.fi‘ 61.2.11, . . . $332.», :53... “it? . . . . .1. it... .3. .r . . . . . . 3:55.51?“ 5 x: s.» 27...; . .5 .. . 1.3mm} . , 1.2:... i .1 {£311} {a . . . . . . .5 :33. ER» . . Airfflflomx. . «use £32.: . . . . .w... ”nun“... a. 2. x: . ‘33.:1.‘ . . $5.... . S .3: .9. . . i a. .531... . . .3. . 55....5. 2.}. . .. _ . ‘ .Elxfiwi: 2.: , :1... 3...: 5.. M 5 v in. I ~ i .213... . . (iv: givfiv 5.1.. : .5 x .L. .5 . {$52. . s E... c . . . .LE‘iuiukrfliitsc. \L {.3 gnu»... s... . . 15.5.: a . {$535.5 .313511 . .i. .1. . a}: . . . 35.51.... 324'. 35.9 .. 5:55:12}: . w §v|pllllohsplai \~kIJ‘~ULSF§- 4 : . 1411-332: .gnip-lnfik . Eff-35:5, rlou';§-;\.: - [‘530: . fl .i-La . nu {.(Jir? .1 . 1.... .1Es..:..i.i 5.15. 35.. 33:5... . . “a! 5x1 :§c?x.€i1‘»i~lx§.}!.ts n 5: .3...&. 5:55:55: 5: .nnfiqrxn..usz 1. . . . .3 i 5 , .. .5. whirl: 1‘ c is x s . PF? 1.3%.}: SE 3. 3...;ti . Si... .: 3:15... 3 x r: . EZSG: 4.5"“ 3, $55.5: 5 : _ 15...... . 1:133:21xii4. @4453“. n-1§~1ri~l v‘thl .ril. ti %2\ 5 g on 5:}..de ,1.._E..“..i{krijitrngsfi .drv . 33.}. 6.1311. 111 1:2. . , {munclgtélflwfinh : . .1S5’....«vluvuq§» w ai}..i¢§ Ext! 3.....{155275 .E.s..«:.:.$i.za§.. .5. . E». t\-l!\og. its< Spicaifir Y: _..1.\;$:2Lcnu.:3. 2.53.!) z :33.) 133’. {5:5 Fifi”; [itivuul‘ t5. .3} .<¥(k§.: . '3‘ I»)! 5.5. i. . 3.3.5:, 2.: 55:23... 122...: 54.55. 4...... :5....i 255%; . a“: Kan». fines». . 5.. (Lt . .. . . . il.....vdlr.v.. , . .3252... . L s p 3!...‘0‘11‘ 5.215»). Kira/{(7. ‘ . . c.5ié..un. . ’23)! . . ,4. 5.3.21 5 .5 555.1}. p24,: 1 t. 2 .( 3:. .12.)... ..)_§~:.)Z. . v J. n .555. 55.5? Ctnxéxfi‘. .. 5.3.1.1131!) 22‘} .{rfgti’xvlrlfilifz . f7. . 3:115:73 71...... E i; I. lllllllllllllllllll This is to certify that the dissertation entitled Structural and Functional Comparison of the Transcriptional Activating Domains of VP16 and RELA presented by Peter J. Horn has been accepted towards fulfillment of the requirements for Ph.D. degree in Biochemistry Maj/or {)rofeéxir Or Date 4 7 y MSUI'J un Aflirmuliw Anion/Equal Opportunity lnrmmmn 042771 LIBRARY __ Miehigan State University STRUCTURAL AND FUNCTIONAL COMPARISON OF THE TRANSCRIPTIONAL ACTIVATING DOMAINS OF VP16 AND RELA By Peter J. Horn A DISSERTATION Submitted to Michigan State University in partial fiilflllment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Biochemistry 1999 ABSTRACT STRUCTURAL AND FUNCTIONAL COMPARISON OF THE TRANSCRIPTIONAL ACTIVATING DOMAINS OF VP16 AND RELA By Peter J. Horn The activation domains of many eukaryotic transcriptional activator proteins are modular in nature. These domains are constructed of multimerized motifs, each independently capable of directing activation. The activation domain of the herpes Simplex virus activator, VP16, consists of two such motifs or "subdomains". Each is sufficient to independently enhance transcription in vivo and in vitro. Similarly, the activation domain‘of the mammalian activator RelA is separable into three independent subdomains. For both activators, the transcriptional activity Of each subdomain is critically dependent upon a Short cluster of hydrophobic residues. These Observations raise at least two pertinent questions. First, what primary structural characteristics contribute to subdomain fimction? Second, does each subdomain function identically, or do different subdomains contribute to activation through distinct mechanisms? To evaluate these questions, both subdomains of VP16 (VP16N and VP16C) and two Of the three subdomains OfRelA (RelAII and RelAIH) were examined by a combination of site-directed and/or random mutational strategies. While each subdomain was found to be dependent upon short hydrophobic segments for function, subtle structural distinctions were observed. The VP16 subdomains each consist of a central, critical F(D/E) dipeptide flanked by hydrophobic residues of variable importance. The precise spacing of the hydrophobic residues within the primary structure is distinct for each subdomain. In contrast to either VP16 subdomain, the RelA subdomains are dependent only upon hydrophobic residues for function. Both RelAII and RelAIII exhibited a pattern of critical hydrophobic residues similar to that present in VP16C. Despite some common structural properties, several results argue that the VP16 subdomains are not functionally identical. VP16N, and not VP16C, stimulated an integrated reporter gene. VP16C efficiently activated a promoter containing an Initiator (Inr) element, but not a promoter containing a TATA element, whereas VP16N directed equal activation of both promoters. Finally, the VP16C subdomain, but not the VP16N subdomain, stimulated formation of a TFIID-TFIIA-DNA (DA) complex. Similar experiments showed that the RelAII and RelAIII subdomains, while effectively activating only an Inr-containing promoter, did not enhance DA complex formation, indicating that these two activities are unlikely to be functionally related. These results demonstrate that the VP16 subdomains fiinction through distinct mechanisms. However, both subdomains have similar structural properties. The VP16N and VP16C subdomains differ only in the spacing of hydrophobic residues within their core sequences. To test whether this Spacing is critical for subdomain-specific function, the spacing of the hydrophobic residues within VP16N, was altered to resemble the spacing observed in VP16C. This VP16N mutant was incapable of stimulating DA complex formation, suggesting that the spacing Of the core hydrophobic residues within the VP16 subdomains is insufficient to confer subdomain-specific function. Residues external to these core motifs must provide functional specificity. to Ellen ACKNOWLEDGMENTS I would like to thank my committee members for their support and assistance. Their time and effort is greatly appreciated. I would also like to thank Steve for his encouragement and patience throughout my time in his lab. His efforts to allow me to develop independently have been appreciated. I would also like tO acknowledge members of the Triezenberg lab past and present, for their companionship, friendship, and scientific help and advice. Iwould also like to thank Naoko Kobayashi and Arnie Berk for their stimulating collaborations and kind hospitality during my brief time at UCLA. Most importantly, I would like to thank my wife, family and friends. They have provided a constant source of support throughout the "ups-and downs" Of scientific research. Without their help, none of this would have been achievable. ACKNOWLEDGMENTS I would like to thank my committee members for their support and assistance. Their time and effort is greatly appreciated. I would also like to thank Steve for his encouragement and patience throughout my time in his lab. His efforts to allow me to develop independently have been appreciated. I would also like to acknowledge members of the Triezenberg lab past and present, for their companionship, friendship, and scientific help and advice. I would also like to thank Naoko Kobayashi and Arnie Berk for their stimulating collaborations and kind hospitality during my brief time at UCLA. Most importantly, I would like to thank my wife, family and friends. They have provided a constant source Of support throughout the "ups-and downs" of scientific research. Without their help, none of this would have been achievable. TABLE OF CONTENTS List of Figures .............................................................................................................. viii List of Symbols and Abbreviations ................................................................................. ix Chapter One Introduction and Literature Review ............................................................................ 1-50 Stepwise Model for Pre-initiation Complex Formation ......................................... 7 Holoenzyme Model of Transcription Initiation ..................................................... 9 Assembly of Pre-initiation Complexes on Inr Promoters .................................... 11 TBP binding constitutes the first regulated step in PIC formation ....................... 12 TBP-associated factors may govern activator response ....................................... 21 TFIIB recruitment may stimulate PIC formation ................................................ 31 The basal transcription factor TFHA is required for activator response ............... 36 Activation may also involve non-basal auxiliary factors ..................................... 41 Other complexes are also potential coactivator candidates .................................. 46 Potential models for the function of activators ................................................... 48 Chapter Two Activation Domins Contain Similar Yet Distinct Hydrophobic Motifs .................... 51-87 Introduction ....................................................................................................... 51 Experimental Results ......................................................................................... 55 The activation domains of VP16 are dependent upon similar hydrophobic motifs ..................................................................... 55 A related “acidic” activator, RelA, also contains similar motifs .............. 64 Discussion and Conclusions ............................................................................... 75 Experimental Methods ....................................................................................... 8O Recombinant DNA manipulations .......................................................... 80 PCR mutagenesis and screening of VP16C mutants ................................ 84 Mouse L cell transfection ....................................................................... 85 Chloramphenicol Acetyltransferase (CAT) Assays ................................. 86 Chapter Three Distinct Activation Domains Function through Distinct Mechanisms ..................... 88-140 Introduction ....................................................................................................... 88 Experimental Results ......................................................................................... 90 VP16N and not VP16C activates transcription from an integrated reporter ....................................................................................... 90 The VP16C activation domain is sensitive to basal promoter structure 93 The VP16C subdomain recruits TFIID and TFHA to a promoter in vitro ........................................................................................ 99 Expression Of VP16N or VP16C activation domains inhibits transcription in vivo .................................................................. 104 Are the RelA subdomains mechanistically similar to VP16C? .............. 111 The RelA activation domain displays Inr promoter preference .............. 111 RelA does not enhance DA recruitment to a promoter in vitro .............. 112 Can alteration of the hydrophobic core of VP16N reconstitute VP16C activity? ........................................................................ 1 13 Discussion and Conclusions ............................................................................. 118 Experimental Methods ..................................................................................... 124 Recombinant DNA Manipulations ........................................................ 124 Tissue Culture Cell Manipulation ......................................................... 126 RNA Isolation ...................................................................................... 127 Primer Extension Analysis ................................................................... 128 Preparation of Gal4-Fusion Proteins ..................................................... 130 Preparation of eTFIID .......................................................................... 132 Preparation of His—tagged TFIIA .......................................................... 136 Preparation Of EMSA Probes ................................................................ 138 Gal4 EMSA Assay ............................................................................... 138 TFHA-TBP EMSA Assay ..................................................................... 139 DA Complex EMSA Assay .................................................................. 140 Chapter Four Activation domains are composed of multiple, small modular elements ........... 141 Activator-target protein-protein complexes appear predominantly hydrophobic ......................................................................................... 142 Modularity of activation domain structure ........................................................ 148 Bibliography ......................................................................................................... 159-184 vii LIST OF FIGURES Figure 1. VP16C mutations identified by genetic selection ............................................ 58 Figure 2. Relative activity of VP16N alanine substitutions in the activation domain core region ......................................................................................................... 62 Figure 3. RelA Activation Domain Alignment ............................................................... 67 Figure 4. Comparison of VP16 and RelA sequences ...................................................... 68 Figure 5. Relative activity OfRelAII substitutions in the activation domain core region ................................................................................................................ 70 Figure 6. Relative activity of RelAIII substitutions in the activation domain core region ................................................................................................................ 72 Figure 7. Activity Of VP16 activation domains on an integrated reporter ....................... 92 Figure 8. Promoter Specificity Of individual VP16 and RelA subdomains ...................... 96 Figure 9. TFIIA stimulates TBP recognition of the AdML promoter TATA box ............ 98 Figure 10. Saturation Of the EMSA probe with Gal-AD fusion proteins ....................... 102 Figure 11. Comparison of the capacity of VP16N, VP16C and a mutant VP16N subdomain to enhance DA complex formation ................................................. 103 Figure 12. Gal4-Vpl6C stimulates formation Of a TFIID-promoter complex through recruitment of TFHA to DNA ............................................................. 105 Figure 13. In vivo competition experiments with NLS-VP16 and Gal4-VP16 expression vectors ............................................................................................ 108 Figure 14. The RelA activation domain does not stimulate DA complex formation ...... 115 Figure 15. Deletion of VP16N amino acids D441 and L444 produces a VP16C-like core motif ........................................................................................................ 116 Figure 16. A VP16N mutant, VP16NAD/L, does not stimulate DA complex formation ......................................................................................................... 117 viii AdML bp CAT EMSA HCA HEPES HSV IPTG kDa OD PAGE PBS RNAP SDS-PAGE TAF TFII tk VP16 LIST OF ABBREVIATIONS activation domain adenovirus major late base pair chloramphenicol acetyltransferase electrophoretic mobility shift assay hydrophobic cluster analysis N-2-hydroxyethylpiperiazine-N'-2-ethanesulfonic acid herpes simplex virus isopropyl-B-D-thiolgalactopyranoside kilODaltons optical density polyacrylamide gel electrophoresis phosphate buffered saline RNA polymerase sodium-dodecyl-sulfate polyacrylamide gel electrophoresis TBP associated factor transcription factor, RNA polymerase H thymidine kinase virion protein 16 ix Chapter One Introduction and Literature Review Accurate control of gene expression is essential for a wide variety Of biological processes including cell growth and development, tissue specialization, metabolic adaptation, immune response, angiogenesis and oncogenesis. As such, a thorough understanding of the fiindarnental mechanisms governing this process is a central issue of current biological research. Regulation of gene expression occurs through a variety of transcriptional and post-transcriptional mechanisms. Transcriptional initiation and elongation, differential mRNA splicing, mRNA stability and transport, translational efiiciency, and poSt-translational protein modifications all modulate gene expression. Transcriptional regulation of gene expression has been one Of the most extensively studied aspects of gene regulation. Unlike prokaryotic systems, the task of RNA production in eukaryotic nuclei is divided between three polymerases. Ribosomal RNAS are transcribed by RNA polymerase I. Transcription of protein—coding genes is performed by RNA polymerase H (RNAPII), while RNA polymerase IH transcribes most small RNAS (some snRNAs are actually RNAPH transcribed). While prokaryotic RNA polymerases are capable of promoter recognition in the absence of additional factors, (the prokaryotic promoter recognition factor, the polymerase 0' subunit, is a component of the polymerase itself), the eukaryotic RNA polymerases require auxiliary factors for promoter recognition. These general transcription factors (GTFs, also called basal transcription factors) are required to coordinate recruitment of RNAPH to the promoter. The RNAPH GTF proteins are thought to assemble in a sequential fashion into a promoter associated complex, referred to as the pre-initiation complex (PIC). This PIC comprises the GTFs TBP (the TATA binding protein of TFHD), TFHA, TFIIB, TFIIE, TFIIF, and TFIIH (Zawel, L. and Reinberg, D., 1993; Orphanides, G. etal., 1996; Conaway, RC. and Conaway, J .W., 1993). The assembly of the PIC is described in more detail in following sections. Regulation of transcription from RNAPII promoters requires an additional set of proteins, often referred to as specific transcription factors. These factors comprise two general groups: transcriptional activators and transcriptional repressors (Hahn, S., 1993; Triezenberg, S.J., 1995; Hanna-Rose, W. and Hansen, U., 1996). Eukaryotic activators, the major topic of this text, are generally bipartite in structure, composed of two independent functiOnal domains (Ptashne, M., 1988; Mitchell, R]. and Tjian, R, 1989). One domain is essential for promoter recognition (either through direct protein-DNA or protein-protein interaction with additional DNA-binding factors), while the second domain fimctions to activate transcription. Artificial chimeric activators composed of firsions of a DNA-binding domain and an activation domain are functional in vitro (Carey, M. et al., 1990; Lin, Y.S. et al., 1988) and in vivo (Sadowski, I. et al., 1988; Walker, S. et al., 1993; Seipel, K. et al., 1992). For example, a fusion of the DNA-binding domain of the yeast activator Gal4 with the herpes simplex virus (HSV) transactivator VP16 produces an unusually potent transcriptional activator (Sadowski, I. et al., 1988). This synthetic activator has been utilized as a prototype in a variety of studies, many of which are reviewed in this text. While the structural features essential for activation domain function remain somewhat poorly understood, their function is clearly dependent upon the hydrophobic arnino acids Of these domains (Triezenberg, S.J., 1995; Hahn, S., 1993). Early experiments with the yeast activator GAL4 suggested that acidic amino acids in one of two GAL4 activation domains were of principal importance to transcriptional activity (Ma, J. and Ptashne, M., 1987; Gill, G. and Ptashne, M., 1987). Deletion analysis of another yeast activator GCN4 also localized the activation domain to a large region of extensive acidic character (Hope, LA. and Struhl, K., 1986). These early observations led to the supposition that the sole structural determinant for activation domain function was a primary structure rich in acidic residues (Hope, IA. and K. Struhl, 1986). Characterization of other transcriptional activators has also resulted in the identification of activation domains rich in glutamine, proline, or isoleucine residues. This led to a proposal for classification of activator classes defined by the predominance of these residues in activation domain sequences (Mitchell, RI. and R. Tjian, 1989). However, studies using the acidic HSV activator VP16 identified hydrophobic residues, and not acidic residues, as critical for transactivation (Cress, W.D. and Triezenberg, S.J., 1991). In addition, more refined analyses of the activation domain of the yeast activator GCN4 by site-directed mutagenesis identified hydrophobic amino acids within this region as the critical residues for transactivation (Drysdale, C.M. et al., 1995; Jackson, B.M. et al., 1996). Other prototypical "acidic" activators also display similar characteristics. The