target. These strategies thus aimed for going from a “working” situation to a “broken” situation and selecting for that. I think that going from “broken” to “working” is a much better way to go and is the aim of the high copy suppression strategy. There are many ways to break something, but when dealing with a specific deficiency there are few ways to remedy the situation. (Note: Going from “working” to “broken” is a problem when using selection strategies, when using screening strategies the “working” to “broken” method is tried and true.) Despite these problems, I do have one possible mutant of interest that is a product of a “working” to “broken” strategy and am currently attempting to characterize it. Among the things that I need to know are whether the defect is the result of a mutation within a single gene, whether or not there are any growth phenotypes associated with this mutation, and whether or not the mutation is recessive or dominant. If the mutation is recessive, in one gene, and has a growth defect associated with it, I will then be able to clone it by complementation. The finding that lexA-Hap4 (aa330-554) is dependent on the SPT7 and SPT20 gene products is very interesting. In the absence of the SPT7 and SPTZO gene products the SAGA complex is disrupted (Sterner, et al., 1999). Hap4 (aa330-554) is able to function independently of many of the SAGA complex members, most notably Ada2, Ada3, and GCN5. However, the ability of Hap4 (aa330-554) to function independently of some of the SAGA complex members does not mean that Hap4 (aa330-554) is not dependent on other members of the complex and thus an the integrity of the complex as a whole. The SAGA complex has functions other than those provided by the ADA gene family, such as SPT 3 and SPT8 interaction with TBP, and perhaps this is a functional 99 hint at that point. It is possible that SPT7 and SPT 20 are either direct targets of Hap4 (aa330-554) or part of another complex that is part of the pathway used by Hap4 (aa330- 554) to activate transcription. However, it should be noted that the effects of SPT7 and SPT20 are very general. No activator tested to date is able to work in the absence of SPT7 and SPT20, a list that includes Gal4, GCN4, and Hap4 tested from a natural promoter (D. McNabb, personal communication). Perhaps SPT7 and SPT20 are general factors needed for all trans-activation. 100 W Future Directions and Persmctives I have established that Hap4 contains two independent activation regions, each of which depend on hydrophobic amino acids to function as a transcriptional activator. The two regions differ in their dependence on GCN5, thus implying that they employ distinct and different mechanisms in the activation of transcription. My attempts to identify a mechanistic target of the Hap4C region have not been as successful. However, I will detail future experiments that may result in a successful identification of a mechanistic target of the Hap4 359-476 region. I will also attempt to put my work into the larger context of studies of activation domains and mechanisms. The results presented in Chapter 2 shows that clusters of hydrophobic amino acids make contributions to the activation potential of Hap4. These clusters contain anywhere from two to four residues. Whether or not all of the residues within an identified cluster make equal contributions to the overall activation potential of Hap4 has not been established. In order to determine the contribution of each residue identified they would have to be changed individually to serine or, in the case where more than two residues exist within a cluster, combinatorial mutants would need to be constructed. I chose to change the hydrophobic residues within the identified clusters to serine because serine is both small and non-hydrophobic as opposed to alanine, which, while it is small, possesses some hydrophobic character. The question of whether or not alanine substitutions at the identified positions would result in an equivalent loss of trans- activation potential would be useful information. If changes to alanine result in similar, 10] yet less severe, phenotypes, then the hydrophobic character of alanine would be useful in the high copy suppression strategy. If it is true that hydrophobic character is necessary for activator-target interactions, then alanine would be able to contribute in this way. In Chapter 3 I described my attempts to identify the target of the Hap4 C-region. A similar approach can be used with the Hap4 N-region to answer the question of whether or not these spatially distinct activation domains have different targets. With the high copy suppression strategy we may identify putative targets of Hap4. These targets may be known transcription factors or be of have a yet unidentified function. In the event that we are able to identify previously undiscovered genes that encode Hap4 activation domain targets, we would construct strains in which the wild- type gene is disrupted. A large deletion would be constructed in the plasmid-bome gene, and a selectable marker (such as URA3 gene) would be inserted. This cassette would be transformed into a diploid strain, selecting for integration of the URA3 marker gene. The presumably heterozygous diploids would be sporulated and tetrads dissected, where a 2:2 ratio of viability would indicate that the gene is essential. If the gene is non-essential, the different growth phenotypes that are associated with deletion of the gene would be catalogued. If the identified gene is non-essential, the efficiency of activation by Hap4wt, lexA-Hap4N, and lexA-Hap4C in the deletion strain could be tested in order to determine whether both activation regions of Hap4 are non-functional in the deletion strain. The determination of whether or not the identified target is Hap4 specific or is required for the function of multiple activation domains could be determined by assaying the ability of 102 lexA-Hap4, Gal4, and GCN4 fusion proteins to activate a lacZ gene with upstream lexA sites. High copy suppression may indicate a genetic interaction between the activation domain tested and the target identified. To determine if the two identified players physically interact, other tests must be carried out. If an epitope-tagged target protein is able to interact with lexA-Hap4, then the lexA monoclonal antibodies that we possess could be used to verify the physical interaction by immunoprecipitation experiments. Quantitative gel shifts or Bia-Core experiments could be used to determine the affinity of the identified target for Hap4. If no direct physical interaction between Hap4 and the genetically identified target is demonstrated, then the exact role of this factor will be an open question, one that will need to be answered. An epitope-tagged target protein would also be useful in determining if the newly identified factor is a part of any larger complexes and what these complexes are. It is possible that it may be part of an already identified complex such as SAGA or SWIISNF or it may be part of a yet identified complex. Conventional chromatography methods such as ion-exchange and size-filtration columns could be used to crudely purify the complex(s) that contain this factor if the factor is part of a previously unidentified complex. Affinity chromatography using the epitope tag would allow further purification of the complex(s). The implication of this work on the field as a whole rests largely on whether or not we are able to establish a target for either of the activation domains resident within Hap4. The fact that these activation domains are dependent on hydrophobic amino acids is a new finding for Hap4, but a well-established feature of many other transcriptional 103 activators. The use of point mutants within Hap4 for the high copy suppression strategy is a novel approach for establishing activation domain targets and allows for immediate biochemical verification of the phenomenon through the establishment of physical interactions if a direct interaction occurs. This links the phenotype to the activation domain itself through the physical interaction data. Whether or not new factors are identified in the search described above or a previously identified factor is implicated in the Hap4 pathway, the result will add another layer of complexity onto the ever-expanding complex universe that is the field of gene expression. Reductionism is giving way to more holistic approaches for attacking the unsolved questions that remain in the field of gene expression; approches that are carried out in natural contexts. The work of the reductionist was valuable, for the in vitro work identified the basal transcription factors necessary for promoter-directed transcription. However, the discovery of activated transcription quickly introduced a new level of complexity into the picture and things have been getting more complicated ever since. For instance, detailed structural analysis of activators has proven very difficult. Simple patterns of residues do not seem to hold the key to how activators are differentiated. It will be necessary to study activators in the presence of targets in a natural context to gain some insight into how specificity is achieved. I hope that my high copy suppression strategy is a step in the right direction. An example of how interconnections may be missed if one is to focus only on a single track can be found in the study of co-activators and co-repressors. Co-activators and co—repressors provide for the integration of signals from diverse signal transduction 104 pathways and translate them into coordinated transcriptional responses. The role of co- activators and co-repressors in the response of the steroid receptors to hormones provides an example of how integration is accomplished in nature. It is also a good example of how complicated the picture can be. For instance, the co-activator GRIPl modulates the activity of the progesterone, estrogen, retinoic acid, and retinoid X receptors (Hong, et al., 1996; Voegel, et al., 1996) while SRC-l modulates the activity of a partially overlapping group that includes the progesterone, estrogen, retinoic acid, retinoid X, glucocorticoid, and thyroid receptors (Kamei, et al., 1996; Onate, et al., 1995; Yao, et al., 1996). Each of these steroid receptors is likely to control the expression of a different regulon, so the fact that they all interact with two common factors does not simply represent a point of convergence. The controlled expression of the co-activator targets provides a means of regulating the action of various receptors. Another interesting possibility is that the ligand-receptor