mammalian tumor suppressor p53 requires hydrophobic residues within the N-terminal transactivation domain for activation (Lin, J. et al., 1994), and the NF-KB family member RelA requires multiple hydrophobic residues for function (Schmitz, ML. et al., 1994; Schmitz, M. et al., 1995; Blair, W.S. et al., 1994b). Unrelated (“non- acidic”) activators such as the glutamine-rich activators Spl (Gill, G. et al., 1994) and Oct-1 (Tanaka, M. and Herr, W., 1994) also require hydrophobic residues for function. Classification of activator subtypes merely by amino acid content may not be particularly meaningful. Instead, classification based upon activator function might be a more useful method. Several studies suggest that activators may function through distinct mechanisms. The Schaffner laboratory compared the ability of various Gal4-activation domain firsions to firnction from cis elements distal (enhancer) and proximal (immediately 5’ of TATA sequences) to an eukaryotic promoter. While some activators (e. g. VP16) were insensitive to promoter positioning, others (6. g. Spl) required a proximal location for efficient function. All activation domains examined functioned from promoter proximal positions (Seipel, K. et al., 1992). Similar results were also observed in yeast (Kunzler, M. et al., 1994). Other Observations also suggest that different activators function through distinct mechanisms. Studies Of several activators by the Bentley (Blau, J. et al., 1996; Yankulov, K. et al., 1994) and Groudine (Krumm, A. et al., 1995) laboratories established that polymerase elongation constitutes a rate—limiting step that some, but not all activators stimulate. Blau et al. proposed classification of activators into three groups based upon their ability to stimulate initiation (e. g. Spl), elongation (e. g. Tat) or both (e. g. VP16, HSF) events. In vitro and in vivo studies by the Smale laboratory identified differential sensitivity of activators to basal promoter architecture, again suggesting that activators may enhance transcription through distinct mechanisms. The VP16N activation domain (see below) demonstrated no dependence upon basal promoter structure. In contrast, the Spl activation domain was highly dependent upon initiator sequences for efficient function (Emami, K.H. et al., 1995). Difl‘erent activation domains also display distinct functions in vitro. TAFs are required for activated transcription in vitro, and when assembled with TBP, comprise the transcription factor TFIID (reviewed in Burley, SK. and Roeder, KG, 1996). TBP is essential for transcription by all three RNA polymerases (Hernandez, N., 1993). Specific TAFs are likely to provide specificity to TBP complexes in vivo, as distinct TBP-TAF complexes have been isolated (Lee, TI. and Young, RA, 1998). TFIID subcomplexes containing partial TAF complexes are also differentially responsive to difierent activators, consistent with the idea that different activators target distinct TAFs for recruitment of TFIID. Distinct partial T AF complexes mediate for NTF-l and Spl function in vitro (Chen, J. et al., 1994). The VP16C subdomain also interacts with the human TAFS hTAFn32 and hTAFu7O (Klemrn, R. et al., 1995), as well as the Drosophila homologue dTAFn40 (Goodrich, J.A. et al., 1993). These interactions are specific for VP16C. The VP16N subdomain does not bind either Of these TAFS. The Epstein-Barr virus activator Zta and HSV VP16 also stimulate formation of a TFIID-TFIIA-promoter (DA) complex in vitro, as assessed by an agarose EMSA technique (Lieberman, RM. and Berk, A.J., 1994; Kobayashi, N. et al., 1995). The C-terminal half of the VP16 activation domain (VP16C, amino acids 456-490) stimulates formation of this complex, while the N-terminal half (VP16N, amino acids 410-456) does not stimulate formation of this complex (Kobayashi, N. et al., 1995, see below). This activity has not been observed in any other activators. This result suggests that even separate domains within a single activator may function through distinct mechanisms. Finally, a variety of studies have observed differential efl‘ects of reporter chromosomal integration on activator firnction in vivo, suggesting a differential capacity of activators to antagonize the repressive effects of chromatin organization (reviewed in Smith, CL. and Hager, G.L., 1997). The precise mechanism of function of any specific activator has remained elusive despite a great deal of experimental effort by a large number of investigators. Conceptually, activators may function at any step in transcription. Chromatin structure is commonly regarded to be inhibitory to transcription. Acetylation of nucleOsomal histones correlates with enhancement of gene expression. Derepression of chromatin packaged genes may be facilitated by recruitment of histone acetyltransferases such as the GCNS/ SAGA or other acetyltransferase complexes (reviewed in Struhl, K., 1998). Activators may also recruit chromatin remodeling complexes such as SWI/SNF or RSC to genes (reviewed in Kingston, RB. et al., 1996). These complexes facilitate access Of nucleosomal DNA to DNA binding proteins and may be required prior to activator and PIC recognition of regulatory or promoter DNA sequences. Original models for activator function largely involved stimulation Of either the rate of formation or stability of the PIC. Consistent with these proposals, protein-protein interactions between activation domains and GTFs have been observed (reviewed in Roeder, KG, 1996). Activators may also induce isomerization of the PIC or a partial PIC to an activated state, essentially acting as allosteric effectors of PIC activity (Chi, T. and Carey, M., 1996). Conceptually, activators could also enhance promoter escape (Kumar, K.P. et al., 1998; Goodrich, IA. and Tjian, R., 1994), facilitating the conversion of the initiation complex to an elongation complex and finally, activators may function through stimulation of elongation (Bentley, BL, 1995). Any or all of these mechanisms may be Operative. The following sections will focus on a brief description Of the initiation- elongation cycle followed by a more detailed survey Of the potential cross talk between activators and specific GTFS or regulatory complexes. Stepwise Model for Pro-initiation Complex Formation In vitro studies utilizing purified and recombinant basal transcription factors have led to the proposal of a stepwise model (Buratowski, S. et al., 1989; reviewed in Roeder, R. G., 1996; Conaway, RC. and J.W. Conaway, 1993) for formation of an RNAPH pre- initiation complex (PIC). These studies have largely been confined to a subset of promoters, primarily strong viral promoters containing either consensus or near consensus TATA elements (i.e. the adenovirus Elb, major late, and E4 promoters). Recent studies have led to an appreciation of the complexity of eukaryotic promoters (reviewed in Smale, ST, 1997). Eukaryotic promoters appear to be composite in nature. In addition to well recognizable TATA sequences, sequences within the initiation site (initiator, or Inr element) (Oshea-Greenfield, A. and Smale, ST, 1992) as well as downstream of the initiation site (DPE; downstream promoter element) contribute to promoter firnction and strength (Burke, T.W. and Kadonaga, J .T., 1996). Thus, current models, which have been elucidated through study of strong viral promoters may be representative Of events at only a subset of eukaryotic promoters. A limited number of studies have been performed with TATA-less (Inr- containing) promoters, and some data suggest that PIC formation pathways may be distinct for TATA and Inr promoters (for a review of Inr-mediated PIC assembly, see Smale, ST, 1997). Pre-initiation complex formation on the adenovirus major late promoter (AdML promoter) initiates with recognition of the TATA box by TFIID. The TFIID complex is composed Of the TATA binding protein (TBP) and a collection Of 8-12 additional polypeptides (TAFs, TBP associated factors). Recognition of the promoter by TFIID (Lieberman, PM. and A]. Berk, 1994; Ozer, J. et al., 1998) or TBP (Cortes, P. et al., 1992; Imbalzano, A.N. et al., 1994b) is facilitated in vitro under certain experimental conditions by the auxiliary transcription factor TFHA. This TFIID-TFIIA-TATA ("DA" complex) can recruit TFIIB to form a "DAB" complex, which in turn recruits RNAPII. Recruitment of RNAPH is thought to occur through the association of TFIIF to the DAB complex. Interestingly, TFHB can be isolated in a RNAPH complex and TFHF is tightly associated with this RNAPII complex. These observations suggest a model whereby RNAPII recruitment is facilitated by TFIIB and TFIIF subunits of a larger RNAPII complex (see Conaway, RC. and J .W. Conaway, 1993; an early proposal of the competing holoenzyrne model for PIC formation). Association Of TFHB and TFIIH with the promoter bound RNAPII complex completes PIC formation, facilitating the release of the promoter bound complex. Formation of the PIC is insufficient for transcription initiation. Efficient transcription of RN APII is unique among eukaryotic RNA polymerases in its requirement for hydrolysis of the [3-7 bond of ATP (Bunick, D. et al., 1982). Transition to an initiation complex may require ATP hydrolysis to facilitate opening of the DNA helix at the start site of transcription, as evidenced by the increased reactivity of promoter sequences with the single-stranded DNA specific reagent KMnO4 (Wang, W.D. et al., 1992). Formation of this "open complex" is followed by a period of abortive initiation, characterized by the production and release of short RNA molecules (Bunick, D. et al., 1982). Tjian and coworkers argue that release from this state, in a TFIIB and TFIIH dependent fashiOn, completes the initiation cycle (Goodrich, J .A. and R. Tjian, 1994). TFIIF may also be involved in this process. The RAP74 subunit is essential for assembly of the PIC but also plays an essential role in early elongation (Chang, C. et al., 1993). Release ofRNAPII correlates with the hyperphosphorylation Of a C-terminal domain (CTD) of the largest subunit ofRNAPII (Payne, J .M. et al., 1989; Layboum, PI. and Dahmus, ME, 1990). The activity Of this elongation complex can be influenced by a number Of additional transcriptional elongation factors including TFIIS, elongin and TFIIF (for review, see Aso, T. et al., 1995). Some activators are also capable of stimulating elongation (Yankulov, K. et al., 1994). Holoengme Model of Transcription Initiation A competing model for initiation complex formation has gained considerable attention in the last three to five years (Koleske, A]. and Young, RA, 1995; Orphanides, G. et al., 1996). AS mentioned above, early Observations suggested the possibility that the RNAPH machinery may exist in a larger complex or "holoenzyme". Genetic and biochemical studies Of the large subunit Of RNAPII by Young and coworkers provided convincing evidence for the existence of a holoenzyme in yeast. Young and colleagues were studying the role of the CTD of RNAPH. Deletion of the CTD results in cell death and an allele with a synthetic truncation of half of the CTD produces a conditional lethal phenotype. Young and coworkers isolated suppressor mutations of this phenotype in what they have called SRB genes (Suppressor ofRNA polymerase B). The proteins encoded by these genes are associated with RN APH in cell-free extracts (Thompson, C.M. et al., 1993), although only a subset of RNAPH appears to be present in these holoenzyme complexes. Young and coworkers have estimated that 2-4,000 SRB molecules exist per yeast cell, while RNAPII numbers 10-20,000 molecules (see Lee, TI. and RA. Young, 1998). Kornberg and coworkers simultaneously identified a high molecular weight complex that associates with the CTD and is sufficient to confer transcriptional response to activators in vitro (Kelleher III, R]. et al., 1990). This "mediator" complex contains the SRB proteins (Kim, Y.-J. et al., 1994). Characterization of the Young holoenzyme has verified that initiation factors TFIIB, TFIIF, TFIIH, SRBS and the previously identified SUGI and GAL]! gene products are present in this complex (see Koleske, A]. and RA. Young, 1995). Since quantitation of cellular levels of proteins by comparative Western blotting of RNAPH and SRB proteins has determined that only a portion of RNAPH (20-40%) is present in holoenzyme complexes, it is possible that a subset of RNAPII (perhaps a specialized species), is present in a pre-assembled initiation complex lacking only TFIID and TFIIA. These results also suggest that RNAPII recognition of a promoter can occur in a two-step fashion: TFIID recognition of the promoter followed by RNAPII holoenzyme recognition of TFIID. Interestingly, the holoenzyme supports activated transcription in vitro (Koleske, Al and Young, RA, 1994). This observation has been confirmed using analogous metazoan preparations (Maldonado, E. et al., 1996). In fact, the mediator activity isolated by Kornberg and colleagues was originally identified as a column fraction capable of restoring activator response to core RNAPH in the presence of TBP (Kelleher III, R]. et al., 1990). These results seem to suggest that TAF presence is unnecessary for activator 10 response in vitro. This observation is contrary to several other in vitro analyses (see below; TAF section). Assembly of Pre-initiation complexes on Inr promoters In contrast to the better understood events occurring on TATA containing promoters, the process of PIC formation on TATA-less promoters is less well understood. Comparatively little work has been performed on Inr containing promoters, and most studies of Inr function have been performed using TATA-containing promoters that also contain Inr elements (for instance the viral AdML promoter). Proposals for PIC assembly on Inr containing promoters have suggested an analogous but distinct step-wise pathway to that elucidated for TATA containing promoters (reviewed in Smale, ST, 1997). Work by Smale and coworkers suggests that TATA-less promoter firnction requires the same ensemble of GTFS, with one notable exception. TATA-less basal transcription requires TAFS, whereas basal TATA fiinction occurs with TBP (Kaufinann, J. and Smale, ST, 1994; Verrijzer, C.P. et al., 1995; Martinez, E. et al., 1994). Since Inr promoter recognition cannot occur through TATA sequences, this result suggests that the TAFs themselves may function in promoter recognition. Roeder and colleagues also Obtained evidence that Inr-function may require factors in addition to TAFs for efficient TAT A-less transcription (Martinez, E. et al., 1998a). The identity of these “TIC” factors remains unknown. Other models for promoter recognition include recognition of the Inr element by Inr-binding proteins, a number of which have been proposed. TFII-I (Roy, AL. et al., 1993), YYl (Usheva, A. and Shenk, T., 1994), and even RNAPII itself (Carcamo, J. et al., 1991) have all been 11 demonstrated to recognize Inr-like elements. In the near future, a clearer picture of PIC formation on TATA-less promoters is likely to emerge. TBP binding constitutes the first regulated step in PIC formation The first step in PIC formation on the majority Of promoters is recognition of the TATA box by the TATA- Binding Protein (TBP). The TBP protein is composed of two domains. The N-terminal domain is less well conserved among eukaryotic species. The C-terminal domain (Often referred to as the "core" domain) is highly conserved in all eukaryotic species analyzed (Hernandez, N., 1993). The DNA sequence of this domain is composed of two direct repeats, suggesting that TBP may have evolved by gene fusion of two monomeric ancestors (Yamamoto, T. et al., 1992). Crystal structures of the core domain of the yeast (Kim, Y.C. et al., 1993), Arabidopsis (Nikolov, D.B. et al., 1992), and later human (Nikolov, D.B. et al., 1996) TBP molecules reinforced this conjecture. The structure of the molecule resembles a saddle, with a convex upper surface and a lower concave surface Of sufficient size to accommodate the DNA helix. The molecule displays two-fold pseudo-symmetry consistent with the sequence similarity of the two repeats. Each repeat folds into a twisted beta sheet, topped by a long C-terrninal alpha helix. The first and second beta strands are interrupted by a short alpha helix, producing a loop between the adjoining strands forming a "stirrup" on either side of the molecule. These stirrups are involved in protein-protein interactions of TBP with TFHA and TFIIB. Although a consensus TATA sequence has been identified, DNA sequences recognized by TBP are relatively degenerate. Even non-TATA sequences at TATA positions in apparently TATA-less promoters contact TBP in some cases (see Martinez, E. 12 et al., 1994 for discussion). The crystal structure Of TBP with DNA suggests a possible molecular basis for this promiscuity (Kim, Y.C. et al., 1993). The concave underside of the TBP molecule, composed of the twisted beta sheets Of the two repeats, binds the TATA motif, producing an extensive distortion of the DNA helix evidenced as a broad widening of the minor groove and introduction Of a severe kink in the path of the DNA helix. Since few sequence Specific contacts were Observed, Sigler proposed that the capacity of a DNA sequence to accommodate this conformation determines the quality of a TATA motif (Kim, Y.C. et al., 1993). The general transcription factors TFHA and TFIIB associate with TBP and DNA, contributing additional protein-DNA contacts likely to contribute significantly to PIC stability. This is apparent fiom the results Of structural studies from the Sigler and Burley laboratories that provide a glimpse of the early stages of PIC assembly. The structures of both a TBP-TFHB-DNA (Nikolov, D.B. et al., 1995) and TBP-TFIIA—DNA (Geiger, J .H. et al., 1996) complex have been solved. Superposition of these complexes has provided a high-resolution model for the TBP- TFIIA-TFHB-DNA complex (Nikolov, DB. and Burley, SK, 1997; Geiger, J .H. et al., 1996). The two GTFS bind Opposite faces of the TBP-DNA complex, making direct protein-protein contacts with TBP. TFIIA binds to TBP near the N-terminal stirrup, associating with DNA upstream of the TATA box, while TFHB clamps the C-terminal stirrup. The TFIIB N-terminus is oriented toward the start site consistent with genetic studies in yeast that have implicated TFIIB in start site selection (Pinto, I. et al., 1992). The mammalian transcription factor TFIID was originally identified as a component Of a chromatographic fraction isolated fiom HeLa nuclear extracts (Reinberg, D. et al., 1987). This activity was demonstrated to protect the promoter region of the 13 AdML promoter, including the TATA box and downstream sites in the vicinity of the initiation site (Horikoshi, M. et al., 1988b). Shortly thereafter, a TATA-binding protein was identified in Saccharomyces cerevisiae, purified to homogeneity and cloned (Cavallini, B. et al., 1989; Hahn, S. et al., 1989; Horikoshi, M. et al., 1989; Schmidt, MC. et al., 1989). Identification of this gene rapidly facilitated the identification ofDrosophila (Hoey, T. et al., 1990; Muhich, ML. et al., 1990) and mammalian TFIID (Homnann, A. et al., 1990; Kao, C.C. et al., 1990; Peterson, MC. et al., 1990). Metazoan TFIID was subsequently identified as a high molecular weight complex of TBP and TAFs (some 8-12 polypeptides dependent upon isolation method and species; reviewed in Tansey, WP. and Herr, W., 1997b; Lee, TI. and RA. Young, 1998). The role of TAFs in PIC formation and transcriptional regulation will be reviewed separately below. Two TBP genes have been identified in Arabidopsis (Gasch, A. et al., 1990). Gene duplication in plants is a relatively common phenomenon, suggesting that the existence of multiple TBP proteins may be unique to this kingdom. However, a second TBP-like gene has also been identified in Drosophila (Crowley, T.E. et al., 1993). This related gene encodes a TBP variant with extensive C-terrninal sequence identity and nonconserved N-terminal domain. The second TBP gene is expressed in neuronal tissues and appears to be involved in tissue-specific transcription. Early efforts to substitute this TBP variant for the ubiquitously expressed basal TBP suggested that the variant was not capable Of substituting as a GTF. However, Tjian and colleagues recently reported that this variant TBP is capable of firnctioning as a GTF. The TBP-like protein appears to be associated with a distinct set of TAFs, suggesting that tissue-specific transcription in Drosophila neuronal tissue is mediated by a tissue-specific TFIID complex (Hansen, S.K. et al., 1997). 14 TBP has been implicated as a target of transcriptional activators (and repressors) by both genetic and biochemical methodologies. Consistent with a possible role for TBP in activator response, TBP has been demonstrated to directly associate with a number of transcriptional activators in vitro. This interaction was first demonstrated using VP16 (Stringer, K.F. et al., 1990). Mutations in VP16 that reduce transactivation in vivo also abolish TBP binding in vitro, consistent with a role for the TBP-VP16 interaction in vivo (Ingles, O]. et al., 1991). Physical interactions between TBP and activators have also been documented with Spl (Emili, A. et al., 1994), the Oct-1 and Oct-2 POU specificity domains (Zwilling, S. et al., 1994), c-Rel (Xu, X. et al., 1993), p53 (Chang, J. et al., 1995), the Epstein-Barr Virus transactivator Zta (Lieberman, PM. and Berk, A.J., 1991), and E2F (Emili, A; and Ingles, C.J., 1995), among many others. Most of these experiments were performed using recombinant TBP without TAF subunits. Since the TFIID complex comprises TBP and TAFs, the relevance of these interactions has been challenged. Given the large number of TAPS, it is reasonable to presume that TAFs might occlude any activator binding site. However, the interaction Of Zta and TBP has been demonstrated to occur in the presence of TAFS (Lieberman, PM. and AI. Berk, 1991). TBP-repressor interactions have also been described. Notably, the p53-TBP interaction observed in vitro has been implicated in p5 3-mediated repression of transcription (Seto, E. et al., 1992). The Observation that repressors are also capable of binding TBP suggests that simple recruitment to a promoter per se, is not sufficient to activate transcription. Interestingly, TBP residues implicated in VP16 binding map to the DNA-binding surface on the underside of the TBP “saddle” (Kim, T.K. et al., 1994). This suggests that VP16- binding and DNA binding may be competitive, an intuitively contradictory result. 15 However, TBP (Coleman, RA. et al., 1995) and TFIID (Taggart, AKP. and Pugh, BE, 1996) can dimerize through the DNA-binding surface. Thus, VP16 may enhance PIC formation by facilitating TBP dimer dissociation. Consistent with this possibility, dissociation Of dirneric TBP has been implicated as a slow step in TATA—recognition in vitro (Coleman, RA. and Pugh, BE, 1997). A number of genetic experiments in yeast have also implicated TBP as a potential target for transcriptional activators. Struhl and co-workers have identified TBP-TATA recognition as a limiting step in transcription in vivo that can be enhanced by activators (Klein, C., Struhl, K., 1994). This result is consistent with mammalian studies suggesting that TFHD recognition of a promoter may be limiting for transcription in vivo. Colgan and Manley foundthat overexpression of TBP in vivo stimulated transcription fi'om TATA containing reporter genes (Colgan, J. and Manley, IL, 1992). Interestingly, artificial tethering of TBP to a composite promoter containing LexA binding sites (by expression of a LexA-TBP fusion protein) activated transcription, suggesting that enhancement of TBP binding to a promoter is sufficient to activate a gene (Chatterjee, Struhl, K., 1995). It is unknown whether the LexA-TBP firnctions as native TBP (to nucleate formation of a PIC), or whether the fusion fortuitously produced an artifactual activation domain. Verification that gene fusions to DNA-binding domains can serendipitously produce artificial activators comes from early studies by the Ptashne laboratory. Synthetic activators were identified by screening a library of Gal4 fiisions produced from random E. coli DNA (Ruden, D.M. et al., 1991). It would be useful to know if TBP mutants that diminish DNA binding also diminish the activity of this LexA-TBP firsion protein. Klages and Strubin produced an artificial recruitment of TBP by introduction Of a leucine zipper l6 into TBP. Binding of a leucine zipper protein upstream Of a promoter activated transcription by the altered TBP, suggesting that TBP recruitment is sufficient to activate transcription in vivo (Klages, Strubin, M., 1995). Classical genetic experiments have also yielded a great deal Of information regarding TBP residues involved in essential protein-protein and protein-DNA interactions. Identified TBP mutant alleles can be generally grouped into four classes (reviewed in Hampsey, M., 1998): (1) DNA-binding specificity mutants, (2) TFHA interaction mutants, (3) TFIIB interaction mutants, and (4) activation-defective mutants with normal basal function. The DNA-binding specificity mutants of TBP were the first identified. SPT15 was originally identified as a gene essential for accurate promoter usage by the transposable Ty element of Saccharomyces cerevisiae. The 579115-122 allele contains a L205F mutation that alters TBP affinity for non-consensus TATA boxes, with transcription switching to a downstream HIS3 non-consensus CATAAA box (Amdt, K.M. et al., 1992). Introduction of an analogous mutation in the N-terminal TBP repeat results in preference for a TATAAG sequence, an indication that the polarity of TATA recognition was inherent to TBP and not conferred in combination with other components of the basal machinery (Amdt, KM. et al., 1994). Interestingly, a genetic selection constructed to identify activation-defective TBP alleles that did not interfere with basal function isolated mutations within the DNA-binding interface (Lee, M. and Struhl, K., 1995). All four mutants displayed alterations in DNA binding and normal TFHB and Gal4-VP16 interactions. Another screen designed to 17 identify Ind and Gal' TBP alleles also identified residues involved in DNA binding (Reddy, P. and Hahn, S., 1991). These phenotypes are commonly associated with defects in transcriptional activation (Amdt, K.M. et al., 1995). Collectively, these results argue that TBP-TATA recognition itself may be an active contributor to activator response in viva. Perhaps certain activators prefer certain core promoter elements, and alteration of TBP- TATA interactions influences activator-core promoter cooperation. The second class of TBP mutants involves disruption Of the TBP-TFHA interface. Perhaps one of the most intriguing TBP mutants is a member of this class. A genetic screen devised to identify temperature-sensitive (ts) TBP alleles with normal RNAPIII transcription identified a K138T/Y 139A double mutation. This double mutant does not respond to several activators, and fails to interact with TFHA, suggesting that TBP-TFHA association is critical for activated transcription in vivo (Stargell, LA. and Struhl, K., 1995). This result correlates with the requirement for TFHA for in vitro transactivation in mammalian systems (see TFHA section below). Interestingly, these residues are located on the H2 helix on the convex "top" of the protein, nowhere near residues involVed in TFIIA contact as indicated in the TBP-TFHA crystal structures (Geiger, J.H. et al., 1996; Tan, S. et al., 1996). However, the TFIIA derivatives used for crystallization lack a central protein segment likely to be disordered in the native protein. This region was excluded to facilitate crystallization. Perhaps a segment of this disordered domain contacts the H2 helix directly. Consistent with this possibility, fluorimetric and NMR Studies using synthetic peptides corresponding to this region of TBP and a TFHA segment from this central lOOp appear to associate in vitro (Ramboarina, S. et al., 1998). These 18 data agree quite well with the phenotypes of similar site-directed substitutions. Two double mutants of TBP (K133/K138 and K133/K145 mutants) are defective for TFIIA binding as evidenced by disruption of TBP-TFHA—DNA complexes in gel shift experiments (Buratowski, S. and Zhou, H., 1992). These studies are also consistent with in vitro Observations in mammalian transcription systems that implicate TFHA as a critical component for activator response (see TFIIA section below). Another interesting TBP allele appears to affect the TBP-TFHA interaction. Blair and Cullen identified a TBP mutant with enhanced RNAPII activity (Blair, W.S. and Cullen, BR, 1997). Gain-of-function mutants of the transcriptional apparatus are extremely rare. The phenotype arises from a single substitution, N69S. This mutation enhanced transcription from weak promoters without altering DNA binding affinity. Secondary mutations that diminished TBP-TFHA interaction blocked the N698 phenotype, consistent with the possibility that TFHA is involved in this process. The third class of TBP mutants involves disruption of the TBP-TFIIB interface. Certain yeast TBP mutations (L1 14K, L189K, or K271L) were found to diminish Gal4- VP16 activator response in vivo (Kim, T.K. et al., 1994). These mutants also displayed in vitro TBP-TFIIB interaction defects. Coupled with previous results suggesting that VP16 may directly bind TBP (Stringer, K.F. et al., 1990), the authors suggest that Gal4-VP16 may stimulate both TBP and TFIIB recruitment (Kim, T.K. et al., 1994). This proposal is consistent with the Observation by Green and coworkers that TFIIB recruitment to a promoter can be enhanced by activators (Lin, Y.S. and Green, M.R., 1991). A similar conclusion was reached by Tansey and Herr. They designed an E284R mutation to impair l9 the TBP-TFIIB interaction. This mutation blocked TBP-TFIIB interaction in vitro and severely impacted activator response in viva. An additional R169E mutation in TFIIB produced an allele-specific suppressor of the TBP mutant. The authors interpret this as evidence that activators function through recruitment of TFIIB (Tansey, WP. and Herr, W., 1997a). In contrast, Lee and Struhl identified several TBP mutants (E186A, E188A, and L189A) that are defective in TBP-TFIIB interaction in vitro, yet are competent to support activated transcription by several activators in viva (Lee, M. and Struhl, K., 1997). The reason for this disparity is unclear, although the nature of the substitutions may play a role. Perhaps the charge reversal introduced by Tansey and Herr produces a more severe phenotype which activators can overcome, whereas the alanine substitutions introduced by Lee and Struhl are less disruptive and therefore do not require activator for compensation. Consistent with this idea, Berk and colleagues have observed similar activator response defects in human TBP mutants with alterations of these corresponding human TBP residues (E284R, E286K or L287E) when charge reversal mutations of the type employed by Tansey and Herr were introduced. This particular study by Berk and colleagues (Bryant, G0. at al., 1996) is the most comprehensive study of TBP residues involved in transcriptional response that has been performed in viva. They produced 89 separate substitutions (charge reversals or introduction of large polar residues for hydrophobic residues) Of residues predicted by crystal structure to be surface exposed in human TBP. With the Objective of identifying TBP residues involved in activator response in viva, they discarded those that abolished 20 DNA-binding in vitro, as well as those that radically altered global protein structure. Activation by Gal4-VP16 and Gal4-E1A(CR3) was assessed with two different basal promoters in viva. The authors identified four clusters of TBP residues essential for activator function. Two patches, one on the upstream side of TBP (relative to the start site) and one on the stirrup of the second repeat of TBP, correspond to identified TFIIA and TFIIB binding sites, respectively. The other two sites, representative of the fourth class of TBP mutations, are less easily explained. One set contains residues in the H2 helix analogous to those described by Stargell and Struhl (Stargell, LA. and K. Struhl, 1995). The second cluster occurs on the downstream side of TBP. The authors postulate that these sites are likely to be involved in interactions with TAFs. However, in light of the observations of Stargell and Struhl, the H2 region could also be involved in secondary contacts with TFHA. TBP-associated factors may govern activator response The observation that activators directly interact with TBP suggests that activators may mediate TBP recruitment to promoters. However, TBP is actually associated with multiple accessory proteins (TAFs) in viva. Two lines of investigation suggest that these TAFS are involved in transcriptional regulation. The first and perhaps most direct evidence evolved from the original analysis and identification of TBP. Yeast TBP can be isolated from cell—free extracts as a single polypeptide. This allowed its purification and identification as the SPT15 gene product. However, metazoan TBP was more difficult to purify to homogeneity from cell-free extracts or nuclear extracts. 21 Additionally, recombinant human TBP or yeast TBP did not support activated transcription in vitro (Pugh, BF. and Tjian, R., 1990), while the impure TFIID fiaction (the 0.85M KCl eluate from a P-ll fractionated nuclear extract) robustly supported activated transcription (Horikoshi, M. et al., 1988a; Horikoshi, M. et al., 1988b; Sawadogo, M. and Roeder, R.G., 1985). This inconsistency was resolved by the identification of a collection of TBP-associated factors (TAFS) that restore activator response in vitro in conjunction with purified or recombinant TBP in Drosophila (Dynlacht, B.D. et al., 1991). These polypeptides were tightly associated with TBP and could only be dissociated from TBP by mildly denaturing conditions (urea plus salt washes). The identification of this high-molecular weight TFIID complex precipitated a series Of attempts to reconstitute activation in vitro using more defined systems composed of recombinant and purified, native TFIID. The Berk and Roeder laboratories have constructed stable HeLa cell lines that express TBP with an N-terminal epitope tag (using an influenza virus hemagluttinin (HA) epitope). This epitope is recognized by a monoclonal antibody (12CA5) that can be used to irnmunopurify the TFIID complex from nuclear extracts (Zhou, P. et al., 1992). The Tjian and Roeder laboratories have also used anti-TBP monoclonals to purify TFHD (Dynlacht, B.D. et al., 1991). Isolation of these complexes has allowed detailed analysis of the role(s) of TAF S in transcriptional activation as well as analysis of the role of TAFS in earlier steps of PIC formation. TAFS have also been demonstrated to be essential for activator stimulated formation of a TFIID-TFHA- DNA complex in vitro (Lieberman, PM. and A]. Berk, 1994). Multiple mechanisms through which TAFs might effect activation have been suggested. TAFs may function 22 through direct protein-protein interaction with activators, acting as adapters for activation domains. They may stabilize the PIC through additional supportive protein-DNA contacts, or may firnction as sensors of activator presence allowing isomerization Of the PIC to an activated state (reviewed in Roeder, R.G., 1996). Evidence consistent with all these possibilities exists. The earliest model for TAF firnction came from studies performed primarily in the Tjian laboratory. The metazoan activator Spl is active only in metazoan organisms. Presumably, yeast lacks a factor(s) essential for Spl function. The initial lack Of evidence for TAF polypeptides in yeast (yeast TAFS were eventually identified) suggested a potential role for TAFS in this species-specific coactivation of Spl. Consistent with this hypothesis, Hoey et a1. identified a 110 kDa Drosophila TAF that specifically binds Spl (Hoey, T. et al., 1993). Since that initial description, several TAF-activator interactions have been described. VP16 (as well as p53 and RelA) binds dTAFn4O (Goodrich, J .A. et al., 1993) and the human homologue, hTAFn32 (Klemrn, RB. et al., 1995). These TAFS dimerize in vitro with dTAFn6O and hTAFn80 (Weinzierl, R.O.J. et al., 1993). This dirnerization occurs through histone H3 and H4 related domains. The acidic activators p53 and RelA also bind both TAF S in vitro (Thut, C]. et al., 1995; Lu, H. and Levine, A.J., 1995; Burley, SK. and KO. Roeder, 1996). The Ile-rich activator NTF-l binds dTAFn150 and dTAFn150 (Chen, J. et al., 1994). Human TAFn3O associates with the estrogen receptor (J acq, X. et al., 1994), and the CREB activator has been found to bind dTAFn110 in vitro (Ferreri, K. et al., 1994). Cloning of the Drosophila TAF S has allowed the reconstitution of completely recombinant partial TFHD complexes. Importantly, reconstitution of these partial TFIID 23 complexes with minimal sets of TAFS has allowed identification Of TAFS necessary and sufiicient for function Of individual activators (Chen, J. et al., 1994). The dTAFu230 and dTAFn150 polypeptides appear to form the core of the complex, as both are required to recruit additional TAFS into these recombinant complexes. The addition of TAFnl 10 to this subcomplex allows activation by Spl, suggesting that Spl and TAFnl 10 interaction is suflicient for activator function in vitro. Subcomplexes assembled with dTAFn230 and dTAFn150, or with dTAFn230 and dTAFu60, support activation by Gal4-NTF-1 fusion proteins, but not by Spl, verifying that the dTAFnl 10-Sp1 interaction is also required for Spl activation in vitro. This result is also consistent with the observation of protein- protein interactions between the NTF-l activation domain and dTAFn150 and dTAFn60. Activation by Gal4-VP16C, Gal4-Jun, and Gal4-CTF fusions required the complete assembly Of the recombinant TFIID complex. These experiments with recombinant TAFS verify that activated transcription requires the presence of specific TAFS in vitro. Furthermore, these required TAFS correspond to those observed to participate in specific activator-TAF interactions. A second set of data suggests that TAFS may be directly involved in formation of the TFIID-promoter nucleoprotein complex, augmenting TBP-TATA DNA contacts. Original studies Of TFIID-promoter binding identified a DNAse protection profile considerably more extensive than that observed with recombinant TBP (Sawadogo, M. and R. G. Roeder, 1985; Horikoshi, M. et al., 1988b). TBP efficiently protected the TATA box, but the TFIID footprint extends downstream beyond the initiation site, while also inducing a number of hypersensitive sites throughout this region (Horikoshi, M. et al., 1988b). Interestingly, this extensive pattern of protection and hypersensitivity could be 24 stimulated by the activator ATF. This suggests that TFIID TAFS may directly contact DNA. Consistent with this hypothesis, dTAFn150 has been Observed to produce a similar pattern of DNAse protection over the initiation site (Verrijzer, C.P. et al., 1994). A second Observation also suggests that TAFS may be involved in control of the DNA binding activity of TFIID. Early Observations by Roeder suggested that hTAFn250 inhibited TBP recognition of the TATA box (Kokubo, T. et al., 1993 a). This inhibition is mediated by a short N-terminal domain of the TAF (Kokubo, T. et al., 1993b). A recent NMR study identifies how the N-temrinal domain mediates this inhibition. Interestingly, the domain appears to mimic TATA DNA, producing a protein surface complementary to the concave DNA-binding surface of TBP (Liu, DJ. et al., 1998). This domain could effectively act as a competitive inhibitor to DNA binding. Consistent with observed interactions between VP16 and the DNA-binding concave surface of TBP, VP16 and hTAFn250 competitively bind the DNA-binding site of TBP (Nishikawa, J. et al., 1997). This produces a simple model whereby activators might stimulate transcription by alleviating the effects of a TBP associated repressor that is actually a component Of TFIID itself. The third possible role of TAFS in transcriptional regulation is perhaps the most intriguing. The hypothesis that TAFS may be involved in isomerizations to and from active complexes was initially suggested by Roeder and colleagues. Since that time, similar ideas have been advanced by other investigators (Chi, T. and M. Carey, 1996). Carey and coworkers have described a DNAse protection pattern similar to that observed by Horikoshi, et al. (Horikoshi, M. et al., 1988a; Horikoshi, M. et al., 1988b). They have proposed that this DNAse sensitivity pattern reflects isomerization of a template 25 associated TFIID complex to a species with enhanced downstream interactions. Hypothetically, this could reflect initiator binding by the human homologue of dTAFnl 50. Formation of this DNAse protection pattern is enhanced by the Epstein- Barr virus activator Zta. This isomerization activity requires the presence of TFIIA, and formation Of the isomerized complex correlates with activation of transcription. Downstream promoter recognition by TAFS is consistent with recognition Of start site sequences by dTAFn150, although several other TAFS may have DNA binding activity. Cloning of the Drosophila TAFS dTAFn40 and dTAFn60, and their human homologues hTAFn32 and hTAFn70, revealed sequence similarity to histones H3 and H4 respectively (Kokubo, T. et al., 1994; Hisatake, K. et al., 1995; Hoffmann, A. et al., 1996). FurthermOre, cloning of hTAFn22 revealed sequence similarity to histone HZB (Choi, B. et al., 1996). Interestingly, crystal