interaction may influence the affinity of the receptor for a co-activator, thus providing another level of control. The challenge that remains for those in the field of gene expression is how to make sense of all of the data; to integrate it, if possible. It is no longer possible to study just one thing to the exclusion of all else - be it one activator, one complex, or even one pathway. One must now fit data into the constantly changing big picture. Genetic techniques will be invaluable in this pursuit. The in viva validation of the in vitro work is a complicated, but necessary, step in our understanding of gene expression. The most recent approaches to comprehensively address these issues are the related fields of genomics, which attempts to establish circuits of gene expression, and proteomics, which attempts establish protein families based on structure or sequence 105 similarity. Studies using these techniques may provide the foundation and context for interpreting mechanistic studies in control of gene expression. However, as with any new avenue of investigation there are pitfalls. Genome-wide analysis will yield vast amounts of data, but how to interpret it will be the challenge. For instance, examples of large fluctuations in specific mRNA levels can occur as a result of a given treatment without a corresponding alteration in protein levels. Are the fluctuations in mRNA levels significant in this case? Despite such issues, these techniques will allow for a broader perspective and will allow for a better understanding of the regulatory circuits that exist within the genome. The variety of genomic sequencing initiatives that make genome-wide analysis possible also allows for cross-species awareness. Homologues of a given protein can be identified in other species, the extent to which it has been conserved can be determined, and an understanding and appreciation for what they might be doing in these other species may be attained. One can find out if they are involved in similar pathways and what context they are found in. This may then influence the nature of the questions asked in other organisms. For instance, if I am able to identify a mechanistic target of Hap4 in S. cerevisiae, does it exist in K. lactis and does the Hap4 homologue in K. lactis target it as well? This may lead to an identification of conserved transcriptional regulatory circuits in K. lactis. If the landscape is different in K. lactis, it may lead to the discovery of new transcriptional regulatory circuits. 106 BIBLIOGRAPHY 107 10. ll. 12. l3. l4. Abate, C., D. Luk, E. Gagne, R. G. Roeder, and T. Curran. 1990. Fos and jun cooperate in transcriptional regulation via heterologous activation domains. Mol. Cell. Biol. 10:5532-5535. Agatep, R., R. D. Kirkpatrick, D. L. Parchaliuk, R. A. Woods, and R. D. Gietz. 1998. Transformation of Saccharomyces cerevisiae by the lithium acetate/single- stranded carrier DNA/polyethylene glycol (LiAc/ss-DNA/PEG) protocol. Technical Tips Online(http://tto.trends.com) . Akoulitchev, S., T. P. Makela, R. A. Weinberg, and D. Reinberg. 1995. Requirement for TFIIH kinase activity in transcription by RNA polymerase 11. Nature 377 :557-560. Alexeev, D. G., A. A. Lipanov, and I. Skuratovskii. 1987. Poly(dA).poly(dT) is a B-type double helix with a distinctively narrow minor groove. Nature 325:821-3. Allfrey, V. G., R. Faulkner, and A. E. Mirsky. 1964. Acetylation and methylation of histones and their possible role in reulation of RNA synthesis. Proc. Natl. Acad. Sci. 51:786-794. Almlof, T., J. A. Gustafsson, and A. P. H. Wright. 1997. Role of hydrophobic amino acid clusters in the transactivation activity of human glucocorticoid receptor. Mol. Cell. Biol. 17:934-945. Aparicio, O. M., B. L. Billington, and D. E. Gottschling. 1991. Modifiers of position effect are shared between telomeric and silent mating—type loci in S. cerevisiae. Cell 66: 1279-87. Apone, L. M., C. A. Virbasius, F. C. HolstegC, J. Wang, R. A. Young, and M. R. Green. 1998. Broad, but not universal, transcriptional requirement for yTAFIIl7, a histone H3-like TAFII present in TFIID and SAGA. Mol Cell 2:653-61. Apone, L. M., C.-M. A. Virbasius, J. C. Reese, and M. R. Green. 1996. Yeast TAFII90 is required for cell—cycle progression through GZ/M but not for general transcription activation. Genes Dev. 10:2368-2380. apRhys, C. M. J., D. M. Ciufo, E. A. O’Neill, T. J. Kelly, and G. S. Hayward. 1989. Overlapping octamer and TAATGARAT motifs in the VF65-response elements in herpes simplex virus immediate-early promoters represent independent binding sites for cellular nuclear factor 111. J. Virol. 63:2798-2812. Archambault, J., F. Lacroute, A. Ruet, and J. D. Friesen. 1992. Genetic interaction between transcription elongation factor TFIIS and RNA polymerase 11. Mol. Cell. Biol 12:4142-4152. Amdt, K. M., S. Ricupero-Hovasse, and F. Winston. 1995. TBP mutants defective in activated transcription in vivo. EMBO J. 14:1490-1497. Amdt, K. T., C. A. Styles, and G. R. Fink. 1989. A suppressor of a HIS4 transcriptional defect encodes a protein with homology to the catalytic subunit of protein phosphatases. Cell 56:527-37. Asa, T., J. W. Conaway, and R. C. Conaway. 1994. Role of core promoter structure in assembly of the RNA polymerase H preinitiation complex. A common pathway for formation of preinitiation intermediates at many TATA and TATA-less promoters. J Biol Chem 269:26575-83. 108 15. 16. 17. l8. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. Attardi, L. D., and R. Tjian. 1993. Drosophila tissue-specific transcription factor NTF-l contains a novel isoleucine-rich activation motif. Genes Dev. 7: 1341-1353. Auble, D. T., Hansen, K. E., Mueller, C. G. F., Lane, W. S., Thorner, J ., Hahn, S. 1994. Motl , a global repressor of RNA polymerase II transcription, inhibits TBP binding to DNA by an ATP-dependent mechanism. Genes Dev. 8: 1920-1934. Barberis, A., C. W. Muller, S. C. Harrison, and M. Ptashne. 1993. Delineation of 2 functional regions of transcription factor-TFIIB. Proc. Natl. Acad. Sci. USA 90:5628-5632. Barlev, N. A., R. Candau, L. Wang, P. Darpino, N. Silverman, and S. B. Berger. 1995. Characterization of physical interactions of the putative transcriptional adaptor, ADA2, with acidic activation domains and TATA-binding protein. J. Biol. Chem. 270: 19337-19344. Bengal, E., O. Flores, A. Krauskopf, D. Reinberg, and Y. Aloni. 1991. Role of the mammalian transcription factor-11F, factor-HS, and factor-11X during elongation by RNA polymerase-II. Mol. Cell. Bio. 11:1195-1206. Berger, 8. L., B. Pina, N. Silverman, G. A. Marcus, J. Agapite, J. L. Regier, S. J. Triezenberg, and L. Guarente. 1992. Genetic isolation of ADA2 - a potential transcriptional adaptor required for function of certain acidic activation domains. Cell 70:25 1-265. Berroteran, R. W., D. E. Ware, and M. Hampsey. 1994. The sua8 supressors of Saccharomyces cerevisiae encode replacements of conserved residues within the largest subunit of RNA polymerase II and affect transcription start site selection similarly to sua7 (T FIIB) mutations. Mol. Cell. Biol. 14:226-237. Bjorklund, S., and Y. J. Kim. 1996. Mediator of transcriptional regulation [see comments]. Trends Biochem Sci 21:335-7. Blair, W. S., H. P. Bogerd, S. J. Madore, and B. R. Cullen. 1994. Mutational analysis of the transcription activation domain of RelA: Identification of a highly synergistic minimal acidic activation module. Mol. Cell. Biol. 14:7226-7234. Blommers, M. J. J ., M. J. J. Fendrich, C. Garcia-Echeverria, and P. Chene. 1997. On the interaction between p53 and MDM2: Transfer NOE study of a p53- derived peptide ligated to MDM2. J. Amer. Chem. Soc. 119:3425-3426. Bonne Andrea, C., F. Harper, J. Sobczak, and A. M. De Recondo. 1984. The role of HMG] protein in nucleosome assembly and in chromatin replication. Adv Exp Med Biol 179:479-88. Bourgarel, D., and C. B.-F. Nguyen, M. 1999. HAP4, the glucose-repressed regulated subunit of the HAP transcriptional complex involved in the ferrnentation- respiration shift, has a functional homologue in the respiratory yeast Kluyveramyces lactis. Molecular Miocrobiology 31: 1205-1215. Bradford, M. M. 1976. A rapid and sensitive method for the quantitation of microgram quantities of protein using the principle of protein-dye binding. Anal. Biochem. 72:248-254. Bradsher, J. N., S. Tan, H. J. McLaury, J. W. Conaway, and R. C. Conaway. 1993. RNA polymerase H transcription factor SIII. 11. Functional properties and role in RNA chain elongation. J Biol Chem 268:25594-603. 109 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 45. Brewster, N. K., G. C. Johnston, and R. A. Singer. 1998. Characterization of the CP complex, an abundant dimer of Cdc68 and Pob3 proteins that regulates yeast transcriptional activation and chromatin repression. J Biol Chem 273:21972-9. Brownell, J. E., and C. D. Allis. 1996. Special HATs for special occasions; linking histone acetylation to chromatin assembly and gene activation. Curr. Opin. Genetics Dev. 6: 176-184. Brownell, J. E., J. Zhou, T. Ranalli, R. Kobayashi, D. G. Edmondson, S. Y. Roth, and C. D. Allis. 1996. Tetrahymena histone acetyltransferase A: A homolog to yeast Gcn5p linking histone acetylation to gene activation. Cell 84:843-851. Bryant, G. 0., L. S. Martel, S. K. Burley, and A. J. Berk. 1996. Radical mutations reveal TATA-box binding protein surfaces required for activated transcription in vivo. Genes Dev. 10:2491-2504. Buermeyer, A. B., L. A. Strasheim, S. L. McMahon, and P. J. Farnharn. 1995. Identification of cis-acting elements that can obviate a requirement for the C-terrninal domain of RNA polymerase II. J Biol Chem 270:6798-807. Buratowski, S., S. Hahn, L. Guarente, and P. A. Sharp. 1989. Five intermediate complexes in transcription initiation by RNA polymerase 11. Cell 56:549-561. Buratowski, S., S. Hahn, P. A. Sharp, and L. Guarente. 1988. Function of a yeast TATA element-binding protein in a mammalian transcription system. Nature 334:37- 42. Buratowski, S., and H. Zhou. 1993. Functional domains of transcription factor- TFIIB. Proc. Natl. Acad. Sci. USA 90:5633-5637. Burley, S. K., and R. G. Roeder. 1996. Biochemistry and structural biology of transcription factor IID (TFIID). Annu Rev Biochem 65:769-99. Burns, L. G., and C. L. Peterson. 1997. Protein complexes for remodeling chromatin. Biochim Biophys Acta 1350: 159-68. Butler, G., and D. J. Thiele. 1991. ACE2, an activator of yeast metallothionein expression which is homologous to SW15. Mol. Cell. Biol. 11:476-485. Cairns, B. R., N. L. Henry, and R. D. Kornberg. 