structures of these H3 and H4-related TAF proteins have been Obtained as a complex (Xie, X.L. et al., 1996). The crystal structure reveals that the two TAFS contain a classical histone-fold core domain, as well as conserved histone-like TAF-TAF interaction domains. This suggests that these two TAFS in conjunction with the H2B-related TAF may form a histone octamer-like structure. Consistent with this possibility, the TAF:TBP stoichiometry of these TAFS within TFIED has been Observed to be 4: 1, and a histone-like pattern of protein-protein interactions has been Observed (Hofi‘mann, A. et al., 1996). While a histone-like octameric TAF complex may form, the DNA-binding capacity of the complex remains an unsolved issue. Studies by Roeder and colleagues suggest that TFIID recognition of the AdML promoter results in wrapping of DNA (Oelgeschlager, T. et al., 1996). Conceivably, this wrapping could be accounted for through association of DNA into a TAF "nucleosome". However, 26 critical arginine residues in both histones H3 and H4 are not conserved in the TAF polypeptides. These arginine residues (H3 residue R83 and H4 residue R45) are both involved in critical histone-DNA contacts. If the TAF octamer binds DNA, it must do so in a fashion distinct from that employed by the histone octamer (see Burley, SK. and R. G. Roeder, 1996; Luger, K. et al., 1997). Another line of investigation suggests that TAFS could play a signal transduction role in transactivation. Work on Drosophila dTAFn230 by Tjian and colleagues has identified enzymatic activity. TFIIF is phosphorylated by dTAFn230 (Dikstein, R. et al., 1996). This activity requires two non-contiguous domains Of the protein, and the protein exhibits no sequence Similarity to any characterized kinase. Drosophila TAFu23O has also been demonstrated to acetylate histones (Mizzen, C.A. et al., 1996). Notably, the kinase activity requires the C-terminal bromodomain, which is not conserved in the homologous yeast TAF (Dikstein, R. et al., 1996). What role(s) these activities may play in activator function remains to be elucidated. Work on yeast TAFS has also provided some Of the most controversial and interesting results with regard to TAF firnction in viva. Despite early results suggesting that yeast lack a TAF-containing TFIID complex, yeast TAFS were eventually identified (Poon, D. et al., 1995; Reese, J.C., Apone, L., Walker, S. S., Griffin, L. A., Green, M. R., 1994). Deletion of several of the yeast TAF genes is lethal (Poon, D. et al., 1995). The yeast TAFS display a reduced affinity for TBP and purify as a separate complex. Reconstitution of this yeast complex produces a yeast TFIID complex similar to its metazoan counterpart. Identification and cloning of the yeast TAFS allowed analysis of TAF function in viva. Walker, et al. isolated a number oftemperature—sensitive alleles of 27 yTAFn145. After a shift to non-permissive temperature, cell growth ceased, but transcription of a large number of analyzed genes was unchanged (Walker, S.S. et al., 1996). Similar results were Obtained using galactose-regulated expression of various TAFS. TAFS were expressed from the GAL] promoter in the presence of galactose. A shift to glucose-containing media results in repression of this promoter and depletion of the expressed TAFS. Western blots confirmed reductions in TAF levels, although extremely low levels of TAF expression below Westem detection levels are difficult to rule out for in such experiments. Moqtaderi, et al. describe a similar set of in viva experiments using a copper- regulated system which selectively degrades specifically tagged TAF proteins (Moqtaderi, Z. et al., 1996). Strains containing ubiquitin fusion proteins of various TAFS were constructed. A copper-inducible promoter was used to induce overexpression of UBRl. UBRl encodes the N-end recognition protein of the ubiquitin degradation pathway. Thus, . TAF expression could be down regulated by copper induction through subsequent degradation of the targeted TAF. Several yeast TAF s were examined, with similar results. Depletion of TAFS did not significantly reduce transcription of several genes examined. Furthermore, TAFS appear non-essential for transcriptional activation in response tO several yeast activators. The authors suggest that TAFS may represent only one potential pathway of activation. While TAF-function might be redundant with multiple pathways in viva, these supporting pathways are non-functional in vitro. This would result in an apparent absolute requirement for TAF s for activation of transcription in vitro, with no apparent requirement in vivo. In contrast, Buratowski and coworkers performed similar experiments with isolated ts alleles Of the yeast histone—like TAFS (Michel, B. et al., 1998). 28 These TAFS are essential for transcription in viva. Shifting to non-permissive conditions disrupted production of a wide variety Of transcripts. The reason for this contradiction is unknown. Perhaps, only the histone-like TAFS are essential. The histone-like TAFS are also components of the SAGA histone acetyltransferase complex (Grant, RA. et al., 1998). They may be essential in this role. Results contradictory to the yeast studies have also been Obtained through genetic studies in Drosophila (Sauer, F. et al., 1996). Sauer et a1. conducted a genetic screen designed to identify EMS-induced dominant mutations that alter the expression of a Rasl reporter construct in the eye. Mutations that reduce or enhance this expression were mapped to the dTAFn110 and dTAFn60 genes. These mutations also reduced expression of several other eye-specific genes, but did not efl‘ect expression of all genes in the eye, suggesting that TAFS may be involved in recognition or regulation of a subset of genes in Drosophila. Furthermore, heterozygous flies expressing either a TAFnl 10 partially deleted allele, or one transdorninant TAFn60 allele displayed reduced expression of two bicoid responsive gene products. These results are consistent with previous Observations that these TAFS are required for synergistic function of the two bicoid activation domains (Sauer, F. et al., 1995). If TAFS are non-essential for activated transcription in yeast in viva, why is the phenotype Of TAF null alleles so dramatic? Recall that deletion of TAF genes is lethal. A potential cause for this phenotype was identified by Walker, et a1. and Apone et a1. (Apone, L.M. et al., 1996), although a hint was actually provided by data Obtained almost a decade earlier (Sekiguchi, T. et al., 1988; Sekiguchi, T. et al., 1991). Apone et a1. identified a temperature-sensitive allele of yTAFn90. Temperature-shift results in ngM 29 cell cycle arrest. Walker et a1. Observed a similar arrest with ts yTAFn145 alleles, although arrest occurred in G]. A ts allele of T SM] (the yeast TAFn150 homologue) also produced cells arrested in 62 (Walker, 8.8. et al., 1996). These phenotypes are similar to those described for the hamster cell-line tsl3. This cell-line was isolated as a cell-cycle ts mutant, incapable of cell cycle progression. The defect was complemented by a human clone designated CCGI (Sekiguchi, T. et al., 1988). Cloning Of dTAF1123O revealed that a close human homologue had previously been cloned as CCGI (Ruppert, S. et al., 1993). Cloning of the human TAF homologue (hTAFn250) verified its identity to CCGI (Hisatake, K. et al., 1993). Overall, these results suggest that TAFS are involved in regulation of only a subset Of genes, probably involved in cell-cycle regulation in yeast and metazoans, and perhaps in development and tissue-pattemed expression in metazoans. Although one cannot rule out the possibility that the cell cycle phenotypes of TAFS are not related to their transcriptional role in vitro, it is difficult to rationalize a mechanism whereby cell-cycle regulatory proteins could allow transactivation in vitro. The ability of partially reconstituted recombinant TFIID complexes to facilitate activator function in vitro is particularly difficult to rationalize, without proposing a true transcriptional firnction for these proteins (Chen, J. et al., 1994). One potentially interesting explanation for TAF activity involves the enzymatic acetyltransferase activity associated with the TFIID complex (dTAFn23 0, hTAFn250, and yTAFn145 are histone acetyltransferases). This is interesting in light of the converse observation that certain TAFS are also present in other histone acetyltransferase (HAT) complexes. TAFS have been identified in both the yeast (Grant, RA. et al., 1998) and 30 human SAGA HAT complexes (Martinez, E. et al., 1998b). The reason for their presence in these complexes is currently unknown. Perhaps the TAFS serve only a structural role. Alternatively, they may be tightly associated substrates. Notably, the current definition of a HAT is somewhat broad, and the specificity of these enzymes for histones should be regarded with caution. Other proteins may also be acetylation targets for these enzymes. The p53 tumor suppressor is specifically acetylated by the P/CAF HAT complex in vitro (Gu, W. and Roeder, R.G., 1997). These same p53 residues are also acetylated in viva (Sakaguchi, K. et al., 1998). Combined, these results suggest that TFIID plays a critical role in modulation of activated transcription. While experiments suggest that TBP recruitment may be limiting for transcription in viva, TBP mutants that are presumed to affect subsequent steps in PIC formation also display activation-defective phenotypes. These Observations reinforce that steps later in PIC formation may also be enhanced by activators. TFIIB recruitment may stimulate PIC formation The transcription factor TFIIB enters the PIC following TFIID recognition of the core promoter. The structure of the protein has been determined both in isolation (Bagby, S. et al., 1995) and in a TBP-TFIIB-DNA complex (Nikolov, D.B. et al., 1995). The factor binds directly to TBP and contacts DNA in the promoter complex immediately upstream and downstream of the TATA box. Consistent with this observation, Reinberg and coworkers have identified promoter sequences that contribute to AdML basal promoter strength, apparently through enhancement of TFIIB-DNA interactions (Lagrange, T. et al., 1998). TFIIB has also been implicated in start site selection in yeast 31 (Pinto, I. et al., 1992), and can be isolated from crude extracts in complex with RNAPII. These results have been interpreted as evidence that TFIIB may play a critical role in recruitment Of RNAPH to promoters (for review see Conaway, RC. and J .W. Conaway, 1993). A study by Lin and Green provided the first suggestion that TFIIB recruitment may be regulated by transcriptional activators (Lin, Y.S. and MR. Green, 1991). Formation of a PIC in vitro is a slow process, requiring 15-30 minutes for completion. The authors attempted to determine the "rate-limiting step" in PIC formation. The reasoning was that activators were likely to activate transcription by enhancing the rate of formation of the PIC, by analogy to enzymatic catalysis. The experiments were performed in a HeLa nuclear extract by first preincubating a promoter fragment in the extract in the presence or absence of activators. Following re-isolation of the promoter fragment, the collection of basal factors required for continued PIC formation was assessed by continuation of PIC assembly with defined combinations of partially purified basal factors. Slow recruitment of any essential factor (a "rate-limiting" step) would presumably be reflected in the lack of that factor in the partially assembled PIC, and a requirement for supplementation of that factor in the second incubation. If activators kinetically enhance assembly of the PIC, then the presence of an activator should alleviate this requirement. Pre-incubation of a Gal4-responsive Adenovirus E4T promoter template with nuclear extract resulted in formation of a partial complex that was still sensitive to activator presence in the second reaction. The second incubation did not require TFIID, but did require TFIIB, TFIIF/RNAPH, and TFIIE for activation. Omission of the activator from the first incubation did not change this requirement. This suggested that 32 TFIID recognition was not rate-limiting (TFIID was not required in the second incubation), and that activator was not required to recruit TFHD (even PICS assembled without activator did not require TFIID in the second reaction). The authors also concluded that PIC formation stalls at the point of incorporation of TFIIB into the complex. More importantly, the assembly of a partial TFIID-TFIIB complex required activator. Based upon these results, Lin and Green concluded that TFIIB incorporation into the PIC was a slow step that required activator participation in vitro, and that activators enhance transcription through recruitment of TFHB to an assembling PIC. This idea was strengthened by the Observation that Gal4-VP16 was capable of depleting an essential component(s) for transcription from nuclear extract. Addition of TFIIB to this depleted extract restored activity, consistent with their proposal. A similar study performed later by Choy and Green, employed Western blots to analyze partially assembled PIC complexes in the presence or absence of activator to confirm these earlier results (Choy, B. and Green, M.R., 1994). The authors verified recruitment of TFIIB by activator (Gal4-AH) to a promoter in vitro. In addition, the authors identified a second slow step after TFIIB association that could be enhanced in vitro by Gal4-AH, consistent with several proposals that activators facilitate multiple steps in PIC formation (Ptashne, M., 1988; Ptashne, M. and Gann, A., 1997; He, S. and Weintraub, S.J., 1998; Blair, W.S. et al., 1996; Brown, S.A. et al., 17; Chi, T.H. et al., 1995). Direct TFIIB-activator interactions have also been demonstrated using several transcriptional activators. As mentioned above, a VP16 protein affinity column depletes a HeLa nuclear extract of an essential factor(s). Western blotting of material eluted from this afiinity column confirms the presence of TFIIB in the bound fiaction. This interaction 33 appears to be direct, as recombinant TFIIB also binds VP16. Furthermore, VP16 mutants which reduce transcriptional activity in vitro and in viva (FP442 and truncation of the C subdomain) abolish this interaction (Lin, Y.S. et al., 1991). Experiments performed by O'Hare and coworkers conflict with these conclusions. Mutations within VP16C, which significantly impair VP16 activity, do not appear to affect VP16 binding to TFIIB in Similar experiments with HeLa nuclear extract (Walker, S. et al., 1993). However, additional factors present in the extract may stabilize TFIIB-VP16 contacts. Experiments assessing VP16 mutants performed by Lin et al. utilized recombinant or purified TFIIB. Significantly, TFIIB mutants defective for the observed TFIIB-VP16 interaction are also defective for activated transcription in vitro (Roberts, S.G.E. et al., 1993). However, Goodrich et al. reported that Drosophila TFIIB failed to interact specifically with VP16 in vitro. Mutations which reduce binding of mammalian TFIIB, do not effect the binding of Drosophila TFIIB (Goodrich, J .A. et al., 1993). This Observation is inconsistent with the robust activity of VP16 in the Drosophila system, suggesting that TFIIB interaction is not essential for VP16 firnction. Consistent with the observations by Lin et al., a number of other transcriptional activators also bind TFIIB in vitro. These include the Drosophila fushi-tarazu activator (Colgan, J. et al., 1993), the yeast activator Gal4 (Wu, Y. et al., 1996), the NF-xB family member RelA (Paal, K. et al., 1997), and several members of the steroid receptor superfamily (COUP-TF, Progesterone (PR) and Estrogen (ER) Receptors) (Ing, N.H. et al., 1992). In this last study, TFHB binding was found to be ligand independent for the PR and ER, suggesting that TFIIB interaction is not sufficient to explain transactivation by the liganded receptor. 34 Interestingly, a very similar experiment to that performed by Lin and Green was performed by John White in Pierre Chambon's laboratory, with a noticeably difi‘erent conclusion (White, J. et al., 1992). Utilizing a HeLa nuclear extract and Gal4-VP16 activator, the authors identified formation Of a partial PIC that was activation domain dependent. They employed an antibody that recognizes the VP16 activation domain. They reasoned that if activators accelerate formation of a critical intermediate in PIC formation, then following the formation of this critical intermediate, the activation domain should be dispensable for further PIC formation. Thus, masking the activation domain by an antibody at this point should have no effect on further PIC formation. Such an intermediate was identified. The kinetics of formation of this intermediate are consistent with those previorisly observed for formation of a template-committed complex, an event attributed to TFIID. Using partially purified basal transcription factor fractions, the authors were able to reconstitute a seemingly identical complex (as judged by antibody resistance) using TFIIA and TFIID. TFIIB had no effect on the formation Of this complex. Furthermore, TFIID and TFIIB could not form this antibody resistant complex. The authors conclude that VP16 acts through a TFIID-TFHA—DNA complex, independent of TFIIB. Another Observation is inconsistent with the conclusions of Lin and Green. Struhl and colleagues have analyzed the effect of a number of TBP mutants on transcription in yeast (Lee, M. and K. Struhl, 1997). These mutants were originally described by Ebright and Reinberg (Tang, H. et al., 1996). Lee and Struhl employed corresponding mutants of yeast TBP to study the importance of these in vitro interactions to transcriptional activation in yeast. Several yeast TBP mutants defective for TFIIB complex formation in 35 vitro, supported activated transcription by several activators in viva. The authors argue that this result makes it unlikely that TFIIB recruitment to a TBP-DNA complex is limiting for transcription in viva. The Basal Transcription Factor TFIIA is reguired for activator response The transcription factor TFHA has had a controversial and enigmatic history. Shortly after TFIIA was identified as a factor essential for specific initiation in vitro, experiments using partially purified systems suggested that the factor was not required for basal transcription and played only a minor stimulatory role in transcription. This suggestion seemed at odds with in viva studies in Saccharomyces cerevisiae. Yeast TFIIA comprises two polypeptides encoded by two genes, T 0A1 and T 0A2. Deletion of either T 0.41 or T 0A2 is lethal (Ranish, J.A. et al., 1992). This severe phenotype suggests that TFIIA plays a criticalrole in transcription in viva. Simultaneous efl‘orts by three laboratories shed light on the molecular basis for these inconsistencies. The initial work utilized either partially purified or recombinant basal transcription factors. Curiously, TFIIA stimulated transcription when partially purified TFIID was used, but not when TBP replaced these TFIID fractions (Ma, D. et al., 1993; Yokomori, K. et al., 1994; Sun, X. et al., 1994). The same result was also Observed with epitope-purified TFIID. These Observations imply that the stimulatory effect of TFIIA requires TAFS (Zhou, Q. et al., 1993). An additional complexity is suggested by recent work from the Smale and Roeder laboratories. Transcription of a core promoter lacking a TATA box, with a firnctional terminal deoxynucleotidyltransferase (TdT) Inr cannot be reconstituted in vitro with the 36 current ensemble of purified basal factors (TFHA, TFIIB, TFIID, TFIIE, TFIIF, and TFIIH). Smale and coworkers identified a column fraction activity which when supplemented with these factors reconstituted Inr promoter activity. They attributed this activity to the human homologue of dTAFn150, a TAF implicated by the Roeder and Tjian laboratories as being involved in downstream promoter recognition by TFIID (Kaufrnann, J. et al., 1996). Significantly, this TAF appears to be missing from hTFIID. Martinez, et al. have also succeeded in reconstituting a cell-free system capable of transcription from a TdT Inr template in vitro (Martinez, E. et al., 1998a). In contrast to the results of Smale and coworkers, Martinez, et al. found that hTAFn150 stabily associates with TFIID. The previous failure to identify the TAF in previous preparatiOn resulted from the fact that the TAF migrates aberrantly as a 135 kDa polypeptide in SDS PAGE. This results in comigration with hTAFn135, a previously identified homologue of dTAFnl 10. This hTFIID preparation, despite containing hTAF315O, did not reconstitute Inr basal activity. Further fractionation of the TFIID- containing crude fraction identified three non-TAF components (which they term TIC- 1/2/3) necessary to firlly reconstitute Inr activity. Thus, Inr function requires accessory factors, in addition to the set of basal factors sufficient for TATA containing promoters, although the precise identity of these factors is unknown. This observation suggests a potential explanation for the variable requirement for TFIIA function in previous studies. The AdML promoter is a composite TATA + Inr promoter. When the function of the AdML promoter is analyzed, only partial activity is observed with purified hTFIID, since the AdML promoter functions only through the TATA box in the absence of TIC and TFHA. Addition of these factors (by use Of cruder 37 fractions) allows the Inr element to function synergistically with the TATA element. Thus, studies employing TFHA-depleted nuclear extracts or crudely purified TFIID would display high levels of TFHA dependent activity due to TATA and Inr synergy. Further purification of TFIID eliminates TIC and results in increasingly reduced TFHA dependence. Yokomori et a1. provide an alternative explanation Of the variable requirement for TFHA (Yokomori, K. et al., 1993). The authors isolated TFIIA as a component Of Drosophila TFIID. Apparently, TFHA associates with TBP with relatively high affinity even in the absence of DNA. The TFIID-associated TFHA could be stripped from TBP by a moderate salt wash (3 00 mM KCl). Thus, variability in the ionic conditions used for TFIID preparation could significantly alter TFHA requirement. Regardless of the potential role of TFIIA in basal transcription, TFIIA is clearly required for activator response in vitro (Ma, D. et al., 1993; Yokomori, K. et al., 1994; Sun, X. et al., 1994; Zhou, O. et al., 1993; Ozer, J. et al., 1994). Reconstitution of activated transcription can be Observed with model activators (either Gal4-AH or Gal4- VP16) and purified basal factors only in the presence TAFS and TFHA. In reconstituted TFIID systems, TAF-dependent activation was determined to be completely dependent upon the addition of TFIIA to transcription reactions. Chi et al. argue that this efi‘ect of TFHA is the result of prevention Of the inhibitory effect of TAFS on TFIID DNA recognition (Chi, T.H. et al., 1995). Consistent with this possibility, Ozer et a1. identified TFIIA mutants incapable of activating transcription. These mutants were still capable of formation of a TBP-TFllA-DNA complex, but incapable of associating with TFHD and DNA (Ozer, J. et al., 1996). This effect could be reproduced with hTAFu250 alone. The authors argue that TFHA, in an activator dependent fashion, competitively displaces 38 hTAFn250, stimulating TFIID recognition of the TATA box on targeted promoters (Ozer, J. et al., 1998) Chiang and coworkers have described a TFHA-dependent, TAF-independent activator response to Gal4-VP16 in vitro (Wu, S. et al., 1998). Interestingly, this activator response does not require holoenzyme. The Kornberg and Young laboratories describe TAF-independent activation in vitro using preparations of holoenzyme (Koleske, A.J. and RA. Young, 1994). However, Chiang and coworkers (Wu, S. et al., 1998) succeeded in generating an activator-responsive in vitro system using extensively purified RNAPII (RNAP II was epitope purified and washed with 0.85M KCl and 1M urea), TFIIA and PC4 (a ssDNA binding protein implicated as a coactivator; see below). This suggests that TFHA may firnction in a TAF-independent capacity as well. A role for TFHA in response to activators has also been indicated by other results utilizing in vitro transcription systems. TFIIA has been demonstrated to enhance the affinity of TBP for DNA in certain EMSA systems (Maldonado, E. et al., 1990). Berk and coworkers (Lieberman, PM. and A.J. Berk, 1994) developed an agarose EMSA system for analysis of complex formation using epitope purified TFIID. In this system, TFIID-DNA binding is strongly dependent upon TFHA, and can be stimulated by activators such as the Epstein-Barr virus activator Zta or VP16 (Kobayashi, N. et al., 1995). The hypothesis is that TFIIA enhances the stability of this TFIID-TFIIA—DNA complex (DA complex) to electrophoretic conditions. The enhancement of this DA complex by Zta is activation domain dependent, and appears to depend upon the ability of Zta to recruit TFIIA. Zta mutants that do not bind TFIIA, do not enhance formation of the DA complex (Ozer, J. et al., 1996). The HSV activator VP16 can also enhance DA 39 complex formation. This activity has been localized to the VP16C subdomain, and VP16C has been shown to bind TFIIA (Kobayashi, N. et al., 1995; Kobayashi, N. et al., 1998). An interesting hypothesis related to this Observation has been proposed by Carey and coworkers (Chi, T. and M. Carey, 1996). They have proposed that TFIIA afl‘ects activator response by inducing an isomerization of the DA complex. This isomerized complex was suggested by the appearance of DNAse protection as well as hypersensitivity downstream of the TATA box extending through the initiation Site (Sawadogo, M. and RC. Roeder, 1985). This pattern was originally observed by Roeder and coworkers when footprinting with crude preparations of TFIID (see above). The appearance of these sites was attributed to DNA binding by TFIID and could be enhanced by the transactivator ATF (Horikoshi, M. et al., 1988b). Similarly, TFHA and Zta enhance the appearance of the downstream pattern of DNAse sensitivity seen by Chi et a1. It has been proposed, by Roeder originally and later by Carey, that this pattern of hypersensitivity reflects an isomerization of TFHD to a more extensive TFHD-DNA nucleoprotein complex. Accumulation of this isomerized complex correlated with transactivation, although an inability to directly trap this intermediate complex precludes any direct observation of its activity relative to unisomerized PIC. While only Zta has been reported to initiate this isomerization, the authors note that VP16 also induces DA isomerization, consistent with the observed activity of VP16 in DA complex formation assays (Kobayashi, N. et al., 1998). While the exact relationship between this isomerized DA complex and the electrophoretically stable DA complex originally identified by Lieberman and Berk is unknown, these parallels are interesting. Perhaps the stable DA complex is in fact an isomerized DA complex. 40 Activation may also involve non-basal auxiliafl factors While the majority Of the early investigations regarding activator function involved studies of activator interactions with basal transcription factors, a variety of genetic and biochemical studies have focussed recent attention on an expanding group of auxiliary cofactors. These coactivators have been identified in both yeast and mammalian systems. One of the earliest observations of coactivator complexes came from genetic studies in yeast. Berger et al. utilized a genetic selection based upon reversal Of the slow growth phenotype of yeast overexpressing a chimeric Gal4-VP16 activator (Berger, S.L. et al., 1992). Mutations {within the activation domain alleviated the phenotype, suggesting that it was transcriptionally linked. The selection has resulted in the identification of several yeast genes proposed to serve as transcriptional adapters: ADA 1, ADA2, ADA3, ADA4 (GCN5), and ADA5 (SPT 20) (Pina, B. et al., 1993; Marcus, G.A. et al., 1994; Marcus, G.A. et al., 1996). Mutations of any of the ADA genes alleviate the slow growth phenotype. The gene proteins appear to form a multi-protein complex, as protein-protein interactions between Ada2p and Ada3p have been reported (Candau, R. and Berger, S.L., 1996; Horiuchi, J. et al., 1995). They are required for efficient function of several activators (i.e. VP16, Gal4, Gcn4) and their deletion has no effect upon basal general transcription. This suggests that the proteins firnction as bridging factors, or adapters, between activators and the basal transcription machinery. Complexes of the Ada proteins and Spt proteins have also been identified in vitro (Grant, RA. et al., 1997). An additional hint as to the firnction of these proteins arose from a line of biochemical investigations in T etrahymena. This multinucleate organism has served as a 41 model for the regulation of gene expression through nuclear structure. The macronucleus is transcriptionally active, while the micronucleus is maintained in a transcriptionally silent state. Biochemical studies aimed at comparison of the structural packaging of micronuclear and macronuclear DNA implicated post-translational modification of histones in this suppression of gene expression. Macronuclear histones are extensively acetylated. Correlations between histone acetylation and transcriptionally active loci have been Observed in other species as well (Csordas, A., 1990; Wolffe, A., 1994; Turner, B.M., 1991; Loidl, P., 1994). This link between acetylation and the transcriptional competence of chromatin precipitated a search for the cellular machinery responsible for this post-translational modification. Biochemical studies of histone acetyltransferases (HATS) in T etrahymena identified both cytoplasmic (type B) and nuclear (type A) enzymes. Allis and coworkers purified the HAT A enzyme and cloned the HAT A gene. Comparison of the predicted gene product with the GenBank database identified similarity to the yeast GCN5 gene product (Brownell, J.E. et al., 1996). This observation provided the first concrete evidence that histone acetyltransferases function as transcriptional regulatory components. Genetic studies of Ty transcription in yeast have also produced additional evidence linking histone acetylation to gene regulation. The SPT genes were identified as suppressors of a TyzHIS3 integrant of Saccharamyces cerevisiae. The Spt mutants favor usage of the natural HIS3 promoter, producing stable HIS3 mRNA and protein. These SPT genes have been divided into two major subgroups: an TBP-associated group and a histone associated group. The ADA and SPT studies converged with the identification of ADA5 as SPT 20 (Marcus, G.A. et al., 1996; Roberts, SM. and Winston, F., 1996). 42 Biochemical analysis identified a high molecular weight complex of ADA and SPT gene products (Grant, RA. et al., 1997). This acetyltransferase complex (the SAGA complex) firnctions as a histone acetyltransferase in vitro, acetylating a mononucleosome containing a Gal4 binding site in a Gal4-VP16N dependent fashion (Utley, R.T. et al., 1998). Mammalian studies have also provided another link between histone acetylation and the firnction of an expanding number of transcriptional activators. The CBP/p3 00 family of coactivators was initially identified as a collection of related proteins capable of binding and enhancing activation by the mammalian CREB (Kwolg R.P.S. et al., 1994) and Adenovirus ElA activators (Arany, Z. et al., 1994). The family has since been implicated in the function of an increasing number Of activators (reviewed in Ogryzko, V.V. et al., 1996) including c-Jun (Arias, J. et al., 1994), c-Myb (Dai, P. et al., 1996), c- Fos (Bannister, A.J. et al., 1995), MyoD (Yuan, W. et al., 1996), and the nuclear hormone receptors (Kamei, Y. et al., 1996). In vitro studies identified CBP as a histone acetyltransferase (Ogryzko, V.V. et al., 1996). Roeder and colleagues also reported that CBP acetylates p53 (611, W. and KO. Roeder, 1997). This acetylation has also been observed in viva (Sakaguchi, K. et al., 1998). Studies with the CBP/p300 proteins identified an additional protein (P/CAF; p3 OO/QBP Associated F actor) which is thought to mediate p300 function. P300/CBP proteins can be found as part of a stable complex with P/CAF. Cloning and sequencing of the P/CAF gene identified extensive sequence homology to yeast GCN5 (Yang, X.-J. et al., 1996). A human GCN5 homologue has also been identified,-suggesting that metazoans may have evolved enhanced regulatory capacity through duplication and specialization of these enzymes (Inoue, M. et al., 1996). The proteins are homologous throughout their C-terrninal HAT domains, but diverge in their 43 N—terminal sequences. While these studies provide evidence for multiple HAT enzymes in higher metazoans, it should be emphasized that demonstration of HAT activity does not necessarily imply in viva relevance for the activity. Additional acetylation targets are likely to exist (for instance, the p53 target of CBP). Acetylation of TFIIB and TFIIF has also been reported (Imhof, A. et al., 1997). In the end, acetylation may prove to be a general nuclear signal transduction pathway. The family of coactivators and adapters is not limited to acetyltransferase components. Tjian and coworkers obtained evidence of a requirement for additional factors for activated transcription in vitro (Dynlacht, B.D. et al., 1991). This observation led to the identification of TAFS (see above) as non-basal factors required for activated transcription in vitro. Recent genetic experiments in yeast have confirmed that at least a subset of these TAFS is required for activator function in viva (Michel, B. et al., 1998). Biochemical studies in yeast have also identified a set of proteins believed to facilitate transcriptional activation. A complex of these proteins was isolated biochemically from Saccharamyces cerevisiae. This mediator complex is necessary and sufficient to allow activator response in vitro (Kelleher III, R]. et al., 1990). Cloning of the mediator genes identified overlap between these genes and components of the yeast holoenzyme. Another class of coactivators has been identified through biochemical studies in mammalian systems. Roeder and colleagues have dissected the transcriptional components biochemically through fractionation and reconstitution of transcription reactions in vitro. This process led to the identification of an Upstream Stimulatory Activity (USA) that enhances activator function. This USA, which is a relatively complex column fraction, contains several factors that differentially modulate transcription (in some 44 cases basal transcription, in other cases only activated transcription). Two positive control (PC) factors have been characterized, PC] and PC4. The best characterized is PC4. Ge and Roeder identified PC4 as a single-stranded DNA binding protein, which enhances transcription in cooperation with activators (Ge, H. and Roeder, R. G., 1994). The DNA binding activity is essential for PC4 function (Kretzschmar, M. et al., 1994). The USA fraction characterization has also led to identification of two negative regulatory factors that may contribute to activator function in vitro. The Dr2 protein was identified as Topoisomerase I (Merino, A. et al., 1993). The better-characterized repressor, Drl, functions as a heterodimer with DRAP]. The protein complex binds TBP preventing TATA binding and represses basal transcription (Inostroza, J .A. et al., 1992; Yeung, K.C., et a1, 1994). Both of these repressors probably augment the function of activators by repressing basal transcription in vitro. The basal factor TFIIA appears to be able to overcome this repression in vitro (Roeder, R.G., 1991). A third TBP-directed repressor, MOT 1 was identified through genetic studies in yeast. The MOT 1 gene product removes TBP from DNA in vitro, in an ATP-dependent reaction. Overexpression of TFHA suppresses a mat] mutation, suggesting that MOT] may function in opposition to TFHA in viva (Auble, D.T., et al., 1994). These repressors may cooperate to diminish transcription in the absence of a stimulatory signal. The ability of activators to derepress this inhibited basal state may be an essential component of activation in viva. Other complexes are also potential coactivator candidates 45 Transcription is also repressed in vitro and presumably in viva by chromatin structure (reviewed in Workman, IL. and Kingston, RE, 1998). DNA is packaged in viva into nucleosomes, each consisting of 146 bp of DNA associated with an octameric complex of four histones. The structure of this core histone particle associated with a 5S rDNA sequence has been solved (Luger, K., Maeder, A.W., Richmond, R.K., Sargent, DE, and Richmond, T.J., 1997). The crystal structure reveals extensive, periodic minor groove contacts between DNA and histones. This protein-DNA interaction obscures one face of the DNA helix as it is wrapped about the histone core. This nucleosomal packaging inhibits DNA binding proteins, significantly reducing affinity for their DNA binding sites. This inhibition of DNA binding likely extends to the PIC and its components. TBP binding is inhibited by nucleosome formation (Imbalzano, A.N. et al., 1994a), and pre-binding of TFIID partially alleviates the inhibitory efl‘ects of nucleosomes in vitro (Felsenfeld, G., 1992). Counteracting these inhibitory efi‘ects of nucleosome structure must be accounted for in any viable model for transcriptional regulation. Histone tail acetylation may provide one means to modify chromatin structure. The histone tails directly bind DNA and are likely to stabilize histone-DNA contacts significantly (Luger, K., et al., 1997), as well as influence higher order chromatin fiber structure. In vitro, under proper ionic strength conditions, arrays spontaneously fold into compact fibers (Fletcher, TM. and Hansen, J .C., 1995). This process is impaired in trypsinized nucleosomal arrays lacking histone tails (Logic, C.L. et al., 1998). Similar, but less dramatic effects are also seen when arrays are constructed with hyperacetylated histones. However, accessibility of DNA binding proteins to mononucleosomes can also be enhanced by acetylation, suggesting that higher 46 order structure alone is insufficient to firlly explain the repressive effects of histones (Lee, D.Y. et al., 1993). While the direct effects of acetylation may be sufficient to allow PIC formation on nucleosomal DNA, a class of nucleosomal remodeling factors may also enhance the accessibility of nucleosomal DNA sequences. These factors were initially identified through genetic studies in yeast. The Herskowitz laboratory described a set of S W] genes essential for mating type switching in yeast (Peterson, CL. and Herskowitz, I., 1992; Stem, M., et al., 1984). Simultaneously, the Winston laboratory characterized a set of yeast genes (SNF genes) essential for growth on sucrose (Winston, F. and Carlson, M., 1992). The products of these genes are found in a single high molecular weight complex in cell-free lysates (Peterson, C.L. et al., 1994; Cairns, B.R. et al., 1994). The SWI/SNF complex alters the DNAse footprint of mononucleosomes, enhancing DNAse cleavage of the histone bound face of the DNA helix in an ATP-dependent fashion. This activity is also manifest by increased accessibility Of nucleosomal DNA sequences to DNA-binding proteins (Cote, J. et al., 1994) and restriction enzymes (Logic, C. and Peterson, CL, 1997). An expanding number of SWI/SNF related ATP-dependent remodeling complexes have also been identified (Kingston, RB. et al., 1996). How these complexes are targeted to DNA loci remains unknown, however activators could conceivably play a direct or indirect role in this recruitment. Potential models for the function of activators How do activators function? The previous sections were intended to provide a survey of the components of the basal transcriptional machinery and regulatory 47 components, as well as evidence for their participation in transcriptional activation. Most models that have been advanced to explain activator function involve activator directed protein-protein interactions that stabilize or modify the basal apparatus. Activators have been proposed to stimulate the kinetics of PIC formation by recruitment of TBP, TFHA, TFIIB, or the holoenzyme itself Alternatively, activators may stabilize the PIC or induce its isomerization to an activated state once it has formed. This isomerization could be induced by conformational changes induced by complex formation between activators and their targets. The composite nature of eukaryotic promoters provides an additional complexity. Individual promoters may require stimulation of different steps of PIC formation. TATA-less promoters may be more dependent upOn activators that stimulate TFIID recruitment and evidence suggests that these promoters are TAF dependent. Activators are also likely to function prior to PIC assembly by facilitating access to DNA binding sites (either promoter elements or activator sites) obscured by the packaging Of DNA into nucleosomes. Recruitment Of histone acetyltransferase (HAT) complexes or chromatin remodeling complexes such as the yeast SWI/SNF complex may be involved in this process. Other derepression models have also been proposed. Activators may antagonize the repressive effects of repressors such as Dr] or Dr2. The discussion above focuses primarily upon activation through stimulation of PIC formation. However, activators are also likely to effect other steps downstream of PIC assembly. Activators such as Tat, SRF, and VP16 can function through stimulation of elongation (Yankulov, K. et al., 1994; Blau, J. et al., 1996; Spencer, CA. and Groudine, M., 1990). SRF has been proposed to recruit TFIIF to promoters (Zhu, H. et al., 1994). How Tat or VP16 stimulate elongation remains unknown. Other factors (TFHS, elongin, 48 etc.) also effect either the rate of RNAPII catalysis or the capacity of the enzyme to progress through pause sites, and loading of these factors could theoretically enhance transcription through RNAPII pause or arrest sites. These activators could also stimulate elongation through stimulation of CTD phosphorylation. Elongation may also be enhanced by recruitment Of TFIIH. The VP16 activation domain binds TFIIH in vitro (Xiao, H. et al., 1994). The TFIIH activity comprises a complex that contains both a helicase and CTD kinase activity (Drapkin, R. and Reinberg, D., 1994). Phosphorylation of the CTD has been proposed to facilitate elongation. This capacity of activators to stimulate elongation could introduce an additional level of promoter-specificity. Genes with promoter proximal pause sites such as the c-myc (Krumm, A. et al., 1992) or hsp70 (O'Brien, T. and Lis, J .T., 1991) promoters are likely more dependent upon activators that stimulate elongation. The mechanism of activator function has been an extensively investigated topic. The majority of the collected evidence suggests that activators work through multiple mechanisms, stimulating multiple steps of transcription upstream and downstream of PIC complex formation. Furthermore, different activators are distinct in their capacity to stimulate these different Steps. Not all activators are capable of stimulating integrated templates, suggesting that some may be incapable of antagonizing the repressive effects of chromatin (reviewed in Smith, CL. and G.L. Hager, 1997). Activators also differ in their capacity to interact with certain coactivator components. Finally, activators are difl‘erentially capable of stimulating elongation of the RNAPH complex (Yankulov, K. et al., 1994; Blau, J. et al., 1996). 