1996. TFG/l‘ AF30/ANC1, a component of the yeast SWIISNF complex that is similar to the leukemogenic proteins ENL and AF-9. Mol Cell Biol 16:3308-16. Cairns, B. R., Y. J. Kim, M. H. Sayre, B. C. Laurent, and R. D. Kornberg. 1994. A multisubunit complex containing the SWIl/ADR6, SWIZ/SNFZ, SW13, SNFS, and SNF6 gene products isolated from yeast. Proc Natl Acad Sci U S A 91: 1950-4. Cairns, B. R., Y. Larch, Y. Li, M. Zhang, L. Lacomis, H. Erdjument-Bromage, P. Tempst, J. Du, B. Laurent, and R. D. Kornberg. 1996. RSC, an essential abundant chromatin-remodeling complex. Cell 87: 1249- l 260. Cao, Y., B. R. Cairns, R. D. Kornberg, and B. C. Laurent. 1997. Sflrl p, a component of a novel chromatin-remodeling complex, is required for cell cycle progression. Mol. Cell. Biol. 17:3323-3334. Carcamo, J., L. Buckbinder, and D. Reinberg. 1991. The initiator directs the assembly of a transcription factor-Hd-dependent transcription complex. Proc. Natl. Acad. Sci. USA 88:8052-8056. Cavallini, B., 1. Fans, H. Matthes, J. M. Chipoulet, B. Winsor, J. M. Egly, and P. Chambon. 1989. Cloning of the gene encoding the yeast protein BTFlY, which can 110 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. substitute for the human TATA box-binding factor. Proc. Natl. Acad. Sci. USA 86:9803-9807. Chambers, R. S., and M. E. Dahmus. 1994. Purification and characterization of a phosphatase from HeLa cells which dephosphorylates the C-terminal domain of RNA polymerase II. J Biol Chem 269:26243-8. Chambers, R. S., B. Q. Wang, Z. F. Burton, and M. E. Dahmus. 1995. The activity of COOH-terrninal domain phosphatase is regulated by a docking site on RNA polymerase II and by the general transcription factors IIF and 11B. J Biol Chem 270: 14962-9. Chang, M., and J. A. Jaehning. 1997. A multiplicity of mediators: alternative forms of transcription complexes communicate with transcriptional regulators. Nucleic Acids Res 25:4861-5. Chatterjee, S. a. S., K. 1995. Connecting a promoter-bound protein to TBP bypasses the need for a transcriptional activation domain. Nature 347 :820-2. Chen, W., and K. Struhl. 1988. Saturation mutagenesis of a yeast hisB "TATA element": genetic evidence for a specific TATA-binding protein. Proc Natl Acad Sci U S A 85:2691-5. Chi, T., and M. Carey. 1996. Assembly of the isomerized TFIIA--TFIID--TATA ternary complex is necessary and sufficient for gene activation. Genes Dev 10:2540- 50. Choder, M., and R. A. Young. 1993. A portion of RNA polymerase II molecules has a component essential for stress responses and stress survival. Mol Cell Biol 13:6984-91. Chodosh, L. A., A. S. Baldwin, R. W. Carthew, and P. A. Sharp. 1988. Human CCAAT-binding proteins have heterologous subunits. Cell 53:11-14. Chou, S., and K. Struhl. 1997. Transcriptional activation by TFIIB mutants that are severely impaired in interaction with promoter DNA and acidic activation domains. Mol. Cell. Biol. 17:6794-6802. Cisek, L. J ., and J. L. Corden. 1989. Phosphorylation of RNA polymerase by the murine homologue of the cell-cycle control protein cdc2. Nature 339:679-84. Clark Adams, C. D., and F. Winston. 1987. The SPT6 gene is essential for growth and is required for delta-mediated transcription in Saccharomyces cerevisiae [published erratum appears in Mol Cell Biol 1987 May;7(5):2035]. Mol Cell Biol 7 :679-86. Coleman, R. A., and B. F. Pugh. 1997. Slow dimer dissociation of the TATA binding protein dictates the kinetics of DNA binding. Proc. Natl. Acad. Sci. USA 94:7221-7226. Coleman, R. A., A. K. P. Taggart, L. R. Benjamin, and B. F. Pugh. 1995. Dimerization of the TATA binding protein. J. Biol. Chem. 270: 13842-13849. Colgan, J ., and J. L. Manley. 1992. TFIID can be rate limiting in vivo for TATA- containing, but not TATA-lacking, RNA polymerase II promoters. Genes Devel. 6:304-315. Conaway, R. C., and J. W. Conaway. 1993. General initiation factors for RNA polymerase 11. Ann. Rev. Biochem. 62: 161-190. 111 61. 62. 63. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. Conaway, R. C., and J. W. Conaway. 1989. An RNA polymerase H transcription factor has an associated DNA-dependent ATPase (dATPase) activity strongly stimulated by the TATA region of promoters. Proc Natl Acad Sci U S A 86:7356-60. Cortes, P., O. Flores, and D. Reinberg. 1992. Factors involved in specific transcription by mammalian RNA polymerase H: purification and analysis of transcription factor HA and identification of transcription factor HJ. Mol. Cell. Biol. 12:413-421. Cote, J., J. Quinn, J. L. Workman, and C. L. Peterson. 1994. Stimulation of GAL4 derivative binding to nucleosomal DNA by the yeast SWIISNF complex. Science 265:53-60. Coulombe, B., J. Li, and J. Greenblatt. 1994. Topological localization of the human transcription factors HA, HB, TATA box-binding protein, and RNA polymerase H-associated protein 30 on a class H promoter. J Biol Chem 269: 19962-7. Coustry, F., S. M. Maity, S. Sinha, and B. de Crombrugghe. 1996. The transcriptional activity of the CCAAT-binding factor CBF is mediated by two distinct activation domains, one in the CBF-B subunit and the other in the CBF-C subunit. J. Biol. Chem. 271: 14485-14491. Cress, W. D., and S. J. Triezenberg. 1991. Critical structural elements of the VP16 transcriptional activation domain. Science 251:87-90. Dahlman-Wright, K., H. Baumann, I. J. McEwan, T. Almof, A. P. H. Wright, J. A. Gustafsson, and T. Hard. 1995. Structural characterization of a minimal functional transactivation domain from the human glucocortoid receptor. Proc. Natl. Acad. Sci. USA 92. Dahmus, M. E. 1996. Reversible phosphorylation of the C-terminal domain of RNA polymerase H. J Biol Chem 271:19009-12. Dahmus, M. E. 1994. The role of multisite phosphorylation in the regulation of RNA polymerase H activity. Prog Nucleic Acid Res Mol Biol 48: 143-79. Damania, B., P. Lieberman, and J. C. Alwine. 1998. Simian virus 40 large T antigen stabilizes the TATA-binding protein-TFHA complex on the TATA element. Mol Cell Biol 18:3926-35. DeJong, J ., R. Bernstein, and R. G. Roeder. 1995. Human general transcription factor TFHA: Characterization of a cDNA encoding the small subunit and requirement for basal and activated transcription. Proc. Natl. Acad. Sci. USA. 92:3313-3317. Donaldson, L., and J. P. Capone. 1992. Purification and characterization of the carboxyl-terminal transactivation domain of Vmw65 from herpes simplex virus type 1. J. Biol. Chem. 267:1411-1414. Darn, A., J. Bollekens, A. Staub, C. Benoist, and D. Mathis. 1987. A multiplicity of CCAAT box-binding proteins. Cell 50:863-72. Drysdale, C. M., E. Duenas, B. M. Jackson, U. Reusser, G. H. Braus, and A. G. Hinnebusch. 1995. The transcriptional activator GCN4 contains multiple activation domains that are dependent on hydrophobic amino acids. Mol Cell Biol 15:1220- 1233. Dvir, A., R. C. Conaway, and J. W. Conaway. 1997. A role for TFIH-1 in controlling the activity of early RNA polymerase H elongation complexes. Proc Natl Acad Sci U S A 94:9006-10. 112 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. Edmondson, D. G., M. M. Smith, and S. Y. Roth. 1996. Repression domain of the yeast global repressor Tupl interacts directly with histones H3 and H4. Genes Dev 10: 1247-59. Edwards, A. M., C. M. Kane, R. A. Young, and R. D. Kornberg. 1991. Two dissociable subunits of yeast RNA polymerase-II stimulate the initiation of transcription at a promoter invitro. J. Biol. Chem. 266:71-75. Eisenmann, D. M., C. Chapon, S. M. Roberts, C. Bollard, and F. Winston. 1994. The Saccharomyces cerevisiae SPT8 gene encodes a very acidic protein that is functionally related to SPT3 and TATA-binding protein. Genetics 137:647-57. Eisenmann, D. M., C. Ballard, and F. Winston. 1989. SPTlS, the gene encoding the yeast TATA binding factor TFHD, is required for normal transcription initiation in vivo. Cell 58:1183-1191. Emami, K. H., and M. Carey. 1992. A synergistic increase in potency of a multimerized VP16 transcriptional activation domain. EMBO J. 11:5005-5012. Emili, A., J. Greenblatt, and J. C. Ingles. 1994. Species-specific interaction of the glutamine-rich activation domains of Spl with the TATA box-binding protein. Mol. Cell. Biol. 14:1582-1593. Estruch, J. J ., L. Crossland, and S. A. Goff. 1994. Plant activating sequences: positively charged peptides are functional as transcriptional activation domains. Nucl. Acids Res. 22:3983-3989. Fang, S. M., and Z. F. Burton. 1996. RNA polymerase H-associated protein (RAP) 74 binds transcription factor (TF) HB and blocks TFHB-RAP30 binding. J Biol Chem 271:11703-9. Feaver, W. J ., O. Gileadi, Y. Li, and R. D. Kornberg. 1991. CTD kinase associated with yeast RNA polymerase H initiation factor b. Cell 67: 1223-1230. Feng, W., R. C. Ribeiro, R. L. Wagner, H. Nguyen, J. W. Apriletti, R. J. Fletterick, J. D. Baxter, P. J. Kushner, and B. L. West. 1998. Hormone-dependent coactivator binding to a hydrophobic cleft on nuclear receptors. Science 280: 1747-9. Finch, J. T., and A. Klug. 1976. Solenoidal model for superstructure in chromatin. Proc Natl Acad Sci U S A 73:1897-901. Flores, O., H. Lu, and D. Reinberg. 1992. Factors involved in specific transcription by mammalian RNA polymerase-II- identification and characterization of factor-11H. J. Biol. Chem. 267:2786-2793. Flores, 0., E. Maldonado, and D. Reinberg. 1989. Factors involved in specific transcription by mammalian RNA polymerase H. Factors HE and HF independently interact with RNA polymerase H. J Biol Chem 264:8913-21. Forget, D., F. Robert, G. Grondin, Z. F. Burton, J. Greenblatt, and B. Coulombe. 1997. RAP74 induces promoter contacts by RNA polymerase H upstream and downstream of a DNA bend centered on the TATA box. Proc Natl Acad Sci U S A 94:7150-5. Forsburg, S. L., and L. Guarente. 1989. Identification and characterization of HAP4: a third component of the CCAAT-bound HAP2/HAP3 heteromer. Genes Dev. 3:1166-1178. Furter Graves, E. M., B. D. Hall, and R. Furter. 1994. Role of a small RNA pol H subunit in TATA to transcription start site spacing. Nucleic Acids Res 22:4932-6. 113 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. Gadbois, E. L., D. M. Chao, J. C. Reese, M. R. Green, and R. A. Young. 1997. Functional antagonism between RNA polymerase H holoenzyme and global negative regulator NC2 in vivo. Proc Natl Acad Sci U S A 94:3145-50. Gansheroff, L. J., C. Bollard, P. Tan, and F. Winston. 1995. The Saccharomyces cerevisiae SPT7 gene encodes a very acidic protein important for transcription in vivo. Genetics 139:523-36. Garcia Martinez, L. F., G. Mavankal, J. M. Neveu, W. S. Lane, D. Ivanov, and R. B. Gaynor. 1997. Purification of a Tat-associated kinase reveals a TFIH-1 complex that modulates HIV-l transcription. EMBO J 16:2836-50. Gavin, I. M., and R. T. Simpson. 