49 We are interested in the primary structural elements that define activator function. We have utilized two model activators, VP16 and the mammalian activator RelA to study the structural requirements for activator function. Each activation domain is composed of multiple, independent subdomains, composed of short segments of hydrophobic amino acids. Each subdomain displays relatively similar structural requirements, with only subtle differences observed. Despite this similarity, evidence suggests that these subdomains do function through distinct mechanisms in vitro and in viva. 50 Chapter Two Activation Domains Contain Similar Yet Distinct Hydrophobic Motifs Introduction Herpes simplex virus (HSV) serves as a useful model for the study of the regulation of gene expression. Lytic infection involves coordinated expression of over 70 viral gene products, requiring intricate temporal control of three sets of viral genes. These genes have been classified according to their time of induction into three major classes: immediate-early, delayed-early, and late genes (Hayward, GS, 1993). Immediate-early genes, activated shortly after infection, encode viral regulatory proteins that control subsequent expression of delayed-early and late genes. Induction of the immediate-early genes is triggered in part by VP16. The protein is bifimctional, serving both this regulatory role, as well as a structural role as a tegument component of the virion. The gene is essential (Weinheimer, S.P. et al., 1992), although an HSV strain bearing a VP16 mutant lacking the activation domain (an amino acid 410-490 deletion) is infective in tissue culture systems (Rath Pichyangkura, unpublished results). The VP16 protein contains 490 amino acids comprising two domains. Deletion analysis of the C-terminus of VP16 has identified an activation domain within the C- terminal 90 amino acids (amino acids 413-490). This C-terminal domain is both necessary for VP16 activation (Triezenberg, S.J. et al., 1988) as well as sufficient to confer activator function to a heterologous DNA binding domain. Fusion of the C-terminal activation domain (amino acids 410-490) to the DNA binding domain of the yeast activator GAL4 (Gal4-VP16) produces a potent activator of transcription both in viva (Sadowski, I. et al., 51 1988) and in vitro in mammals (Carey, M. et al., 1990) and yeast (Chasman, D.I. et al., 1989). The N-terminal 410 amino acid domain is suflicient to recognize VP16 responsive cis-elements in the viral irnmediate-early enhancers. These elements are hybrid Oct-1 binding sites recognized by a trimeric complex consisting of Oct-1, a host cell factor (HCF), and VP16 (Preston, C.M. et al., 1988; Gerster, T. and Roeder, R.G., 1988). The N-terminal domain is also sufficient for tegument assembly (Rath Pichyangkura, unpublished results). The activation domain of VP16 contains an unusually high content of acidic amino acids. This observation has led to its classification as an "acidic" activator, a class for which it has served as prototype. Deletion of the C-terminal 40 amino acids results in approximately 50 per cent residual activity, while further deletion of an additional 40 residues results in full loss of activity in viva (Triezenberg, 5.]. et al., 1988), suggesting the existence of two functional motifs within the intact activation domain. These two "subdomains" will be referred to as VP16N (amino acids 410-456) and VP16C (amino acids 456-490). Both of these domains are capable of activating transcription independently in viva (Emami, K.H. and Carey, M., 1992; Seipel, K. et al., 1992). Initial mutational studies on the VP16 activation domain were performed on the VP16N subdomain. Cress and Triezenberg identified a critical hydrophobic residue, F442 that was essential for activation by VP16 (Cress, W.D. and SJ. Triezenberg, 1991). A rough correlation between acidity and the transcriptional activity of the domain was also suggested by the deleterious efl‘ect of asparagine substitutions for aspartic acid residues. However, this effect required simultaneous alteration of multiple residues; no single acidic residue was required for function. Apparently, only overall acidity of the domain is 52 important for function. Regier et al. demonstrated the importance of two additional hydrophobic residues within the VP16N subdomain (Regier, J .L. et al., 1993). These two residues, L439 and L444, flank the critical F442 residue and partially contribute to VP16N function. Neither appears as critical for firnction as F 442. Attempts to identify critical residues within the VP16C subdomain demonstrated weak efi‘ects of alteration of two phenylalanine residues (F473 and F475) within the subdomain, but interpretation of these results was complicated by the presence of an impaired VP16N subdomain in these constructs. We are interested in precisely defining the essential primary structural requirements for activation domain function. This topic has been extensively studied. Studies of VP16 (Cress, W.D. and SJ. Triezenberg, 1991; Regier, J .L. et al., 1993; Walker, S. et al., 1993), RelA (Blair, W.S. et al., 1994b; Schmitz, ML. et al., 1994), Spl (Gill, G. et al., 1994), p53 (Lin, J. et al., 1994), as well as the yeast activators Gcn4 (Drysdale, C.M. et al., 1995; Jackson, B.M. et al., 1996) and Gal4 (Leuther, K.K. et al., 1993), have each identified common structural features. These activators appear to require multiple hydrophobic residues for activation domain function. Furthermore, in many of these cases, activators appear to contain multiple activation motifs (subdomains) capable Of independent function (e. g. VP16, RelA, p53, GCN4). These studies imply that activation domains are constructed fiom repeated motifs, each functionally dependent upon their hydrophobic character. Work by Schaflher and colleagues on VP16N illustrates that these functional motifs may actually be surprisingly small. Seipel et al. found that a synthetic Gal4-fusion composed of a direct repeat of an 11 amino acid motif based upon the VP16N subdomain 53 sequence (surrounding the critical F442 residue) is sufiicient to direct activated transcription (Seipel, K. et al., 1994). This result implies that this 11 amino acid motif contains all sequence information necessary to confer function. Baeuerle and coworkers have also reconstituted activator function using a Gal4-fusion of a similarly compact segment of RelA (Schmitz, ML. et al., 1994). Collectively, these results suggest that activation domains are composite in nature, containing multiple compact hydrophobic motifs. Despite these similarities, examination of activation domains has failed to identify any consensus motif characteristic of these domains. The only apparent similarity revealed is a common requirement for the overall amino acid composition of these motifs. This has led to the proposal that the critical structural features for activation domain function are somewhat flexible. - Perhaps any DNA-binding domain associated, surface-exposed sequence rich in hydrophobic residues is sufficient for activator function (Ruden, D.M. et al., 1991; Sigler, P., 1988). This is poignantly demonstrated by the observation that insertion of an alanine residue into the center of the C—terminal activation domain of RelA does not significantly disrupt the strength of this activation domain (Schmitz, ML. et al., 1994). Alternatively, the apparent lack of conservation between different activation domains may be the result of comparison of unrelated and firnctionally distinct activation domains. A lack of structural conservation between functionally unrelated activation domains would not be surprising. To examine these structural questions, we utilized a strategy of mutational analysis to investigate the primary structural requirements for activation domain firnction, using two well—characterized activators, VP16 and RelA. Both are members of the acidic class 54 of transcriptional activators. Our results indicate that VP16 and RelA share a similar requirement for hydrophobic and aromatic amino acids for function. Furthermore, both activation domains are composite in nature. Despite these similarities, mutational analysis suggests subtle differences in the structural requirements of the VP16 and RelA subdomains. Together, these results suggest that activation domains are reliant upon similar, yet individually unique, core elements for function. Experimental Results The activation domains of VP16 are dependent upon similar hydrophobic motifs T o characterize the essential primary structural features of the VP16C subdomain, we employed a genetic strategy based on reversal of a slow-grth phenotype in yeast. Overexpression of a Gal4-VP16 fusion is relatively toxic to yeast (Berger, S.L. et al., 1992). This results in a slow growth phenotype, exhibited by reduced colony size of yeast expressing a Gal4-VP16C fusion from a high-copy yeast vector. The vector used in the experiments reported here (pPJHl3) is derived from a yeast 2p circle. This DNA is a recombinant derivative of an endogenous yeast plasmid that maintains a copy number of 25-50 DNA molecules per cell. The plasmid expresses a Gal4-VP16 firsion from the strong yeast ADH] promoter. A Similar selection strategy has previously been used to identify yeast mutants deficient for activator firnction in viva (Berger, S.L. et al., 1992). Berger et al. demonstrated that mutations within the VP16 activation domain that reduce transcriptional activity also reduce the severity of the toxicity phenotype. This linkage suggested that reversal of toxicity could be used to isolate potential VP16 transcriptional mutants. 55 Our isolation strategy is briefly outlined as follows: (1) introduction of random mutations within the VP16C subdomain, (2) subcloning of these mutants into pPJH13 to produce a library of Gal4-VP16C fusions, (3) transformation of this library into yeast, (4) selection of yeast bearing a transformed plasmid and displaying robust growth, (5) isolation of the plasmid from identified colonies and sequencing of the activation domain, (6) assay of the transcriptional activity of the VP16C mutant in yeast. To introduce mutations within the VP16C subdomain, Taq DNA polymerase was used to amplify VP16C under conditions that reduce polymerase fidelity. The VP16C PCR products were subcloned into pPJH13, producing a library of Gal4-PCR product fusions. This pool of pPJH13 variants was transformed into yeast strain BP] (It/L4 Ta ade1-100 ura3-52 leu2-3,112 his4-519) and cells were plated onto Leu' media for selection of pPJH13 transforrnants. High level expression of this wild-type VP16C firsion results in reduced colony size relative to empty vector controls ("wild-type VP16C" will be used to refer to Gal4-fusions containing VP16 sequences which correspond to the native VP16 gene sequence). Colonies were selected for secondary screening based upon visual inspection of colony size. Large colonies were selected and patched onto selective media to verify this phenotype. DNA was isolated from candidate colonies and transformed into E. coli for amplification. The plasmids were then isolated from E. coli and sequenced to identify any activation domain mutations. To assess the transcriptional phenotype of each mutant Obtained, candidate VP16C activation domain mutants were subcloned into low copy yeast expression vectors (ARS/CEN vectors). These vectors also express Gal4-VP16C firsions under the control of the ADH] promoter, but are maintained at copy numbers of 56' 1-2 per cell. This allows lower levels of expression and reduces the toxicity of the fusion proteins, allowing more accurate quantitation of their transcriptional activity. Activity was quantitated by cotransformation of these Gal4-VP16C fusion plasmids with the plasmid pLGSDS (Guarente, L. et al., 1982). This reporter plasmid contains a bacterial lacZ gene under the control of the GAL] -] 0 UAS and a CYC] TATA box. The GAL] -] 0 Upstream Activating Sequence (U AS) corresponds to the natural yeast sequences regulating the GAL] and GALIO genes. The UAS contains four Gal4-binding sites. Three individual transformants of each mutant were assayed by the method of Miller et a1. (Miller, J H 1972) The collection of mutations identified as well as their relative transcriptional activities appears in Figure 1. Of 14 mutants isolated, 11 have mutations occurring between amino acids 470 and 480, highlighting the probable importance of this region for activation domain firnction. Nine Of 13 missense mutations have lesions at one or more phenylalanines within the activation domain. Eight of these phenylalanine mutations identified occur within the core 470-480 region (F473, F475 or F479). Three of four mutants classified as exhibiting a severe phenotype (less than 25% wild-type activity) contained substitutions of one or more hydrophobic residues within this core region by polar amino acids. Significantly, conservative mutations of these hydrophobic residues (e. g. F479L, F457L/F479L, F47 5L) exhibited little deleterious effect, suggesting that retention of hydrophobic character at these positions is sufficient to confer activation domain function. The remaining mutant exhibiting a severe phenotype consisted of a prematurely terminated open reading frame due to mutation of the glutamine 477 codon to a stop codon, reinforcing that the entire core region is essential for function. 57 $an 5 woueomvfi 8e 89596:; 3238.595 9.8 58m 3 3322: 8a Soho 3:23:85: 0: no 23: £3, 3:832 en: E 938me 03 53:3 Raozmtomgb E 82838 838 warms—“Ev 2384 Locum 28: Emu 05 E 33% 3—38 2E. 25502 E concave me 5338 3:23:85: Com cabana 803 3883 couezoe 38:8 Baa—onsm .260 Box 05:8 Bro—-m—wfim E 8:865 225533 Bow 058“ 2: £3» 65883 begun £5 323 328me 8a 203838 eoaucoa .38 Bow 05:3 cote— -2wEm 5 a9 2.: $82.... Deanna EeEoBa 03% 05 .«o 8325a SEE 2F 05% oczwm 05 E can: 08 @3352 30:8an 05. 303838 Enact 858,38 33:28 @962 9 30538 203 mEeEoe c2338 2:. 30508 E wantomow me @8023. 803 moabocona 9.22033 2: 8m oratoomoc $5338 USE/-225 £2338 ocosw .3 35:52 28:85: US?» ._ ouawwm AXVN LOPWN‘FVO DDDDDDDDDDDDDNDDDDDDDDDDDDDDDDDDDDDDDDDDD D D D D D D D D D D D U) D D D D D D D :02 85> -.»».>»»...»».»p.»»»».».»O»»».»»»....»».» gm? wcvmd DDDDDDDDDDDDDDDDDDDDDDQDDDDDDDDDDDDDDDDDD :2 P38 trrrtpptprpppmppp.-p»..->.»>>»p»»».»»rrpt :2 $422.8va »»»»»>»»»».»>»»»p..pprmprpp».».»».»»»»».r as: mus: --prprppprrprrtqtppprp»»»»>>»»»»»»»»».-»» $8 fiEaéSOm »»»»»».».»».»»»AO»»..rrpppprrprrprrtrtptp :8 Eng ..ppprrrtppppppptmtrr.>ttpttprtrtrr».i-» $2 $48.?va ..p»...».r.»»p>»Om..r»..-»>»»>»..p»».p»r» §v asvmmfifizfivma »»».»».ptprmhptmrpr...t.»»p»»»».».p»»p>»» $2 £48.8va tiprpppttopmrtppptrtrr.-,.--»ptprpptrptpp 32: 84.5.5315 pppprpp..trq».»»....»».»»..»>.r»»a»»»>»»» $8 stud .»»»»»».»»»4»»»».r.»».»»>>»»»>>>»»»»pr»pp E353 mzoflfibs oceans“Oam_0:-_0:-t0: 25> 2:030: =5 25— 885.: <30: 50m 4.2098: :5 <>_0:- 85088 ESE/450 .8 010 :88 $5.55 0228 oz .30: 2 02 so: 01 m e5 .N ._ .30 ._ .8 0202430 5 .20.: e :2 so: .0: 8m :5 .02 .2: .8 .o: 52.5-28 .20: 9 02 so: 0: om Pa .2 .m .N .0 02.3-38 .20: 8 :2 so: .01 m Pa .N ._ .8: .2 .8 2292-350 .8535 0 Z ”Emu 0:. 0:0— 80: 85.: 508555: E 552:5 2 8:00.55: :50 :_ 285—: 8:300:20 5:00.55: .8 >850: 05. 228:0: 03-8: :5 2T32 :5 08:: 0:609: 0: 5 m5? 8:80:00 5:8: .3 :0880: 0:03 32:85.: 0m_00:m .:0:00:m5b-:8: 200: m: 2.00 :080mm5b 80:: :0505: 53 <72 2:5 .320: 352:5 5v 885—: 008:0: A05 05::— 05253: 08:88-8: 8:5 .U 88> 88> . m :duzo :dozo 8 <8: .4 5288 97 17 ng yTBP 33 ng yTBP TFHA TFHA free probe on » mun , _ A '3‘ , k, 3 s. 2‘ ‘1- Figure 9. TFIIA stimulates TBP recognition of the AdML promoter TATA box. To test the activity of recombinant TFIIA, TFIIA was incubated with two quantities (17 ng, lanes 2-7; 33 ng, lanes 8-13) of yeast TBP (yTBP). Under these electrophoretic conditions (see Methods) yTBP does not associate with the AdML TATA box in the absence of TFIIA (lanes 2 and 8). Addition of TFIIA results in stable binding to the TATA box in a TFIIA dose dependent fashion (see lanes 3-7 and 9-13). TFIIA alone does not bind DNA (lane 14). 98 The polyoma enhancer produces not only the technical problem of upstream start sites, but also a possible scientific complexity. The two VP16 subdomains may differ not in promoter recognition, but rather in their ability to synergize with the polyoma enhancer. To test the contribution of polyoma sequences, we removed all of the polyoma sequences from the reporter constructs. Reporters bearing or lacking polyoma sequences were cotransfected into mouse L cells. The polyoma bearing reporters were included as internal controls for transfection and primer extension. In each transfection, a Gal4-VP16FL expression vector served as activator (VP16FL does not display basal promoter sensitivity; see above). The results appear in Figure 8C. In all cases, primer extension signals were evident only from the polyoma bearing reporter plasmids. No primer extension products were detected from cotransfected plasmids lacking the polyoma enhancer. These results confirm that polyoma sequences are essential to observe detectable signal from these synthetic reporters. Either these results reflect a direct effect of basal promoter structure on activator function, or they reflect a basal promoter effect on activator-enhancer synergy. Regardless, the primer extension results suggest differences in the mechanism of function of the two subdomains The VP16C subdomain recruits TFIID and TFHA to a promoter in vitro A number of observations suggest a role for TFHA in response to activators (reviewed in introduction). Berk and coworkers have developed an in vitro TFIID electrophoretic mobility shift assay (EMSA) system. The size of the TFHD complex precludes the use of traditional polyacrylamide EMSA systems, since TFIID-DNA complexes do not migrate from the loading wells in these systems. To circumvent this 99 difficulty, Lieberman et al. performed EMSA in a 1.4% agarose gel in the presence of Mg2+ to stabilize TFIID-DNA complexes (Lieberman, RM. and A.J. Berk, 1994). In this EMSA system, TFHA dramatically enhances the apparent affinity of TFIID for DNA. The TFHA polypeptides appear to enhance the stability of TFIID on DNA by forming a TFIID-TFIlA-DNA complex (DA complex). Results from the Berk laboratory suggest that certain activators can also stimulate formation of this DA complex. The Epstein-Barr Virus activator Zta was the first activator demonstrated to enhance DA complex formation (Lieberman, P., 1994). Gal4-VP16 fusions can also stimulate DA complex formation (Kobayashi, N. et al., 1995; Kobayashi, N. et al., 1998). This activity appears to localize to the VP16C subdomain, since the VP16C subdomain strongly stimulates DA complex formation while the VP16N subdomain has no effect (see below and Kobayashi, N. et al., 1998). We wished to examine the capacity of the VP16 and RelA subdomains to stimulate DA complex formation. We first needed to establish appropriate conditions for the EMSA experiments. Human TFHA, human epitope-tagged TFIID (eTFIID), and Gal4- activation domain fiJsions were prepared as outlined in Methods. The TFHA was assayed for the ability to enhance the affinity of recombinant yeast TBP for the AdML promoter in a polyacrylamide EMSA system (Figure 9). Radiolabeled AdML promoter DNA was incubated with yeast TBP (yTBP) in the absence (lane 2, 17 ng yTBP and lane 8, 33 ng of yTBP) or presence of an increasing concentration of recombinant human TFIIA (lanes 3-7 and lanes 9-13). Under these electrophoretic conditions, DNA-binding by yTBP is completely dependent upon TFHA (Maldonado, E. et al., 1990). The TFHA preparation efficiently stimulated yTBP recognition of the promoter. 