1997. Interplay of yeast global transcriptional regulators Ssn6p-Tuplp and Swi-Snf and their effect on chromatin structure. EMBOJ 16:6263-71. Ge, H., and R. G. Roeder. 1994. The high mobility group protein HMG] can reversibly inhibit class H gene transcription by interaction with the TATA-binding protein. J Biol Chem 269: 17136-40. Ge, H., Y. Zhao, B. T. Chait, and R. G. Roeder. 1994. Phosphorylation negatively regulates the function of coactivator PC4. Proc Natl Acad Sci U S A 91: 12691-5. Georgakopoulos, T., N. Gounalaki, and G. Thireos. 1995. Genetic evidence for the interaction of the yeast transcriptional co-activator proteins GCN5 and ADA2. Mol. Gen. Genet. 246:723-728. Georgakopoulos, T., and G. Thireos. 1992. Two distinct yeast transcriptional activators require the function of the GCN5 protein to promote normal levels of transcription. EMBO J 11:4145-4152. Gill, G., E. Pascal, Z. H. Tseng, and R. Tjian. 1994. A glutamine-rich hydrophobic patch in transcription factor Spl contacts the dTAFHl 10 component of the Drosophila TFHD complex and mediates transcriptional activation. Proc. Natl. Acad. Sci. USA 91:192-196. Gill, G., and M. Ptashne. 1988. Negative effect of the transcriptional activator GAL4. Nature 334:721-724. Goodrich, J. A., T. Hoey, C. J. Thut, A. Admon, and R. Tjian. 1993. Dr050phila TAF(H)40 interacts with both a VP16 activation domain and the basal transcription factor TFIIB. Cell 75:519-530. Gotta, M., and S. M. Gasser. 1996. Nuclear organization and transcriptional silencing in yeast. Experientia 52:1136-47. Gottschling, D. E., O. M. Aparicio, B. L. Billington, and V. A. Zakian. 1990. Position effect at S. cerevisiae telomeres: reversible repression of Pol H transcription. Cell 63:751-62. Grant, P. A., L. Duggan, J. Cote, S. M. Roberts, J. E. Brownell, R. Candau, R. Ohba, T. Owen-Hughes, C. D. Allis, F. Winston, S. L. Berger, and J. L. Workman. 1997. Yeast Gcn5 functions in two multisubunit complexes to acetylate nucleosomal histones: Characterization of an Ada complex and the SAGA (Spt/Ada) complex. Genes Dev. 11: 1640-1650. Grant, P. A., D. Schieltz, M. G. Pray Grant, D. J. Steger, J. C. Reese, J. R. Yates, 3rd, and J. L. Workman. 1998. A subset of TAF(H)s are integral components of the SAGA complex required for nucleosome acetylation and transcriptional stimulation [see comments]. Cell 94:45-53. 114 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. Greenblatt, J. 1991. RNA polymerase-associated transcription factors. TIBS 16:408- 411. Grosschedl, R., and M. L. Birnstiel. 1980. Identification of regulatory sequences in the prelude sequences of an H2A histone gene by the study of specific deletion mutants in vivo. Proc Natl Acad Sci U S A 77: 1432-6. Guarente, L. 1983. Yeast promoters and lacZ fusions designed to study expression of cloned genes in yeast. Methods Enzymol. 101: 181-191. Guarente, L., P. Lalonde, P. Gifford, and E. Alani. 1984. Distinctly regulated tandem upstream activation sites mediate catabolite repression of the CYCI gene of S. cerevisiae. Cell 36:503-51 1. Gugneja, S., C.-M. A. Virbasius, and R. C. Scarpulla. 1996. Nuclear respiratory factors 1 and 2 utilize similar glutamine containing clusters of hydrophobic residues to activate transcription. Mol. Cell. Biol. 16:5708-5716. Guthrie, C., and G. R. Fink. 1991. Guide to Yeast Genetics and Molecular Biology, vol. 194. Academic Press. Ha, 1., W. S. Lane, and D. Reinberg. 1991. Cloning of a human gene encoding the general transcription initiation factor- HB. Nature 352:689-695. Haber, J. E., and J. P. George. 1979. A mutation that permits the expression of normally silent copies of mating-type information in Saccharomyces cerevisiae. Genetics 93: 13-35. Hahn, S., S. Buratowski, P. A. Sharp, and L. Guarente. 1989. Yeast TATA- binding protein TFHD binds to TATA elements with both consensus and nonconsensus DNA sequences. Proc. Natl. Acad. Sci. USA. 86:5718-5722. Hahn, S., and L. Guarente. 1988. Yeast HAP2 and HAP3: transcriptional activators in a heteromeric complex. Science 240:317-321. Hahn, S., E. T. Hoar, and L. Guarente. 1985. Each of three "TATA elements" specifies a subset of the transcription initiation sites at the CYC-l promoter of Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 82:8562-6. Hampsey, M. 1998. Molecular genetics of the RNA polymerase H general transcriptional machinery. Microbiol Mol Biol Rev 62:465-503. Hampsey, M. 1997. A SAGA of histone acetylation and gene expression. Trends Genet 13:427-9. Hansen, S. K., and R. Tjian. 1995. TAFs and TFHA mediate differential utilization of the tandem Adh promoters. Cell 82:565-75. He, 8., and S. J. Weintraub. 1998. Stepwise recruitment of components of the preinitiation complex by upstream activators in vivo. Mol Cell Biol 18:2876-83. Healy, A. M., T. L. Helser, and R. S. Zitomer. 1987. Sequences required for transcriptional initiation of the Saccharomyces cerevisiae CYC7 genes. Mol Cell Biol 7:3785-91. Hekmatpanah, D. S., and R. A. Young. 1991. Mutations in a conserved region of RNA polymerase H influence the accuracy of mRNA start site selection. Mol Cell Biol 11:5781-91. Henry, N. L., A. M. Campbell, W. J. Feaver, D. Poon, P. A. Weil, and R. D. Kornberg. 1994. TFHF-TAF-RNA polymerase H connection. Genes Dev. 8:2868- 2878. 115 125. 126. 127. 128. 129. 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. Henry, N. L., M. H. Sayre, and R. D. Kornberg. 1992. Purification and characterization of yeast RNA polymerase H general initiation factor g. J Biol Chem 267:23388-92. Hernandez, N. 1993. TBP, a universal eukaryotic transcription factor? Genes Dev 7:1291-1308. Hirschhorn, J. N., S. A. Brown, C. D. Clark, and F. Winston. 1992. Evidence that SNF2/SW12 and SNFS activate transcription inyeast by altering chromatin structure. Genes & Development 6:2288-2298. Hollenberg, S. M., R. Stemglanz, P. F. Cheng, and H. Weintraub. 1995. Identification of a new family of tissue-specific basic helix-loop-helix proteins with a two-hybrid system. Mol Cell Biol 15:3813-22. Holstege, F. C., E. G. JenningS, J. J . Wyrick, T. 1. Lee, C. J. Hengartner, M. R. Green, T. R. Golub, E. S. Lander, and R. A. Young. 1998. Dissecting the regulatory circuitry of a eukaryotic genome. Cell 95:717-28. Holstege, F. C., D. Tantin, M. Carey, P. C. van der Vliet, and H. T. Timmers. 1995. The requirement for the basal transcription factor HE is determined by the helical stability of promoter DNA. EMBO J 14:810-9. Hong, H., K. Kohli, A. Trivedi, D. L. Johnson, and M. R. Stallcup. 1996. GRIPl , a novel mouse protein that serves as a transcriptional coactivator in yeast for the hormone binding domains of steroid receptors. Proc Natl Acad Sci U S A 93:4948- 52. Horikoshi, N., A. Usheva, J. Chen, A. J. Levine, R. Weinmann, and T. Shenk. 1995. Two domains of p53 interact with the TATA-binding protein, and the adeno virus 13S ElA protein disrupts the association, relieving p53-mediated transcriptional repression. Mol. Cell. Biol. 15:227-234. Horiuchi, J ., N. Silverman, G. A. Marcus, and L. Guarente. 1995. ADA3, a putative transcriptional adaptor, consists of two separable domains and interacts with ADA2 and GCN5 in a trimeric complex. Mol. Cell. Biol. 15:1203-1209. Horiuchi, J ., N. Silverman, B. Pina, G. A. Marcus, and L. Guarente. 1997. ADAl , a novel component of the ADA/GCN5 complex, has broader effects than GCN5, ADA2, or ADA3. Mol. Cell. Biol. 17:3220-3228. Horn, P. 1999. Ph.D. thesis. Michigan State University, East Lansing. Huang, L., W. Zhang, and S. Y. Roth. 1997. Amino termini of histones H3 and H4 are required for al-alpha2 repression in yeast. Mol Cell Biol 17:6555-62. Hull, M. W., K. McKune, and N. A. Woychik. 1995. RNA polymerase H subunit RPB9 is required for accurate start site selection. Genes Dev 9:481-90. Imbalzano, A. N., K. S. Zaret, and R. E. Kingston. 1994. Transcription factor TFHB and TFHA can independently increase the affinity of the TATA-binding protein for DNA. J. Biol. Chem. 269:8280-8286. Ingles, C. J., M. Shales, W. D. Cress, S. J. Triezenberg, and J. Greenblatt. 1991. Reduced binding of TFHD to transcriptionally compromised mutants of VP16. Nature 351:588-590. Iniguez-Lluhi, J. A., D. Y. Lou, and K. R. Yamamoto. 1997. Three amino acid substitutions selectively disrupt the activation but not the repression function of the glucocorticoid receptor N terminus. J. Biol. Chem. 272:4149-4156. 116 141. 142. 143. 144. 145. 146. 147. 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. Inostroza, J. A., F. H. Mermelstein,1. Ha, W. S. Lane, and D. Reinberg. 1992. Dr] , a TATA-binding protein-associated phosphoprotein and inhibitor of class H gene transcription. Cell 70:477-489. Iyer, V., and K. Struhl. 1995. Poly (dA:dT), a ubiquitous promoter element that stimulates transcription via its intrinsic DNA structure. EMBO J. 14:2570-2579. Izban, M. G., and D. S. Luse. 1992. Factor-stimulated RNA polymerase H transcribes at physiological elongation rates on naked DNA but very poorly on chromatin templates. J Biol Chem 267 :13647-55. Jackson, B. M., C. M. Drysdale, K. Natarajan, and A. G. Hinnebusch. 1996. Identification of seven hydrophobic clusters in GCN4 making redundant contributions to transcriptional activation. Mol. Cell. Biol 16:5557-5571. Javahery, R., A. Khachi, K. Lo, B. Zenzie Gregory, and S. T. Smale. 1994. DNA sequence requirements for transcriptional initiator activity in mammalian cells. Mol Cell Biol 14:116-27. Jiang, Y., S. J. Triezenberg, and J. D. Gralla. 1994. Defective transcriptional activation by diverse VP16 mutants associated with a common inability to form open promoter complexes. J. Biol. Chem. 269:5505-5508. Johnson, A. D. 1995. The price of repression. Cell 81:655-658. Johnson, P. F., E. Sterneck, and S. C. Williams. 1993. Activation domains of transcriptional regulatory proteins. J. Nutr. Biochem. 4:386-398. Kadosh, D., and K. Struhl. 1997. Repression by Ume6 involves recruitment of a complex containing Sin3 corepressor and de3 histone deacetylase to target promoters. Cell 89:365-71. Kadosh, D., and K. Struhl. 1998. Targeted recruitment of the Sin3-de3 histone deacetylase complex generates a highly localized domain of repressed chromatin in vivo. Mol Cell Biol 18:5121-7. Kamei, Y., L. Xu, T. Heinzel, J. Torchia, R. Kurokawa, B. Gloss, S. C. Lin, R. A. Heyman, D. W. Rose, C. K. Glass, and M. G. Rosenfeld. 1996. A CBP integrator complex mediates transcriptional activation and AP-l inhibition by nuclear receptors. Cell 85:403- 14. Kaufman, J ., and S. T. Smale. 1994. Direct recognition of initiator elements by a component of the transcription factor HD complex. Genes Dev 8:821-9. Kaufmann, J ., C. P. Verrijzer, J. Shao, and S. T. Smale. 1996. CIF, an essential cofactor for TFIID-dependent initiator function. Genes Dev. 10:873-886. Kim, T. K., S. Hashimoto, R. J. Kelleher, III, P. M. Flanagan, R. D. Kornberg, M. Horikoshi, and R. G. Roeder. 1994. Effects of activation-defective TBP mutations on transcription initiation in yeast. Nature 369:252-255. Kim, T. K., T. Lagrange, Y. H. Wang, J. D. Griffith, D. Reinberg, and R. H. Ebright. 1997. Trajectory of DNA in the RNA polymerase H transcription preinitiation complex. Proc Natl Acad Sci U S A 94: 12268-73. Kim, T. K., and R. G. Roeder. 1997. Critical role of the second stirrup region of the TATA-binding protein for transcriptional activation both in yeast and human. J. Biol. Chem. 272:7540-7545. Kim, Y. C., J. H. Geiger, S. Hahn, and P. B. Sigler. 1993. Crystal structure of a yeast TBP/TATA-Box complex. Nature 365:512-520. 117 158. 159. 1 60. 161. 162. 163. 164. 165. 166. 167. 168. 169. 170. 171. 172. 173. 174. Kim, Y.-J., S. Bjorklund, Y. Li, M. H. Sayre, and R. D. Kornberg. 1994. A multiprotein mediator of transcriptional activation and its interaction with the c- terminal repeat domain of RNA polymerase H. Cell 77:599-608. Klages, N. a. S., M. 1995. Stimulation of RNA polymerase H transcription initiation by recruitment of TBP in vivo. Nature 374:822-3. Klar, A. J. S., S. Fogel, and K. Macleod. 1979. MARI—a regulator of HMa and HMalpha loci in Saccharomyces cerevisiae. Genetics 93:37-50. Klebanow, E. R., D. Poon, S. Zhou, and P. A. Weil. 1997 . Cloning and characterization of an essential Saccharomyces cerevisiae gene, TAF40, which encodes yTAFH40, an RNA polymerase H-specific TATA-binding protein-associated factor. J Biol Chem 272:9436-42. Klebanow, E. R., D. Poon, S. Zhou, and P. A. Weil. 1996. Isolation and characterization of TAF25, an essential yeast gene that encodes an RNA polymerase H-specific TATA-binding protein-associated factor. J Biol Chem 271:13706-15. Klein, C., Struhl, K. 1994. Increased recruitment of TATA-binding protein to the promoter by transcriptional activation domains in vivo. Science 266:280-282. Kobayashi, N., T. G. Boyer, and A. J. Berk. 1995. A class of activation domains interacts directly with TFHA and stimulates TFHA-TFHD-promoter complex assembly. Mol. Cell. Biol. 15:6465-6473. Kobayashi, N., P. J. Horn, S. M. Sullivan, S. J. Triezenberg, T. G. Boyer, and A. J. Berk. 1998. DA-complex assembly activity required for VP16C transcriptional activation. Mol. Cell. Biol. 18:4023-4031. Koleske, A. J., and R. A. Young. 1994. An RNA polymerase H holoenzyme responsive to activators. Nature 368:466-469. Kolodziej, P., and R. A. Young. 1989. RNA polymerase H subunit RPB3 is an essential component of the mRN A transcription apparatus. Mol Cell Biol 9:5387-94. Kolodziej, P. A., N. Woychik, S. M. Liao, and R. A. Young. 1990. RNA polymerase H subunit composition, stoichiometry, and phosphorylation. Mol. Cell. Biol. 10: 1915-1920. Kolodziej, P. A., and R. A. Young. 1991. Mutations in the three largest subunits of yeast RNA polymerase H that affect enzyme assembly. Mol Cell Biol 11:4669-78. Kristie, T. M., and B. Roizman. 1988. Differentiation and DNA contact points of host proteins binding at the cis site for virion-mediated induction of alpha genes of herpes simplex virus 1. J. Virol. 62:1145-1157. Kruger, W., C. L. Peterson, A. Sil, C. Coburn, G. Arents, E. N. Moudrianakis, and I. Herskowitz. 1995. Amino acid substitutions in the structured domains of histones H3 and H4 partially relieve the requirement of the yeast SWIISNF complex for transcription. Genes Dev 9:2770-9. Kuldell, N. H., and S. Buratowski. 1997. Genetic analysis of the large subunit of yeast transcription factor IIE reveals two regions with distinct functions. Mol Cell Biol 17:5288-98. Kunkel, G. R., and H. G. Martinson. 1981. Nucleosomes will not form on double- stranded RNa or over poly(dA).poly(dT) tracts in recombinant DNA. Nucleic Acids Res 9:6869-88. Kuo, M. H., J. E. Brownell, R. E. Sobel, T. A. Ranalli, R. G. Cook, D. G. Edmonson, S. Y. Roth, and C. D. Allis. 1996. GCN5p, a yeast nuclear histone 118 175. 176. 177. 178. 179. 1 80. 181. 182. 183. 184. 185. 186. 187. 188. 189. 190. acetyltransferase, acetylates specific lysines in histone H3 and H4 that differ from deposition-related acetylation sites. Nature 383:261-272. Kussie, P. H., S. Gorina, V. Marechal, B. Elenbaas, J. Moreau, A. J. Levine, and N. P. Pavletich. 1996. Structure of the MDM2 oncoprotein bound to the p53 tumor suppressor transactivation domain. Science 274:948-953. Labow, M. A., S. B. Baim, T. Shenk, and A. J. Levine. 1990. Conversion of the lac repressor into an allosterically regulated transcriptional activator for mammalian cells. Mol. Cell. Biol. 10:3343-3356. Lagrange, T., A. N. Kapanidis, H. Tang, D. Reinberg, and R. H. Ebright. 1998. New core promoter element in RNA polymerase H-dependent transcription: sequence-specific DNA binding by transcription factor HB. Genes Dev 12:34-44. Landsman, D., and M. Bustin. 1991. Assessment of the transcriptional activation potential of the HMG chromosomal proteins. Mol Cell Biol 11:4483-9. Lee, J. M., and A. L. Greenleaf. 1991. CTD kinase large subunit is encoded by CT K1 , a gene required for normal growth of Saccharomyces cervisiae. Gene Expr. 1: 149-167. Lee, M., and K. Struhl. 1995 . Mutations on the DNA-binding surface of TATA- binding protein can specifically impair the response to acidic activators in vivo. Mol. Cell. Biol. 15:5461-5469. Lei, L., D. Ren, A. Finkelstein, and Z. F. Burton. 1998. Functions of the N- and C- terminal domains of human RAP74 in transcriptional initiation, elongation, and recycling of RNA polymerase H. Mol Cell Biol 18:2130-42. Leuther, K. K., D. A. Bushnell, and R. D. Kornberg. 1996. Two dimensional crystallography of TFHB- and HE-RNA polymerase H complexes: Implications for start site selection and initiation complex formation. Cell 85:773-779. Leuther, K. K., J. M. Salmeron, and S. A. Johnston. 1993. Genetic evidence that an activation domain of GAL4 does not require acidity and may form a beta sheet. Cell 72:575-585. Li, W. Z., and F. Sherman. 1991. Two types of TATA elements for the CYC 1 gene of the yeast Saccharomyces cerevisiae. Mol. Cell. Biol. 11:666-676. Li, Y., S. Bjorklund, Y. J. Kim, and R. D. Kornberg. 1996. Yeast RNA polymerase H holoenzyme. Methods Enzymol 273: 172—5. Li, Y., P. M. Flanagan, H. Tschochner, and R. D. Kornberg. 1994. RNA polymerase H initiation factor interactions and transcription start site selection. Science 263:805-807. Liao, S., J. Zhang, D. A. Jeffery, A. J. Koleske, C. M. Thompson, D. M. Chao, M. Viljoen, H. J. J. van Vuuren, and R. A. Young. 1995. A kinase-cyclin pair in the RNA polymerase H holoenzyme. Nature 374: 193-196. Licht, J. D., M. Ro, M. A. English, M. Grossel, and U. Hansen. 1993. Selective repression of transcriptional activators at a distance by the Drosophila Kruppel protein. Proc Natl Acad Sci U S A 90:11361-5. Liljelund, P., C. J. Ingles, and J. Greenblatt. 1993. Altered promoter binding of the TATA box-binding factor induced by the transcriptional activation domain of VP16 and suppressed by TFHA. Mol. Gen. Genet. 241:694-699. Lin, J., J. Chen, B. Elenbaas, and A. J. Levine. 1994. Several hydrophobic amino acids in the p53 amino-terminal domain are required for transcriptional activation, 119 191. 192. 193. 194. 195. 196. 197. 198. 199. 200. 201. 202. 203. 204. 205. 206. binding to mdm-2 and the adenovirus 5 ElB 55-kD protein. Genes Dev. 8: 1235- 1246. Lin, Y. S., and M. R. Green. 1991. Mechanism of action of an acidic transcriptional activator invitro. Cell 64:971-981. Lin, Y. S., 1. Ha, E. Maldonado, D. Reinberg, and M. R. Green. 1991. Binding of general transcription factor TFHB to an acidic activating region. Nature 353:569-571. L00, 8., and J. Rine. 1995. Silencing and heritable domains of gene expression. Annu. Rev. Biol. Dev. 11:519-548. Lowell, J. E., and L. Pillus. 1998. Telomere tales: chromatin, telomerase and telomere function in Saccharomyces cerevisiae. Cell Mol Life Sci 54:32-49. Lu, H., R. P. Fisher, P. Bailey, and A. J. Levine. 1997. The CDK7-cycH-p36 complex of transcription factor IH-I phosphorylates p53, enhancing its sequence- specific DNA binding activity in vitro. Mol Cell Biol 17:5923-34. Lu, H., O. Flores, R. Weinmann, and D. Reinberg. 1991. The nonphosphorylated form of RNA polymerase H preferentially associates with the preinitiation complex. Proc. Natl. Acad. Sci. USA 88: 10004-10008. Lu, H., L. Zawel, L. Fisher, J. M. Egly, and D. Reinberg. 1992. Human general transcription factor 1H1 phosphorylates the C-temrinal domain of RNA polymerase H. Nature 358:641-645. Lue, N. F., A. R. Buchman, and R. D. Kornberg. 1989. Activation of yeast RNA polymerase II transcription by a thymidine-rich upstream element in vitro. Proc Natl Acad Sci U S A 86:486-90. Luger, K., A. W. Mader, R. K. Richmond, D. F. Sargent, and T. J. Richmond. 1997. Crystal structure of the nucleosome core particle at 2.8 angstrom resolution. Nature 6648:251-260. Lycan, D., G. Mikesell, M. Bunger, and L. Breeden. 1994. Differential effects of Cdc68 on cell cycle-regulated promoters in Saccharomyces cerevisiae. Mol Cell Biol 14:7455-65. Ma, D., H. Watanbe, F. Mermelstein, A. Admon, K. Oguri, X. Sun, T. Wada, T. lmai, S. Toshifumi, D. Reinbeg, and H. Handa. 1993. Isolation of a cDN A encoding the largest subunit of TFHA reveals functions important for activated transcription. Genes Devel 7:2246-2257. MacDonald, G. H., Y. Itoh Lindstrom, and J. P. Ting. 1995. The transcriptional regulatory protein, YB-l, promotes single-stranded regions in the DRA promoter. J Biol Chem 270:3527-33. Maire, P., J. Wuarin, and U. Schibler. 1989. The role of cis-acting promoter elements in tissue-specific albumin gene expression. Science 244:343-6. Maity, S. N., P. T. Golumbek, G. Karsenty, and B. de Crombrugghe. 1988. Selective activation of transcription by a novel CCAAT binding factor. Science 241:582-5. Majello, B., G. Napolitano, P. De Luca, and L. Lania. 1998. Recruitment of human TBP selectively activates RNA polymerase H TATA-dependent promoters. J Biol Chem 273: 16509-16. Malik, S., K. Hisatake, H. Sumirnoto, M. Horikoshi, and R. G. Roeder. 1991. Sequence of general transcription factor TFHB and relationships to other initiation factors. Proc. Natl. Acad. Sci. USA 88:9553-9557. 120 207. 208. 209. 210. 211. 212. 213. 214. 215. 216. 217. 218. 219. 220. 221. 222. 223. Malone, E. A., C. D. Clark, A. Chiang, and F. Winston. 1991. Mutations in SPTl6/CDC68 suppress cis— and trans-acting mutations that affect promoter function in Saccharomyces cerevisiae. Mol Cell Biol 11:5710-7. Marcus, G. A., J. Horiuchi, N. Silverman, and L. Guarente. 