100 Gal4-VP16N and Gal4-VP16C firsions were prepared as described in Methods, and tested for DNA-binding activity. The EMSA probe used for this and all subsequent EMSA experiments is a 225 bp restriction fragment of pG5E4TCAT containing five Gal4 sites and the adenovirus E4T promoter upstream of a segment of the bacterial CAT gene. Performance of the agarose EMSA experiments requires prior determination of the quantity of each Gal4-fusion protein necessary to saturate all five Gal4 binding sites. This was determined using a conventional polyacrylamide EMSA system. An example of such an experiment appears in Figure 10. Conditions saturating for all five sites in this experiment were used to determine appropriate Gal4-fusion quantities for the agarose EMSA experiments. To test the capacity of each VP16 subdomain to stimulate DA complex formation, we performed an agarose EMSA (Figure 11). Radiolabeled DNA was incubated with immunopurified TFIID and recombinant TFIIA in the absence or presence of Gal4-VP16N or Gal4-VP16C (one, two, or four times the amount of protein necessary to saturate the probe). DA complexes were resolved by electrophoresis in a Mg” containing electrophoresis buffer on a 1.4% (w/v) agarose gel followed by autoradiography. A small degree of DA complex formation was apparent even in the absence of activator (lane 3). In the presence of Gal4-VP16C, the extent of this DA complex formation was significantly enhanced (lanes 5-7). However in the presence of Gal4-VP16N, DA complex formation was actually inhibited (lanes 11-13). In each case, formation of these complexes was completely TFIIA dependent (see lanes 4, 9, and 15). These results agree with those previously observed by Kobayashi, et al. (Kobayashi, N. et al., 1995). 101 Gal4-VP16N Gal4-VP16C free probe Figure 10. Saturation of the EMSA probe with Gal4-AD fusion proteins. Gal4- VP16N or Gal4-VP16C fusion proteins were incubated with radiolabeled adenovirus E4T DNA under conditions identical to those employed in the agarose EMSA experiments, followed by electrophoresis on 4% PAGE in leBE. Lane 1: free probe. Lanes 2-9; increasing concentration of the Gal4-VP16N fusion protein. Lanes 10-18; increasing concentration of the Gal4-VP16C fusion protein. The probe contains five Gal4 binding sites. Species with 1,2,3,4 or 5 occupied sites are each resolved. VP16C VP16N VP16NAD/L A AAAA AAAA AAAA D DDD D DDD D DDD D Figure 11. Comparison of the capacity of VP16N, VP16C and a mutant VP16N subdomain to enhance DA complex formation. Recombinant Gal4-VP16C (lanes 4-9), Gal4-VP16N (lanes 10-15), or Gal4-VP16NAD/L (lanes 11-16) were incubated with a radiolabeled adenovirus E4T promoter probe, in the presence or absence of TFIIA (A; 20 ng) and/or TFIID (D; 1 uL). Following incubation, complexes were separated on a 1.4% agarose gel in a Mg—containing TBE buffer. The Gal4-fusions were present at a concentration as indicated by the filled triangles above each set of lanes (0.5 times, 1 times, or 2 times the concentration required to saturate the Gal4 sites under identical incubation conditions). TFIIA and TFIID without activator (lane 3) weakly associated with the probe. Addition of VP16C (lanes 5-7), but not VP16N (1 1-13) or VP16NAD/L fusions (17-19), enhanced DA complex formation. In all cases, efiicient DA complex formation required both TFIIA (lanes 2, 9, 15 and 21 lack TFIIA) and TFIID (lanes 8, 14 and 20 lack TFIID). l03 In collaboration with Arnold Berk and Naoko Kobayashi, we assessed a series of VP16C mutants for the ability to stimulate DA complex formation. Using a collection of our VP16C mutants, Naoko Kobayashi performed DA complex agarose EMSA assays. The VP16C mutants were chosen to display a range of transcriptional strength from nearly wild type to completely defective as assayed in yeast and mammalian cells in vivo. The results appear in Figure 12. The VP16C mutants are arranged in order left to right, from strongest to weakest (as determined by the in vivo transcription assays). The in vivo transcriptional activities of the mutants correspond well with their ability to stimulate DA complex formation in vitro. We further tested the ability of wild type and mutant VP16C to bind directly to TFIIA. Gal4-VP16C proteins were in vitro translated, incubated with myc-tagged TFIIA and coimmunoprecipitated using an anti-myc epitope MAb. The results demonstrated a direct correlation between the ability of the VP16C mutants to bind TFHA, and their transcriptional activity (see Figure 12). In contrast, coimmunoprecipitation experiments performed with TAFn3 2, TBP, TFIIB and the p65 subunit of TFIIH did not display this correlation. These results suggest that VP16C stimulates DA complex formation through recruitment of TFHA to an assembling PIC. Furthermore, the VP16N activation domain does not stimulate DA complex formation, demonstrating that VP16N and VP16C can fiinction through distinct mechanisms in vitro. Expression of VP16N or VP16C activation domains inhibits transcription in vivo To determine whether or not the VP16 subdomains function through. similar mechanisms in vivo, we also employed a strategy originally used by Chambon 104 Figure 12. Gal4-VP16C stimulates formation of a TFIID-promoter complex through recruitment of TFHA to DNA. Panel A. An agarose EMSA experiment was performed using Gal4-fusions of wild type VP16C and several VP16C mutants. Gal4-VP16C fusions were incubated for 20 minutes at 30°C with radiolabeled adenovirus E4T promoter probe in the presence of TFIIA and/or TFIID (as indicated above each lane). Following the incubation, complexes were separated on a 1.4% agarose gel using a Mg-containing TBE buffer. The mutants are arranged in descending order of in vivo transcriptional potency, fi'om left to right. The positions of free probe and various complexes are indicated at the right of the figure. Panel-B. Phosphor Imager quantitation of the image of panel A. Mutants: C1, D472A; C2, Q477A; C3, F479A; C5, F473 S/E474G; C6, F47 5S/M478T/F479S; C7, F473A/F475A; C8, F479L; C9, F479S/D486G. Panel C. Interaction of TFIIA with Gal4-VP16C mutants. The mutants are arranged in descending order of in viva transcriptional potency, from left to right. Gal4-VP16C (or the indicated mutant) was incubated in the presence (+, lanes 3, 6, 9, 12, 15, 18, 21, 24, and 27) or absence (-, lanes 2, 5, 8, 11, 14, 17, 20, 23, and 26) of Myc-epitope tagged TFIIA. Following incubation, monoclonal 9E10 (a Myc-epitope monoclonal) was used to immunoprecipitate TFIIA. Co-irnmunoprecipitated material was subjected to SDS-PAGE. Bound 3SS-Met labeled Gal4-VP16C was quantified by Phosphorimager, using 5% of input Gal4-VP16C (lanes 1) as a reference. Panel D. Numerical quantitation of the autoradiogram in Panel C. Means and standard deviations of three experiments are shown. (Figures obtained from Kobayashi, et al., 1998) 105 eTFIlD_+++++++++++++—+— A. rTFllA--+++++++++++-—+— G4VP160 — — —wrcs C2C1CSC3 C9c7csw-wrwrwr- Q (—|ID+IIA+G4VP16C . (lIDHIA ‘ * <—G4VP160 ”LufFreeprobe 12 3 4 5 6 7 8 91011121314151617 B 100- 80‘ 60‘ %Wild type N A o 9 WT CB C2 C1 CS C3 CS C7 CS C 5‘” “”50 WT CB C2 C1 C5 C3 C9 C7 CG . I — +Il — +I“-\._. +ll| >— +lll — +ll| — +Ill — +lll — +lll — +1 “‘ ~ and... “Masai “Can... a . ..,... w p. 12 3 4 5 6 7 8 9101112131415161718192021222324252627 D. 120 1 00 80 60 96 Wild type 40 20 F‘ 12 ° lgure ' WT C8 C2 C1 C5 C3 C9 C7 C6 106 and coworkers (Tasset, D. et al., 1990) to analyze the glucocorticoid receptor. An in vivo competition strategy was employed to assess the mechanistic similarity of activation domains from the glucocorticoid receptor (GR) and estrogen receptor (ER). Activation domains lacking DNA binding domain sequences were overexpressed as chimeric proteins containing N-terminal fusions of the SV40 T antigen nuclear localization sequence (NLS) to activation domain sequences. They reasoned that if these constructs are still capable of recognition of their nuclear target yet lack DNA-binding domains, then their expression should result in the progressive titration of the target from nuclear pools. This titration should prevent activation by any activator that utilizes the same target protein. Using this strategy, Tasset, et al. argued that the N-terminal activation domains of the GR and ER function through the same mechanism. The strategy has been employed in several other cases (Schmitz, M. et al., 1995; Carruth, L.M., et al., 1994; De Felice, M. et al., 1995; Folkers, G.E. et al., 1995). Using plasmids obtained from Chambon, we produced mammalian expression vectors for NLS-VP16N (pNLS-VP16N) and NLS-VP16C (pNLS-VP16C) chimeric proteins (“competitors”) to test their ability to interfere with the activity of either Gal4-VP16N or Gal4-VP16C (“efi‘ectors”). An example of the results of an experiment using Gal4-VP16C is shown in Figure 13A. Competitor and effector DNAs were cotransfected into mouse L cells by the DEAE-dextran method. Cotransfections were performed with a Gal4-VP16C effector (100 ng), varied competitor DNA quantities (as indicated in the figure legend) and 1 ug of the pGST+I/CAT reporter construct (described previously; see Chapter 2). In each transfection, DNA levels were maintained constant by the use of pSGS as carrier DNA. The pSGS plasmid is the parental DNA 107 Figure 13. In viva competition experiments with NLS-VP16 and Gal4-VP16 expression vectors. Panel A. Mouse L cells were cotransfected with 100 ng of a Gal4- VP16C expression vector, 1 ug of a reporter construct which expresses chloramphenicol acetyltransferase (CAT) from a composite TATA + Inr promoter downstream of five Gal4 binding sites, and an increasing dose (0 to 10 ug) of one of two constructs expressing either the VP16N (set of bars on the left) or VP16C activation domains (bars on right) fused to a nuclear localization sequence. The total DNA quantity transfected was held constant by inclusion of the parental NLS-fusion expression vector pSGS as a carrier DNA. Bars represent the average of three separate transfections plus or minus the standard deviation. CAT activity was determined by a floor diffusion method, using 3H- acetyl-CoA and unlabelled chloramphenicol. The activity is expressed as arbitrary units normalized to the control transfection (0 ug of competitor and 10 ug of pSGS). Black bars: 0 ug of competitor and 10 ug of pSGS. White bars: 1 ug of competitor and 9 pg of pSGS. Grey bars: 5 ug of competitor and 5 ug of pSGS. Right cross-hatched bars: 10 ug of competitor and 0 ug of pSGS. Panel B. same as Panel A, except that the experiment was performed with 100 ng of a Gal4-VP16N expression vector. 108 NLS-VP16N NLS-VP16C NLS-Competitor nmol acetylcoA conv./h/mg NLS-VP16N NLS-VP16C NLS—Competitor Figure 13. 109 of the competitor plasmids, and contains an SV40 enhancer upstream of an empty multiple cloning site. This polylinker is the site of insertion of activation domain sequences in the competitor constructs. While 1 pg of pNLS-VP16C was sufficient to inhibit Gal4-VP16C activity 65%, the same quantity of pNLS-VP16N had little observable effect (<6% inhibition). Thus specificity, as evidenced by the greater effect of the pNLS-VP16C competitor on Gal4-VP16C, was observed at low levels of competitor DNA. However, at higher quantities of DNA, both competitors inhibited Gal4-VP16C activity (compare 10 ug bars, Figure 13A). Results from a similar experiment performed with a Gal4-VP16N effector DNA displayed different results. Inhibition was not observed at competitor levels below 5 ug, and at 5 ug both competitors inhibited VP16N activity slightly (see Figure 13B). If competitOr levels were increased to 10 ug, pNLS-VP16N displayed greater inhibition than pNLS-VP16C. These results are consistent with partially independent mechanisms of activation by VP16N and VP16C. The specific effect of the NLS-VP16C competitor on Gal4-VP16C activity at low competitor dose (Figure 13A) suggests that the VP16C competitor titrates a limiting factor essential for VP16C firnction. Furthermore, the absence of an effect of the VP16N competitor does not appear to result from a lack of expression of the NLS- VP16N fiision, since the VP16N competitor is capable of inhibition of Gal4-VP16C at higher concentrations (Figure 13B). The inhibition of Gal4-VP16C observed by both competitors at high competitor DNA levels, could be explained by either of two models. The activation domains might contact multiple targets and share at least one target. This hypothetical common target would have low affinity for VP16N and high affinity for VP16C. Alternatively, the inhibition observed at higher levels of competitor could be non- 110 specific due to the presence of the strong SV40 enhancer/promoter on the competitor plasmids resulting in lowered expression of the chimeric activator and reduced transactivation of the reporter construct. This second model points out a potential limitation in the experimental strategy. A primary assumption is that cotransfection of each of these DNA constructs results in coexpression of these constructs in each transfected cell. Another critical assumption is that alteration of the amount of competitor DNA transfected results in a corresponding alteration of competitor protein synthesized in each cell. Both assumptions are difficult to verify. Thus, while these results suggest that VP16N and VP16C utilize distinct cellular targets, clear interpretation of these results remains complex. We RelA spbdomains mechanistically similar to VP16C? The mutagenesis studies described in the previous chapter suggest that VP16C and RelA subdomains H and III share similar structural requirements for function. Specifically, both contain a similar sequence motif of critical residues: Fxx‘IAI’ (where F denotes phenylalanine and ‘1’ denotes a hydrophobic amino acid). This structural similarity raises the possibility that VP16C and RelA may be functionally similar. To examine this possibility we examined some of the fiinctional characteristics of the RelA activation domain. We chose to examine RelA using two assays previously used to examine VP16N and VP16C. The RelA activation domain displays Inr promoter preference 111 The VP16C subdomain displays core promoter specific function. Gal4-VP16N efficiently activates promoters containing either TATA or Inr elements in viva. In contrast, Gal4-VP16C activates only Inr containing promoters. We tested the sensitivity of the RelAII and RelAIII subdomains to promoter structure, as described for previous experiments with Gal4-VP16 fusions. A Gal4-RelAII expression vector was cotransfected with either the TATA or Inr-containing reporter plasmid. The Gal4-RelAII results are shown in Figure 8A. The RelAII fusion displayed transcriptional activity only when cotransfected with the Inr reporter plasmid (lane 10). No activity was detected when the TATA reporter plasmid was used (lane 5). We also tested the ability of a Gal4-RelAHI fiision protein to activate the reporter constructs. A Gal4-RelAIII expression vector was cotransfected with both the TATA and Inr reporter plasmids simultaneously. The result of the cotransfection experiments appears in Figure 8B. The Gal4-RelAIII fiision protein directs Inr transcription preferentially, as was observed with VP16C fusions. These results argue that the RelAII, RelAIII and VP16C subdomains share similar mechanism(s) of activation. RelA does not enhance DA recruitment to a promoter in vitro The previous results suggest that the RelA and VP16C activation domains may function at least in part through similar mechanisms. In addition to the promoter specificity observed in primer extension experiments, the VP16C subdomain also enhances DA complex formation. To firrther test for functional similarities between VP16C and the RelA activation domain, we evaluated RelA for this DA stabilization activity. A plasmid was constructed that encodes a fusion protein containing a full-length RelA activation 112 domain (amino acids 417-551) fused to a minimal Gal4 DNA binding domain (amino acids 1-92). The full length activation domain was chosen for two reasons. First, use of this fusion protein allows evaluation of both VP16C-like domains without the preparation of two fiision proteins. Second, the strength of the intact domain is likely to be greater than the individual subdomains, improving the potential signal strength. This efi'ect was previously observed with VP16. Despite the fact that VP16N does not stimulate DA complex formation, the full length VP16 activation domain enhances DA complex formation more than VP16C alone (N aoko Kobayashi, personal communication). Epitope purified TFIID (0.5 to 2 uL) was incubated with recombinant TFHA either in the presence or absence of a recombinant Gal4-RelA fusion protein, and separated on an agarose gel. The results are shown in Figure 14. As seen previously, some DA complex was observed in the absence of activator (lane 4). The presence of Gal4-RelA reduced the extent of DA complex formation under all TFIID conditions examined (see lanes 7-9). This phenotype is common with activators that do not stimulate DA complex formation, as noted previously for Gal4-VP16N (Figure 11). In contrast, a Gal4-VP16C fusion protein included as a positive control significantly enhanced the extent of DA complex formation (lane 12). As observed previously, formation of the DA complex is TFHA dependent (see lanes 5, 10 and 11). These results strongly argue that RelA does not stimulate DA complex formation. Can alteration of the hydrophobic core of VP16N reconstitute VP16C activity? Although the RelA activation domain contains two subdomains, RelAII and 113 RelAIII, that share a critical sequence motif (Fxxww) with VP16C, the RelA activation domain does not stabilize DA complex formation. This result implies that the Fxx‘P‘I’ motif is insufficient to confer this in vitro activity to an activation domain. To directly test this hypothesis, we altered VP16N sequences such that the subdomain contained a sequence with this motif. Codons specifying Asp 441 and Leu 444 of VP16N were deleted, producing a primary structure that resembles VP16C with respect to the positioning of the acidic and hydrophobic residues within the sequence. For reference, the sequences of the core regions of VP16N, VP16C, and the altered subdomain (VP16NAD/L) appear in Figure 15. We assayed a Gal4-VP16NAD/L firsion protein for the ability to stabilize DA complex formation. As outlined above for the RelA agarose EMSA assay, epitope purified TFIID (0.5 to 2 1.1L) was incubated with recombinant TFHA either in the presence or absence of a recombinant Gal4-VP16NAD/L firsion protein, followed by separation on an agarose gel. The results are shown in Figure 16. As observed previously, high TFIID levels allowed some DA complex formation even without activator (see lanes 2-4). The presence of Gal4-VP16NAD/L reduced the extent of DA complex formation (see lanes 7-9). Under identical conditions, a Gal4-VP16C fusion protein enhanced DA complex formation (lane 12). Formation of the DA complex was in all cases, TFIIA dependent (see lanes 5 and 10). These results confirm the conclusions suggested by the RelA DA complex experiments. The presence of the F xx‘P‘I’ primary structural element does not confer the capacity to stabilize formation of a ternary DA promoter complex in vitro. 114 RelA VP16C a, D D D g A A.- 0. AAA AAA A Figure 14. The RelA activation domain does not stimulate DA complex formation. Gal4-RelA (lanes 6-10) or Gal4-VP16C (lanes 11-13) was incubated in the presence of TFIID and/or TFIIA (as indicated by the text D or A; 20 ng) with a radiolabeled adenovirus E4T promoter fragment. Samples were separated on 1.4% agarose gel in a Mg—containing buffer. TFIID quantities were varied (0.5 11L, 1 uL, and 2 uL) as indicated by the filled triangles. Two microliters of TFIID and 20 ng of TFIIA readily formed DA complex in the absence of activator (lane 4). The presence of Gal4-VP16C enhanced the formation of this complex (lane 12). In contrast, Gal4-RelA reduced the extent of DA complex formation (lanes 7-9). In all cases, TFIIA was required to observe significant complex formation (see lanes 5, 10 and 11). VP16N: (Amino acids 439 to 447) LDDFDLDML VP16C: (Amino acids 470 to 479) MADFEFEQMF VPl 6N AD/ L: (Amino acids 439 to 447) LDFDDML Figure 15. Deletion of VP16N amino acids D441 and L444 produces a VP16C-like core motif. The amino acid sequences of VP16N and VP16C are represented by single- letter amino acid code. Codons 441 and 444 were deleted from the VP16N core sequence to produce the VP16N mutant VP16NAD/L (bottom line). This core sequence resembles the VP16C core sequence, but contains VP16N sequences flanking the altered core sequence. 