1996. ADA5/SPT20 links the ADA and SPT genes, which are involved in yeast transcription. Mol. Cell. Biol. 16:3197-3205. Marcus, G. A., N. Silverman, S. L. Berger, J. Horiuchi, and L. Guarente. 1994. Functional similarity and physical association between GCN5 and ADA2: putative transcriptional adaptors. EMBO J. 13:4807-4815. Markovtsov, V., A. Mustaev, and A. Goldfarb. 1996. Protein-RNA interactions in the active center of transcription elongation complex. Proc Natl Acad Sci U S A 93:3221-6. Marshall, N. F., J. Peng, Z. Xie, and D. H. Price. 1996. Control of RNA polymerase H elongation potential by a novel carboxyl-terminal domain kinase. J Biol Chem 271:27176-83. Martinez, E., C. M. Chiang, H. Ge, and R. G. Roeder. 1994. TATA-binding protein-associated factor(s) in TFHD function through the initiator to direct basal transcription from a TATA-less class H promoter. EMBO J 13:3115-26. Maxon, M. E., and R. Tjian. 1994. Transcriptional activity of transcription factor HE is dependent on zinc binding. Proc Natl Acad Sci U S A 91:9529-33. McKune, K., P. A. Moore, M. W. Hull, and N. A. Woychik. 1995. Six human RNA polymerase subunits functionally substitute for their yeast counterparts. Mol Cell Biol 15:6895-900. McNabb, D. S., Y. Xing, and L. Guarente. 1995. Cloning of yeast HAPS: a novel subunit of a heterotrimeric complex required for CCAAT binding. Genes Dev. 9:47- 58. Means, A. L., J. E. Slansky, S. L. McMahon, M. W. Knuth, and P. J. Famham. 1992. The HIP] binding site is required for growth regulation of the dihydrofolate reductase gene promoter. Mol Cell Biol 12: 1054-63. Meisteremst, M., and R. G. Roeder. 1991. Family of proteins that interact with tfiid and regulate promoter activity. Cell 67:557-567. Meisteremst, M., A. L. Roy, H. M. Lieu, and R. G. Roeder. 1991. Activation of class H gene transcription by regulatory factors is potentiated by a novel activity. Cell 66:981-993. Melcher, K., and S. A. Johnston. 1995. GAL4 interacts with TATA-binding protein and coactivators. Mol. Cell. Biol. 15:2839-2848. Merino, A., K. R. Madden, W. S. Lane, J. J. Champoux, and D. Reinberg. 1993. DNA topoisomerase I is involved in both repression and activation of transcription. Nature 365:227-234. Michel, B., P. Komamitsky, and S. Buratowski. 1998. Histone-like TAFs are essential for transcription in vivo. Mol Cell 2:663-73. Miller, J. H. 1972. Experiments in Molecular Genetics, Cold Spring Harbor, New York. Mitchell, P. J., and R. Tjian. 1989. Transcriptional regulation in mammalian cells by sequence-specific DNA binding proteins. Science 245:371-378. 121 224. 225. 226. 227. 228. 229. 230. 231. 232. 233. 234. 235. 236. 237. 238. 239. Moqtaderi, Z., Y. Bai, D. Poon, P. A. Weil, and K. Struhl. 1996. TBP-associated factors are not generally required for transcriptional activation in yeast. Nature 383: 188-191. Moqtaderi, Z., M. Keaveney, and K. Struhl. 1998. The histone H3-like TAF is broadly required for transcription in yeast. Mol Cell 2:675-82. Moqtaderi, Z., J. D. Yale, K. Struhl, and S. Buratowski. 1996. Yeast homologues of higher eukaryotic TFHD subunits. Proc Natl Acad Sci U S A 93: 14654-8. Mulder, W., I. H. Scholten, R. W. de Boer, and L. A. Grivell. 1994. Sequence of the HAP3 transcription factor of Kluyveromyces lactis predicts the presence of a novel 4-cysteine zinc-finger motif. Mol Gen Genet 245:96-106. Mustaev, A., M. Kozlov, V. Markovtsov, E. Zaychikov, L. Denissova, and A. Goldfarb. 1997. Modular organization of the catalytic center of RNA polymerase. Proc Natl Acad Sci U S A 94:6641-5. Myers, L. C., C. M. Gustafsson, D. A. Bushnell, M. Lui, H. Erdjument Bromage, P. Tempst, and R. D. Kornberg. 1998. The Med proteins of yeast and their function through the RNA polymerase H carboxy-terminal domain. Genes Dev 12:45-54. Nagawa, F., and G. R. Fink. 1985. The relationship between the "TATA" sequence and transcription initiation sites at the HIS4 gene of Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 82:8557-61. Nakatani, Y., S. Bagby, and M. Ikura. 1996. The histone folds in transcription factor TFHD. J Biol Chem 271:6575-8. Neigebom, L., and M. Carlson. 1984. Genes affecting the regulation of SUC2 gene expression by glucose repression in Saccharomyces cerevisiae. Genetics 108:845-58. Neigebom, L., J. L. Celenza, and M. Carlson. 1987. SSN20 is an essential gene with mutant alleles that suppress defects in SUC2 transcription in Saccharomyces cerevisiae. Mol Cell Biol 7 :672-8. Neish, A. S., S. F. Anderson, B. P. Schlegel, W. Wei, and J. D. Parvin. 1998. Factors associated with the mammalian RNA polymerase H holoenzyme. Nucleic Acids Res 26:847-53. Nelson, H. C., J. T. Finch, B. F. Luisi, and A. Klug. 1987. The structure of an oligo(dA).oligo(dT) tract and its biological implications. Nature 330:221-6. Nguyen, C., M. Bolotin Fukuhara, M. Wesolowski Louvel, and H. Fukuhara. 1995. The respiratory system of Kluyveromyces lactis escapes from HAP2 control. Gene 152:113-5. Nightingale, K., S. Dimitrov, R. Reeves, and A. P. Wolffe. 1996. Evidence for a shared structural role for HMG] and linker histones B4 and H1 in organizing chromatin. EMBO J 15:548-61. Nikolov, D. B., H. Chen, E. D. Halay, A. A. Usheva, K. Hisatake, D. K. Lee, R. G. Roeder, and S. K. Burley. 1995. Crystal structure of a TFIIB-TBP-TATA-element ternary complex. Nature 377:1 19-128. Nolte, R. T., G. B. Wisely, S. Westin, J. E. Cobb, M. H. Lambert, R. Kurokawa, M. G. Rosenfeld, T. M. Willson, C. K. Glass, and M. V. Milburn. 1998. Ligand binding and co-activator assembly of the peroxisome proliferator-activated receptor- gamma. Nature 395: 137-43. 122 240. 241. 242. 243. 244. 245. 246. 247. 248. 249. 250. 251. 252. 253. 254. 255. Nonet, M. L., and R. A. Young. 1989. Intrgenic and extragenic suppressors of mutations in the heptapeptide repeat domain of Saccharomyces cerevisiae RNA polymerase H. Genetics 123:715-724. O’Brien, T., S. Hardin, A. Greenleaf, and J. T. Lis. 1994. Phosphorylation of RNA polymerase H C-terminal domain and transcriptional elongation. Nature 370:75-77. O’Hare, P., and C. R. Goding. 1988. Herpes simplex virus regulatory elements and the immunoglobulin octamer bind a common factor and are both targets for virion transactivation. Cell 52:435-445. O’Hare, P., and G. Williams. 1992. Structural studies of the acidic transactivation domain of the Vmw65 protein of herpes simplex virus using 1H NMR. Biochemistry 31:4150-4156. Ohkuma, Y., S. Hashimoto, C. K. Wang, M. Horikoshi, and R. G. Roeder. 1995. Analysis of the role of TFHE in basal transcription and TFHB-mediated carboxy- terminal domain phosphorylation through structure-function studies of TFHB-alpha. Mol Cell Biol 15:4856-66. Ohkuma, Y., and R. G. Roeder. 1994. Regulation of Tfiih ATPase and kinase activities by Tfiie during active initiation complex formation. Nature 368: 160-163. Olesen, J. T., and L. Guarente. 1990. The HAP2 subunit of yeast CCAAT transcriptional activator contains adjacent domains for subunit association and DNA recognition - model for the HAP2/3/4 complex. Genes Dev. 4: 1714-1729. Olesen, J. T., S. Hahn, and L. Guarente. 1987. Yeast HAP2 and HAP3 activators both bind to the CYCl upstrem activation site UASZ in an interdependent manner. Cell 51:953-961. Onate, S. A., S. Y. Tsai, M. J. Tsai, and O. M. BW. 1995. Sequence and characterization of a coactivator for the steroid hormone receptor superfamily. Science 270: 1354-7. Ozer, J ., L. E. Lezina, J. Ewing, S. Audi, and P. M. Lieberman. 1998. Association of transcription factor HA with TATA binding protein is required for transcriptional activation of a subset of promoters and cell cycle progression in Saccharomyces cerevisiae. Mol Cell Biol 18:2559-70. Ozer, J., P. A. Moore, A. H. Bolden, A. Lee, C. A. Rosen, and P. M. Lieberman. 1994. Molecular cloning of the small (gamma) subunit of human TFHA reveals functions critical for activated transcription. Genes Dev. 8:2324-2335. Pan, G., and J. Greenblatt. 1994. Initiation of transcription by RNA polymerase H is limited by melting of the promoter DNA in the region immediately upstream of the initiation site. J Biol Chem 269:30101-4. Parvin, J. D., and P. A. Sharp. 1993. DNA topology and a minimal set of basal factors for transcription by RNA polymerase H. Cell 73:533-540. Parvin, J. D., B. M. Shykind, R. E. Meyers, J. Kim, and P. A. Sharp. 1994. Multiple sets of basal factors initiate transcription by RNA polymerase H. J Biol Chem 269:18414-21. Pauli, T. T., M. Carey, and R. C. Johnson. 1996. Yeast HMG proteins NHP6A/B potentiate promoter-specific transcriptional activation in vivo and assembly of preinitiation complexes in vitro. Genes Dev 10:2769-81. Pazin, M. J., and J. T. Kadonaga. 1997 . What’s up and down with histone deacetylation and transcription. Cell 89:325-328. 123 256. 257. 258. 259. 260. 261. 262. 263. 264. 265. 266. 267. 268. 269. 270. 271. Peck, L. J ., and J. C. Wang. 1981. Sequence dependence of the helical repeat of DNA in solution. Nature 292:375-8. Perrella, M. A., A. Pellacani, P. Wiesel, M. T. Chin, L. C. Foster, M. Ibanez, C. M. Hsieh, R. Reeves, S. F. Yet, and M. E. Lee. 1999. High mobility group-KY) protein facilitates nuclear factor-B binding and transactivation of the inducible nitric- oxide synthase promoter/enhancer. J. Biol. Chem. 274:9045-9052. Pina, B., S. Berger, G. A. Marcus, N. Silverman, J. Agapite, and L. Guarente. 1993. ADA3: A gene, identified by resistance to GAL4-VP16, with properties similar to and different from those of ADA2. Mol. Cell. Biol. 13:5981-5989. Pinkham, J. L., and L. Guarente. 1985. Cloning and molecular analysis of the HAP2 locus, a global regulator of respiratory genes in Saccharomyces cerevisiae. Mol. Cel. Biol. 5:3410-3416. Pinkham, J. L., J. T. Olesen, and L. Guarente. 1987. Sequence and nuclear localization of the Saccharomyces cerevisiae HAP2 protein, a transcriptional activator. Mol. Cel. Biol. 7:578-587. Pinto, 1., D. E. Ware, and M. Hampsey. 1992. The yeast SUA7 gene encodes a homolog of human transcription factor TFHB and is required for normal start site selection in vivo. Cell 68:977-988. Poon, D., Y. Bai, A. M. Campbell, S. Bjorklund, Y. J. Kim, S. Zhou, R. D. Kornberg, and P. A. Weil. 1995. Identification and characterization of a TFHD-like multiprotein complex from Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 92:8224-8228. Postnikov, Y. V., V. V. Shick, A. V. Belyavsky, K. R. Khrapko, K. L. Brodolin, T. A. Nikolskaya, and A. D. Mirzabekov. 1991. Distribution of high mobility group proteins 1/2, E and 14/ 17 and linker histones H1 and H5 on transcribed and non- transcribed regions of chicken erythrocyte chromatin. Nucleic Acids Res 19:717-25. Powell, W., and D. Reines. 1996. Mutations in the second largest subunit of RNA polymerase 11 cause 6-azauracil sensitivity in yeast and increased transcriptional arrest in vitro. J Biol Chem 271:6866-73. Prelich, G., and F. Winston. 1993. Mutations that suppress the deletion of an upstream activating sequence in yeast: involvement of a protein kinase and histone H3 in repressing transcription in vivo. Genetics 135:665-76. Preston, C. M., M. C. Frame, and M. E. M. Campbell. 1988. A complex formed between cell components and an HSV structural polypeptide binds to a viral immediate early gene regulatory DNA sequence. Cell 52:425-434. Price, D. H., A. E. Sluder, and A. L. Greenleaf. 1989. Dynamic interaction between a Drosophila transcription factor and RNA polymerase H. Mol Cell Biol 9: 1465-75. Prunell, A. 1982. Nucleosome reconstitution on plasmid-inserted poly(dA) . poly(dT). EMBO J 1:173-9. Pugh, B. F., and R. Tjian. 1992. Diverse transcriptional functions of the multisubunit eukaryotic TFHD complex. J. Biol. Chem. 267:679-682. Quinn, J ., A. M. Fyrberg, R. W. Ganster, M. C. Schmidt, and C. L. Peterson. 1996. DNA-binding properties of the yeast SWIISNF complex. Nature 379:844-7. Radhakrishnan, 1., G. C. Perez-Alvarado, D. Parker, H. J. Dyson, M. R. Montminy, and P. E. Wright. 1997. Solution structure of the KD( domain of CBP 124 272. 273. 274. 275. 276. 277. 278. 279. 280. 281. 282. 283. 284. 285. 286. 287. 288. bound to the transactivation domain of CREB: A model for activatorzcoactivator interactions. Cell 91:741-752. Raghavan, S., P. K. Burma, and S. K. Brahmachari. 1997. Positional preferences of polypurine/polypyrimidine tracts in Saccharomyces cerevisiae genome: implications for cis regulation of gene expression. J Mol Evol 45:485-98. Ranish, J. A., and S. Hahn. 1991. The yeast general transcription factor TFHA is composed of two polypeptide subunits. J. Biol. Chem. 266: 19320-19327. Ranish, J. A., W. S. Lane, and S. Hahn. 1992. Isolation of two genes that encode subunits of the yeast transcription factor HA. Science 255:1127-1129. Reese, J. C., Apone, L., Walker, S. S., Griffin, L. A., Green, M. R. 1994. Yeast TAF(H)s in a multisubunit complex required for activated transcription . Nature 371:523-527. Regier, J. L., F. Shen, and S. J. Triezenberg. 1993. Pattern of aromatic and hydrophobic amino acids critical for one of two subdomains of the VP16 transcriptional activator. Proc. Natl. Acad. Sci. USA 90:883-887. Renz, M., P. Nehls, and J. Hozier. 1977. Involvement of histone H1 in the organization of the chromosome fiber. Proc Natl Acad Sci U S A 74:1879-83. Rhodes, D., and A. Klug. 1981. Sequence-dependent helical periodicity of DNA. Nature 292:378-80. Rine, J., and I. Herskowitz. 1987. Four genes responsible for a position effect on expression from HML and HMR in Saccharomyces cerevisiae. Genetics 116:9-22. Rine, J., J. N. Strathern, J. B. Hicks, and I. Herskowitz. 1979. A suppressor of mating-type locus mutations in Saccharomyces cerevisiae: evidence for and identification of cryptic mating-type loci. Genetics 93:877-901. Robert, F., D. Forget, J. Li, J. Greenblatt, and B. Coulombe. 1996. Localization of subunits of transcription factors HE and HF immediately upstream of the transcriptional initiation site of the adenovirus major late promoter. J Biol Chem 271:8517-20. Roberts, S. G. E., Green, M. R. 1994. Activator-induced conformational change in general transcription factor TFHB. Nature 371:717-720. Roberts, S. G. E., 1. Ha, E. Maldonado, D. Reinberg, and M. R. Green. 1993. Interaction between an acidic activator and transcription factor HB is required for transcriptional activation. Nature 363:741-744. Roberts, S. M., and F. Winston. 1996. SPT20/ADA5 encodes a novel protein functionally related to the TATA-binding protein and important for the transcription in Saccharomyces cerevisiae. M01. Cell. Biol. 16:3206-3213. Rochette-Egly, C., 8. Adam, M. Rossignol, J.-M. Egly, and P. Chambon. 1997. Stimulation of RARalpha activation function AF-l through binding to the general transcription factor TFIH-I and phosphorylation by CDK7. Cell 90:97-107. Ronne, H. 1995. Glucose repression in fungi. Trends Genet 11:12-7. Rose, M. D., F. Winston, and P. Hieter. 1990. Methods in Yeast Genetics: a laboratory course manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Roth, S. Y. 1995. Chromatin-mediated transcriptional repression in yeast. Curr Opin Genet Dev 5:168-73. 125 289. 290. 291. 292. 293. 294. 295. 296. 297. 298. 299. 300. 301. 302. 303. 304. Rowley, A., R. A. Singer, and G. C. Johnston. 1991. CDC68, a yeast gene that affects regulation of cell proliferation and transcription, encodes a protein with a highly acidic carboxyl terminus. Mol Cell Biol 11:5718-26. Roy, A. L., H. Du, P. D. Gregor, C. D. Novina, E. Martinez, and R. G. Roeder. 1997. Cloning of an inr- and E-box-binding protein, TFH-I, that interacts physically and functionally with USFl. EMBO J 16:7091-104. Roy, A. L., S. Malik, M. Meisteremst, and R. G. Roeder. 1993. An alternative pathway for transcription initiation involving TFH-I. Nature 365:355-359. Roy, R., L. Schaeffer, S. Humbert, W. Vermeulen, G. Weeda, and J. M. Egly. 1994. The DNA-dependent ATPase activity associated with the class-H basic transcription factor BTF2/TFIH. J. Biol. Chem. 269: 9826-9832. Rudolph, H., and A. Hinnen. 1987. The yeast PHOS promoter: phosphate-control elements and sequences mediating mRNA start-site selection. Proc Natl Acad Sci U S A 84: 1340-4. RuizGarcia, A. B., R. Sendra, M. Pamblanco, and V. Tordera. 1997. Gcn5p is involved in the acetylation of histone H3 in nucleosomes. FEBS Letters 403:186-190. Sadovsky, Y., P. Webb, G. Lopez, J. D. Baxter, P. M. Fitzpatrick, E. Gizang- Ginsberg, V. Cavailles, M. G. Parker, and P. J. Kushner. 1995. Transcriptional activators differ in their responses to overexpression of TATA-box-binding protein. Mol. Cell. Biol. 15: 1554-1563. Sainz, M. B., S. A. Goff, and V. L. Chandler. 1997. Extensive mutagenesis of a transcriptional activation domain identifies single hydrophobic and acidic amino acids important for activation in vivo. Mol. Cell. Biol. 17:115-122. Sakurai, H., T. Ohishi, and T. Fukasawa. 1994. Two alternative pathways of transcription initiation in the yeast negative regulatory gene GAL80. Mol Cell Biol 14:6819-28. Saleh, A., V. Lang, R. Cook, and C. J. Brand]. 1997. Identification of native complexes containing the yeast coactivator/repressor protiens NGGl/ADA3 and ADA2. J. Biol. Chem. 272:5571-5578. Sauer, F., J. D. Fondell, Y. Ohkuma, R. G. Roeder, and H. Jackle. 1995. Control of transcription by Kruppel through interactions with TFHB and TFHB beta. Nature 375: 162-164. Sauer, F., S. K. Hansen, and R. Tjian. 1995. Multiple TAFHs direction synergistic activation of transcription. Science 270:1783-1788. Sauer, F., D. A. Wassarman, G. M. Rubin, and R. Tjian. 1996. TAFHs mediate activation of transcription in the Drosophila embryo. Cell 87 : 1271-1284. Sayre, M. H., H. Tschochner, and R. D. Kornberg. 1992. Reconstitution of transcription with five purified initiation factors and RNA polymerase H from Saccharomyces cerevisiae. J Biol Chem 267 :23376-82. Schaeffer, L., R. Roy, S. Humbert, V. Moncollin, W. Vermeulen, J. H. J. Hoeijkmakers, P. Chambon, and J .-M. Egly. 1993. DNA repair helicase: A component of BTF2 (TFHH) basic transcription factor. Science 260:58-63. Schmitz, M. L., M. A. dos Santos Silva, H. Altrnann, M. Czisch, T. A. Holak, and P. A. Baeuerle. 1994. Structural and functional analysis of the NF—kappaB p65 C terminus: an acidic and modular transactivation domain with the potential to adopt an alpha-helical conformation. J. Biol. Chem. 269:25613-25620. 126 305. 306. 307. 308. 309. 310. 311. 312. 313. 314. 315. 316. 317. 318. 319. 320. Serizawa, H., R. C. Conaway, and J. W. Conaway. 1992. A carboxyl-terminal- domain kinase associated with RNA polymerase H transcription factor delta from rat liver. Proc. Natl. Acad. Sci. USA 89:7476-7480. Serizawa, H., R. C. Conaway, and J. W. Conaway. 1993. Multifunctional RNA polymerase H initiation factor delta from rat liver. Relationship between carboxyl- terrninal domain kinase, ATPase, and DNA helicase activities. J Biol Chem 268: 17300-8. Severinov, K., A. Mustaev, A. Kukarin, O. Muzzin, 1. Bass, S. A. Darst, and A. Goldfarb. 1996. Structural modules of the large subunits of RNA polymerase. Introducing archaebacterial and chloroplast split sites in the beta and beta’ subunits of Escherichia coli RNA polymerase. J Biol Chem 271:27969-74. Shaw, S. P., J. Wingfield, M. J. Dorsey, and J. Ma. 1996. Identifying a species- specific region of yeast TFIIB in vivo. Mol. Cell. Biol. 16:3651-3657. Shen, F., S. J. Triezenberg, P. Hensley, D. Porter, and J. R. Knutson. 1996. Transcriptional activation domain of the herpesvirus protein VP16 becomes conforrnationally constrained upon interaction with basal transcription factors. J. Biol. Chem. 271:4827-4837. Shi, X., M. Chang, A. J. Wolf, C. H. Chang, A. A. Frazer Abel, P. A. Wade, Z. F. Burton, and J. A. Jaehning. 1997. Cdc73p and Paf 1p are found in a novel RNA polymerase H-containing complex distinct from the Srbp-containing holoenzyme. Mol Cell Biol 17:1160-9. Shi, X., A. Finkelstein, A. J. Wolf, P. A. Wade, Z. F. Burton, and J. A. Jaehning. 1996. Paf 1p, an RNA polymerase H-associated factor in Saccharomyces cerevisiae, may have both positive and negative roles in transcription. Mol Cell Biol 16:669-76. Shore, D. 1994. RAP]: a protean regulator in yeast. Trends Genet 10:408-12. Shykind, B. M., J. Kim, and P. A. Sharp. 1995. Activation of the TFHD-TFHA complex with HMG-2. Genes Dev 9:1354-65. Sikorski, R. S., and P. Hieter. 1989. A system of shuttle vectors and yeast host strains designed for efficient manipulation of DNA in Saccharomyces cerevisiae. Genetics 122:19-27. Silverman, N., J. Agapite, and L. Guarente. 1994. Yeast ADA2 protein binds to the VP16 protein activation domain and activates transcription. Proc. Natl. Acad. Sci. USA 91:11665-11668. Singer, V. L., C. R. Wobbe, and K. Struhl. 1990. A wide variety of DNA sequences can functionally replace a yeast TATA element for transcriptional activation. Genes Dev 4:636-45. Singh, J., and G. H. Dixon. 1990. High mobility group proteins 1 and 2 function as general class H transcription factors. Biochemistry 29:6295-302. Smale, S. T., and D. Baltimore. 1989. The "initiator" as a transcription control element. Cell 57: 103-13. Smale, S. T., M. C. Schmidt, A. J. Berk, and D. Baltimore. 1990. Transcriptional activation by Spl as directed through TATA or initiator: specific requirement for mammalian transcription factor HD. Proc Nat] Acad Sci U S A 87 :4509-13. Smith, J. S., and J. D. Boeke. 1997. An unusual form of transcriptional silencing in yeast ribosomal DNA. Genes Dev 11:241-54. 127 321. 322. 323. 324. 325. 326. 327. 328. 329. 330. 331. 332. 333. 334. 335. 336. Stelzer, G., A. Goppelt, F. Lottspeich, and M. Meisteremst. 1994. Repression of basal transcription by HMG2 is counteracted by TFIH-I-associated factors in an ATP- dependent process. Mol Cell Biol 14:4712-21. Stern, M., R. Jensen, and I. Herskowitz. 1984. Five SW1 genes are required for expression of the H0 gene in yeast. J Mol Biol 178:853-68. Sterner, D. E., P. A. Grant, S. M. Roberts, L. J. Duggan, R. Belotserkovskaya, L. A. Pacella, F. Winston, J. L. Workman, and S. L. Berger. 1999. Functional organization of the yeast SAGA complex: distinct components involved in structural integrity, nucleosome acetylation, and TATA-binding protein interaction. Mol Cell Biol 19:86-98. Stringer, K. F., C. J. Ingles, and J. Greenblatt. 1990. Direct and selective binding of an acidic transcriptional activation domain to the TATA-box factor TFHD. Nature 345:783-786. Strubin, M., and K. Struhl. 1992. Yeast and human TFHD with altered DNA- binding specificity for TATA elements. Cell 68:721-730. Struhl, K. 1984. Genetic properties and chromatin structure of the yeast gal regulatory element: an enhancer-like sequence. Proc. Natl. Acad. Sci. 81:7865-7869. Struhl, K. 1985. Naturally occurring poly(dA-dT) sequences are upstream promoter elements for constitutive transcription in yeast. Proc Natl Acad Sci U S A 82:8419- 23. Struhl, K. 1995. Yeast transcriptional regulatory mechanisms. Annu Rev Genet 29:651-74. Sullivan, S. M., P. J. Horn, V. A. Olson, A. H. Koop, W. Nu, R. H. Ebright, and S. J. Triezenberg. 1998. Mutational analysis of a transcriptional activation region of the VP16 protein of herpes simplex virus. Nucleic Acids Res. 26:4487-4496. Sun, X., D. Ma, M. Sheldon, K. Yeung, and D. Reinberg. 1994. Reconstitution of human TFHA activity from recombinant polypeptides: a role in TFHA-mediated transcription. Genes Dev 8:2336-2348. Sun, Z., and M. Hampsey. 1996. Synthetic enhancement of a TFHB defect by a mutation in SSU72 an essential yeast gene encoding a novel protein that affects transcription start site selection in vivo. Mol. Cell. Biol. 16: 1557-1566. Sun, Z. W., and M. Hampsey. 1995. Identification of the gene (SSU71/TFG1) encoding the largest subunit of transcription factor TFHF as a suppressor of a TFHB mutation in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 92:3127-31. Sun, Z. W., A. Tessmer, and M. Hampsey. 1996. Functional interaction between TFHB and the pr9 (Ssu73) subunit of RNA polymerase H in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. 24:2560-2566. Taggart, A. K. P., and B. F. Pugh. 1996. Dimerization of TFHD when not bound to DNA. Science 272: 1331-1333. Tantin, D., and M. Carey. 1994. A heteroduplex template circumvents the energetic requirement for ATP during activated transcription by RNA polymerase H. J Biol Chem 269: 17397-400. Thanos, D., and T. Maniatis. 1992. The high mobility group protein HMG I(Y) is required for NF-kappa B-dependent virus induction of the human IFN-beta gene. Cell 71:777-89. 128 337. 338. 339. 340. 341. 342. 343. 344. 345. 346. 347. 348. 349. 350. 351. 352. Thoma, F., T. Koller, and A. Klug. 1979. Involvement of histone H1 in the organization of the nucleosome and of the salt-dependent superstructures of chromatin. J Cell Biol 83:403-27. Thomas, J. O., and V. Furber. 1976. Yeast chromatin structure. FEBS Lett 66:274- 80. Tremethick, D. J., and P. L. Molloy. 1988. Effects of high mobility group proteins 1 and 2 an initiation and elongation of specific transcription by RNA polymerase H in vitro. Nucleic Acids Res 16:11107-23. Triezenberg, S. J. 1995. Structure and function of transcriptional activation domains. Curr. Opin. Genetics Dev. 5: 190-196. Triezenberg, S. J ., R. C. Kingsbury, and S. L. McKnight. 1988. Functional dissection of VP16, the trans-activator of herpes simplex virus immediate early gene expression. Genes Dev. 2:718-729. Tse, C., E. I. Georgieva, A. B. Ruiz Garcia, R. Sendra, and J. C. Hansen. 1998. Gcn5p, a transcription-related histone acetyltransferase, acetylates nucleosomes and folded nucleosomal arrays in the absence of other protein subunits. J Biol Chem 273:32388-92. Tyree, C. M., C. P. George, L. M. Lira-DeVito, S. L. Wampler, M. E. Dahmus, L. Zawel, and J. T. Kadonaga. 1993. Identification of a minimal set of proteins that is sufficient for accurate initiation of transcription by RNA polymerase H. Genes Dev. 7: 1254-1265. Usheva, A., and T. Shenk. 1994. TATA-binding protein-independent initiation: YYl, TFHB, and RNA polymerase H direct basal transcription on supercoiled template DNA. Cell 76:11 15-1121. Van Hoy, M., K. K. Leuther, T. Kodadek, and S. A. Johnston. 1993. The acidic activation domains of the GCN4 and GAL4 proteins are not or helical but form B sheets. Cell 72:587-594. Verrijzer, C. P., J.-L. Chen, K. Yokomori, and R. Tjian. 1995. Binding of TAFs to core elements directs promoter selectivity by RNA polymerase H. Cell 81:1115-1 125. Verrijzer, C. P., and R. Tjian. 1996. TAFs mediate transcriptional activation and promoter selectivity. Trends Biochem Sci 21:338-42. Voegel, J. J ., M. J. Heine, C. Zechel, P. Chambon, and H. Gronemeyer. 1996. TIF2, a 160 kDa transcriptional mediator for the ligand-dependent activation function AF-2 of nuclear receptors. EMBO J 15:3667-75. Wade, P. A., W. Werel, R. C. Fentzke, N. E. Thompson, J. F. Leykam, R. R. Burgess, J. A. Jaehning, and Z. F. Burton. 1996. A novel collection of accessory factors associated with yeast RNA polymerase H. Protein Expr Purif 8:85-90. Walker, 8., R. Greaves, and P. O’Hare. 1993. Transcriptional activation by the acidic domain of Vmw65 requires the integrity of the domain and involves additional determinants distinct from those necessary for TFHB binding. Mol. Cell. Biol. 13:5233-5244. Walker, S. S., J. C. Reese, L. M. Apone, and M. R. Green. 1996. Transcriptional activation in cells lacking TAF(H)s. Nature 383: 185-188. Walker, S. S., W. C. Shen, J. C. Reese, L. M. Apone, and M. R. Green. 1997. Yeast TAF(H)145 required for transcription of G1/S cyclin genes and regulated by the cellular growth state. Cell 90:607-14. 129 353. 354. 355. 356. 357. 358. 359. 360. 361. 362. 363. 364. 365. 366. 367. 368. 369. Wang, L., B. Turcotte, L. Guarente, and S. L. Berger. 1995. The acidic transcriptional activation domains of herpes virus VP16 and yeast HAP4 have different co-factor requirements. Gene 158: 163-170. Wang, W. D., M. Carey, and J. D. Gralla. 1992. Polymerase-H promoter activation - closed complex formation and ATP-driven start site opening. Science 255:450-453. Weis, L., and D. Reinberg. 1992. Transcription by RNA polymerase-H - initiator- directed formation of transcription-competent complexes. FASEB J. 6:3300-3309. West, M. L., and J. L. Corden. 1995. Construction and analysis of yeast RNA polymerase H CTD deletion and substitution mutations. Genetics 140: 1223-33. Wilson, A. C., K. LaMarco, M. G. Peterson, and W. Herr. 1993. The VP16 accessory protein HCF is a family of polypeptides processed from a large precursor protein. Cell 74:] 15-125. Wilson, C. J ., D. M. Chao, A. N. Imbalzano, G. R. Schnitzler, R. E. Kingston, and R. A. Young. 1996. RNA polymerase H holoenzyme contains SWIISNF regulators involved in chromatin remodeling. Cell 84:235-44. Winston, F., and M. Carlson. 1992. Yeast SNF/SW1 transcriptional activtors and the SPT/SIN chromatin connection. Trends in Genetics 8:387-391. Winston, F., C. Dollard, and S. L. Ricupero Hovasse. 1995. Construction of a set of convenient Saccharomyces cerevisiae strains that are isogenic to 8288C. Yeast 11:53-5. Winter, E., and A. Varshavsky. 1989. A DNA binding protein that recognizes oligo(dA).oligo(dT) tracts. EMBO J 8: 1867-77. Wobbe, C. R., and K. Struhl. 1990. Yeast and human TATA-binding proteins have nearly identical DNA sequence requirements for transcription in vitro. Mol Cell Biol 10:3859-67. Woychik, N. A., and R. A. Young. 1994. Exploring RNA polymerase H structure and function, p. 227-242. In R. C. Conaway, and J. W. Conaway (ed.), Transcription: mechanisms and regulation. Raven Press, New York, NY. Xiao, H., J. D. Friesen, and J. T. Lis. 1995. Recruiting TATA-binding protein to a promoter: Transcriptional activation without an upstream activator. Mol. Cell. Biol. 15:5757-5761. Xiao, H., A. Pearson, B. Coulombe, R. Truant, S. Zhang, J. L. Regier, S. J. Triezenberg, D. Reinberg, O. Flores, C. J. Ingles, and J. Greenblatt. 1994. Binding of basal transcription factor TFIIH to the acidic activation domains of VP16 and p53. Mol. Cell. Biol. 14:7013-7024. Xie, X. L., T. Kokubo, S. L. Cohen, U. A. Mirza, A. Hoffmann, B. T. Chait, R. G. Roeder, Y. Nakatani, and S. K. Burley. 1996. Structural similarity between TAFs and the heterotetrameric core of the histone octamer. Nature 380:316-322. Xing, Y. Y., J. D. Fikes, and L. Guartente. 1993. Mutations in yeast HAP2/HAP3 define a hybrid CCAAT box binding domain. EMBO J. 12:4647-4655. Yamashita, S., K. Hisatake, T. Kokubo, K. Doi, R. G. Roeder, M. Horikoshi, and Y. Nakatani. 1993. Transcription factor TFHB sites important for interaction with promoter-bound TFIID. Science 261:463-466. Yang, X.-J., V. V. Ogryzko, J.-I. Nishikawa, B. H. Howard, and Y. Nakatani. 1996. A p300/CBP-associated factor that competes with the adenoviral oncoprotein ElA. Nature 382:319-324. 130 370. 371. 372. 373. 374. 375. 376. Yao, T. P., G. Ku, N. Zhou, R. Scully, and D. M. Livingston. 1996. The nuclear hormone receptor coactivator SRC-l is a specific target of p300. Proc Natl Acad Sci U S A 93:10626-31. Yokomori, K., A. Admon, J. A. Goodrich, J. Chen, and R. Tjian. 1993. Drosophila TFHA-L is processed into two subunits that are associated with the TBP/T AF complex. Genes Dev. 7:2235-2245. Yokomori, K., M. P. Zeidler, J. Chen, C. P. Verrijzer, M. Mlodzik, and R. Tjian. 1994. Drosophila TFHA directs cooperative DNA binding with TBP and mediates transcriptional activation. Genes Dev. 8:2313-2323. Young, R. A. 1991. RNA polymerase-II. Annu. Rev. Biochem. 60:689-715. Zaychikov, E., E. Martin, L. Denissova, M. Kozlov, V. Markovtsov, M. Kashlev, H. Heumann, V. Nikiforov, A. Goldfarb, and A. Mustaev. 1996. Mapping of catalytic residues in the RNA polymerase active center. Science 273:107-9. Zhou, Q., T. G. Boyer, and A. J. Berk. 1993. Factors (TAFs) required for activated transcription interact with TATA box-binding protein conserved core domain. Genes Dev. 7:180-187. Zhu, A., and M. A. Kuziora. 1996. Homeodomain interaction with the beta subunit of the general transcription factor TFHB. J Biol Chem 271:20993-6. 131 HICHIGAN STATE UNIV. LIBRRRIES 11111111 11111111111111111 11111111 11111111111111 31293018200224