116 VP16AD/L D D C D (D '8“- 5 A Figure 16. A VP16N mutant, VP16NAD/L,does not stimulate DA complex formation. Gal4-VP16NAD/L (lanes 6-11) or Gal4-VP16C (lanes 12) was incubated in the presence of TFIID and/or TFIIA (as indicated by the text D or A; 20 ng) with a radiolabeled adenovirus E4T promoter fragment. Samples were separated on 1.4% agarose gel in a Mg-containing TBE buffer. TFIID quantities were varied (0.5 uL, l uL, and 2 uL) as indicated by the filled triangles. Two microliters of TFIID and 20 ng of TFIIA readily formed DA complex in the absence of activator (lane 4). The presence of Gal4-VP16C enhanced the formation of this complex (lane 12). In contrast, Gal4-VP16NAD/L reduced the extent of DA complex formation (lanes 7-9). In all cases, TFIIA was required to observe significant complex formation (see lanes 5 and 10). TFIIA alone did not produce an observable shift (lane 11). ll7 Discussion and Conclusions The results presented in the previous chapter highlight the critical features of the primary structures of the VP16N and VP16C subdomains, identifying similar features essential for fiinction. However, minor differences in activation domain structure were also observed. Specifically, the spacing of hydrophobic residues about the central critical phenylalanine residue in the two subdomains differ. Sequence and mutational analysis of RelA identified a similar motif in two segments of the activation domain. These motifs within RelA more closely resemble VP16C than VP16N suggesting a possible fiinctional parallel between RelA and VP16C. Sequence analysis and mutational results identified a motif (Fxx‘P‘I’) within VP16C, RelAII and RelAIII that is essential for function. We were interested in determining whether or not the subtle structural differences in the VP16N and VP16C subdomains were paralleled by differences in fimction. Does this difference, however subtle, confer distinguishable fimction to each subdomain? Alternatively, the VP16N and VP16C subdomains may be functionally redundant. The structural differences observed between the two subdomains may simply reflect an inherent flexibility in the structural requirements for activation domain fiinction. As an initial means to assess the possibility that VP16N and VP16C activate through distinct mechanisms, we tested the relative activity of VP16N and VP16C in the cell line NII-1208, which contains an integrated pGSBCAT reporter. The VP16N but not the VP16C fusion proteins displayed robust activity in NIH208 cells. In contrast, both activators strongly stimulated transcription of a cotransfected pGSBCAT reporter in the parental NIH3 T3 cell line. This result suggests a difference in the capacity of VP16N and VP16C to efiiciently antagonize the repressive effects of chromatin in viva. 118 We also examined the capability of VP16N and VP16C to stimulate transcription from two reporters with different basal promoter architecture. Results from the Smale (Emami, K.H. et al., 1995) and Gralla (Chang, C. and JD. Gralla, 1993) laboratories implied that the VP16N and VP16C subdomains differ in their capacity to mediate transcriptional activation through different promoter elements. Analysis of VP16N and VP16C subdomains by cotransfection of two reporters containing either a TATA element or Inr element confirmed this inference. While either activator efficiently transactivated the Inr-containing reporter, only VP16N activated the TATA reporter. These results are also consistent with the idea that VP16N and VP16C mediate transcriptional activation by distinct mechanisms. Another result reinforces this conclusion. While neither previous observation suggests a specific mechanism for either subdomain, experiments performed in collaboration with Berk and coworkers suggest a mechanism for VP16C in vitro (Kobayashi, N. et al., 1998). Lieberman et a1. developed an in vitro assay for TFIID recruitment (Lieberman, PM. and A.J. Berk, 1994). This assay employs recombinant TFIIA and immunoaffinity purified native TFIID complex. Formation of the DA complex in this assay system is highly TFIIA dependent and can be stimulated by the Epstein-Barr virus activator Zta (Lieberman, P., 1994; Kobayashi, N. et al., 1995) and HSV VP16. The VP16 activity has been localized to the VP16C subdomain (Kobayashi, N. et al., 1995). Using a panel of our VP16C mutants, Kobayashi et al. (Kobayashi, N. et al., 1998) established a correlation between activation potential in viva and DA complex formation in vitro. A correlation was also observed between the capacity to stimulate DA complex formation and TFHA binding by these mutants. These experiments suggest that VP16C 119 activates transcription through TFIID recruitment to promoters via a TFIIA dependent mechanism. Furthermore, the correlation between the in viva and in vitro phenotypes of the VP16C mutants implies that this mechanism may be significant in viva. This result is also interesting with regard to the observation that VP16C displays Inr-containing promoter specificity. The in vitro EMSA results establish that VP16C is capable of recruiting TFIID to a promoter in a TFIIA dependent fashion. Since the primary sequence specific DNA-binding subunit of TFIID is TBP, removal of TATA sequences would be expected to markedly reduce TFIID affinity for TATA-less promoters. An activator that recruits TFIIA could stabilize TBP interaction with non- TATA sequences by facilitating a TATA-like interaction of TBP with non—TATA sequences. If the VP16C subdomain fimctions by recruitment of TFHA, this could explain the preference for Inr-containing promoters. In the presence of a TATA box, TFIIA activity may be of lesser importance due to the presence of already strong TBP-DNA interactions. This would result in activators that recruit TFHA having a reduced effect on PIC stability. This model would predict that TATA-containing promoters do not require TFHA and are not activated by TFHA-recruiting activators. In contrast, Inr-containing factors, limited by their afiinity for TFIID, would be highly dependent upon TFHA and TFHA-recruiting activators. The DA complex assay employs an agarose gel EMSA separation of nucleoprotein complexes in the presence of Mg”. This assay system appears to be particularly sensitive to the stability of the DA complex, as prolonged electrophoresis appears to dissociate the complex significantly. This suggests that some activators may function by stabilizing a TFIID-TFIIA-DNA complex, rather than increasing the rate of TFIID recruitment. While 120 the precise nature of this stabilized complex is unknown, it is interesting to note that Chambon and coworkers (White, J. et al., 1992) implicated an early step in PIC formation as a step stimulated by VP16. This "template-committed" complex was sufficient to enhance transcription even after blockage of the VP16 activation domain by a monoclonal antibody. Carey and colleagues (Chi, T. and M. Carey, 1996) have also suggested that formation of an altered TFIID-DNA complex, as evidenced by an extended DNAseI footprint, correlates with the enhancement of activation provided by Zta. This event also requires TFIIA. It is inviting to speculate that all three of these observations may be related. Regardless of the identity of the EMSA stabilized DA complex, the assay provides a USBfiJl means to determine the fimctional similarity of activation domains. We utilized the DA assay to assess the firnction of the subdomains of VP16 as well as the RelA activation domain. The VP16C and not the VP16N subdomain efficiently stimulates DA complex formation. The RelA activation domain contains two sequence segments that contain a VP16C-like Fxx‘P‘P motif. When a Gal4-RelA full length activation domain was tested for DA complex stabilization, the activator actually reduced DA complex formation. This phenotype is common for activators that do not stimulate DA complex formation, and is also exhibited by a Gal4-VP16N fusion protein. This result demonstrates that the Fxx‘P‘I’ motifs present in the RelA activation domain are not sufficient to confer DA complex stabilization activity to the activation domain. Another observation supports the conclusion that the Fxx‘P‘P motif is not a sufiicient determinant for DA complex stabilization. A VP16N mutant was designed to simulate the primary structure of the VP16C subdomain. Deletion of D441 and L444 of VP16N produces a VP16N mutant (VP16AD/L) with a primary structure that resembles 121 the Fxx‘P‘I’ motif of VP16C (see Figure 15 for sequences). The Gal4-VP16NAD/L fusion protein did not enhance DA complex stability. This result confirms that the Fxx‘I’T motif is not sufiicient to confer DA complex stabilization activity. Collectively, these results suggest that while the Fxx‘P‘P motif may be necessary for transcriptional activity, it is not sufficient for specifying the DA complex stimulatory activity. Although the RelA activation domain did not stimulate DA complex formation, the RelAIII subdomain did resemble VP16C with regard to the promoter specificity test. Both Gal4-VP16C and Gal4-RelAIII fiisions directed transcription from an Inr-containing promoter, and not from a TATA-containing promoter in viva. In contrast, Gal4-VP16N stimulated transcription from either an Inr-containing or TATA-containing promoter. The lack of a correlation between the promoter specificity results and DA complex results is surprising. The capability of the VP16C subdomain to stabilize TFIID recognition of a promoter via TFHA recruitment seems to present a reasonable explanation for the observed Inr specificity (see above). The observed Inr-specificity and lack of DA complex stabilization activity observed with RelA suggests that these two activities may not be related. This conclusion is reinforced by inspection of the Adenovirus E4 promoter (the promoter used for the DA complex EMSA studies). The AdE4 promoter contains a thymidine-rich tract of nucleotides at the initiation site. This sequence does not resemble the TdT Inr sequence (Smale, ST, 1997). Since the DA complex EMSA assay is performed using an Int-lacking promoter, it is unlikely that this activity can be described as Inr-specific. Consistent with this objection, Smale and colleagues failed to demonstrate TFIID recognition of a TdT Inr promoter using this DA complex EMSA system (Emami, 122 K.H. et al., 1997). However, these studies were performed in the absence of activators. It remains possible that an activator will facilitate TFIID recognition of an Inr containing promoter in vitro. Another line of investigation suggests a link between Inr function, TAFS, and TFIIA. The Tjian laboratory obtained evidence that TFHA may facilitate promoter specific function in a TAF dependent fashion (Hansen, SK. and Tjian, R., 1995). The Drosophila Adh promoter is transcribed from two tandem promoters. These promoters are differentially utilized temporally throughout development. This promoter specific function can be reconstituted in vitro from appropriate stage embryos. The stage specific function of the distal promoter is Inr element dependent. In vitro this Inr-mediated specificity is dependent upon TAFs and TFIIA (Hansen, SK. and R. Tjian, 1995). These results suggest that Inr function is mediated by TFHA through TAFs In summary, our experiments lead to the following conclusions. The VP16N and VP16C subdomains display functional properties both in viva and in vitro that suggest distinct mechanisms of fiinction. These include distinct basal promoter architecture preferences, differential performance on an integrated reporter, and a differential capacity to stabilize TFIID promoter recognition through TFIIA recruitment. The RelA activation domain, which contains two Fxx‘I’fl’ motifs similar to VP16C (one in RelAH and one in RelAIII), displays the same promoter specificity as VP16C. Despite this common promoter-specificity, the RelA activation domain lacks the ability to stimulate DA complex formation, an observation inconsistent with the supposed similarity of function between VP16C and RelA. Furthermore, a VP16N mutant (VP16NAD/L) which contains a FxxW motif similar to VP16C does not direct DA complex stabilization. These results 123 suggest that while the identified F xx‘IJ‘I’ motif may be necessary for activity, it is insufficient to confer specific fimction to an activation domain. Experimental Methods Recombinant DNA Manipulations The pNLS-VP16N fusion vector was produced by modification of the plasmid VP16(N) previously constructed by Tasset et al. (Tasset, D. et al., 1990). This vector contains VP16 410-490 sequences. To delete VP16C sequences, the plasmid was cut with Smal and KpnI. The digested DNA was blunted with T4 DNA polymerase, followed by agarose gel electrophoresis and ligation of the plasmid DNA to recircularize with T4 DNA ligase. pNLS-VP16C was prepared from VP16(N) by deletion of VP16 410-452 sequences. The VP16(N) plasmid was digested with AvaI. AvaI cleaves both at the SV40 NLS junction as well as at the VP16N-VP16C boundary (VP16 codon 453). The 5' overhangs were filled with Klenow enzyme. Plasmid DNA was purified by agarose gel electrophoresis and recircularized with T4 DNA ligase. The resulting constructs express either an NLS-VP16N or NLS-VP16C fusion from the SV40 early enhancer. The alternate promoter reporter constructs were obtained from Steven Smale (UCLA). They consist of synthetic Oligonucleotides with appropriate TATA or Inr sequences inserted upstream of HSV thymidine kinase (tk) sequences. The plasmids also contain an SV40 minimal origin of replication and a polyoma fragment which encompasses the entire early transcript unit. To remove these sequences, we excised the entire polyoma coding region by BamI-II digestion and religated the linearized plasmid. 124 The VP16NAD/L E. coli expression plasmid, pG4(1-92)VP16NAD/L, was constructed by PCR using a modification of the three primer method used to construct RelA point mutants (see Methods, Chapter 2). An activation domain PCR product lacking L444 was first constructed using primer ST-27 7 that contains activation domain coding sequences lacking a L444 codon. This DNA was amplified from pSGVPN using ST-277 and a second primer ST-246, which hybridizes within SV40 polyadenylation sequences 3' of the VP16 coding sequences in pSGVPN. The product was gel purified and reamplified using ST-246 and ST-278, an Oligonucleotide containing VP16 coding sequences, but lacking both codon D441 and codon L444. This PCR product was agarose gel purified and used as a maxiprimer for amplification from the E. coli expression vector pJGRl 10-1 using PCR primers ST-246 (SV40 primer) and ST-281, a primer which contains Gal4 sequences upstream of the pJGRl 10-1 cloning junction with VP16. This PCR converted the PCR product to a Gal4(1-92)-VP16 chimeric sequence. The PCR product was digested with XhaI in the Gal4 coding sequences, and Bell within polylinker sequences at the 3' end of VP16 sequences in pSGVPN. The fragment was introduced into XhaI-BamHI cut pVP16CX, an E. coli Gal4-VP16C expression vector, and sequenced to verify the codon 441 and 444 deletions. The Gal4(1-92)-RelA(417-551) expression plasmid (pG4(1-92)-RelAFL) was constructed in several steps. The activation domain (amino acids 417-551) was amplified from pSGp65(307-551) (a gift of Paul Moore and Brian Cullen; Moore, RA. et al., 1993) by PCR using Oligonucleotides ST-205 and ST-206. The ST-205 Oligonucleotide contains a BamHI site placed in frame with pSG424 polylinker and RelA sequences. The ST-206 Oligonucleotide amplifies from outside of the XbaI cloning site in pSGp65(307-551). This 125 PCR product was ligated into the EcaRV polylinker site of pB S/KS+ to generate the plasmid pBSp65(417-551). The pBSp65(417-551) BamI-II-XbaI excised activation domain sequences were then subcloned into BamI-II and XbaI cut pSGp65(307-551) to generate the plasmid pPJH65(417-551), a mammalian expression vector. The E. coli expression vector was produced from pPJH65(417-551) and pJL2. pPJH65(417-551) was linearized with BamHI and the 5'-overhang was blunted by treatment with Klenow enzyme. The activation domain sequences were then excised with XbaI and gel purified by agarose electrophoresis. The E. coli expression vector pJL2 was cut with HpaI and XbaI, excising VP16 and HSV tk sequences, and the RelA activation domain sequences were introduced by ligation to generate pG4(1-92)-RelAFL. Tissue Culture Cell Manipulations All mouse L cell transfections were performed by a standard DEAE-dextran technique as described in Chapter Two. All CAT assays were also performed as described in Chapter Two. NIH 208 and NIH 3T3 cell transfections were performed by a Lipofectarrrine (GibcoBRL) technique as advised by the manufacturer. In a P60 tissue culture plate, cells were seeded at 2x105 cells per plate. Cells were grown in DMEM media plus 10% PBS at 37°C in a 10% C02 environment to approximately 60%-80-% confluence. DNA (see Figure legends) was suspended in 100 uL of DMEM in a 1.5 mL microfuge tube. Lipofectamine (4 uL per transfection) was suspended in DMEM (100 uL per transfection). A volume of 100 uL of the Lipofectamine/DMEM suspension was transferred to each DNA suspension, followed by mixing of the contents by flicking of the microfuge tube several times with a finger. These suspensions were incubated at room 126 temperature approximately 30 minutes, followed by addition of 0.8 mL of serum-free DMEM. Cell cultures were rinsed with 2 mL of serum-flee DMEM, and overlaid with the diluted DNA-lipid complexes. Cell cultures were incubated at 37 °C/ 10% C02 for five hours followed by addition of 1 mL of DMEM containing 20% FBS. Lipofectamine containing media was removed following grth overnight and replaced with fresh DMEM containing 10% FBS. Cultures were harvested and analyzed after an additional 2 days of growth. RNA Isolation RNA from transfected monolayers was isolated using TRI Reagent (Molecular Research Center, Inc.) following the manufacturer’s instructions with the exception that volumes throughout were reduced two-fold for use in P60 tissue culture plates. Monolayers were aspirated to remove media. A 500 uL aliquot of TRI reagent was pipetted into each plate. The TRI Reagent was pipetted repeatedly onto the monolayer, until cell disruption was observed. Treated monolayers first displayed a glassy appearance, and were subsequently observed to detach. Detached cells were transferred to a microfuge tube. Tubes were either stored at -80°C for subsequent use or immediately lysed by incubated at room temperature for 5 minutes. Frozen cells were thawed at room temperature and incubated an additional 15 minutes to be certain cellular dissociation had occurred. A 100 uL volume of chloroform was added to each lysate. Tubes were vortexed 15 seconds and centrifuged at 14K in an Eppendorf fixed angle microfilge for 15 minutes at 4°C. The upper, aqueous phase of ~280 uL was collected from each sample, and 500 uL of isopropanol was added to each tube. Tubes were incubated at room 127 temperature five to ten minutes and RNA was precipitated by centrifiigation in a fixed angle microfuge at 14K for 8 minutes at 4°C. Isolated RNA was washed with 80% ethanol and dissolved in 20 uL of RNAse free TE buffer. Primer Extension Analysis Preparation of the radiolabeled probe was performed using T4 DNA kinase. A 10 pmol aliquot of Oligonucleotide ST-136 was dried in a Speed-vac apparatus to remove residual acetonitrile. The Oligonucleotide was dissolved in 7 uL of water and 30 uCi 32P- y-adenosine-triphosphate (8000 Ci/mmol; New England Nuclear; Amersham), 10 units of T4 kinase (GibcoBRL), and 3 uL of 5x Forward Reaction Buffer (GibcoBRL) were added (15 [IL total volume). The solution was incubated at 37°C for 1 hour. After incubation, the reaction was stopped by addition of 2 DL of 0.5 M EDTA and 50 uL of TE buffer. A DE-52 (Whatman) column was prepared in a 1 mL pipette tip. A 20 uL volume of AG 50W-X8 (BioRad) cation exchange (100-200 mesh) resin was poured onto a glass wool plug as a support matrix for the DE-52. Approximately 50 uL (bed volume) of DE-52 (Whatman) was overlaid and the column was equilibrated with 10 mL of TEN100 (10 mM Tris-HCl, pH 8, 1 mM EDTA, 100 mM NaCl). The labeled DNA was applied to the column and flowthrough was reloaded onto the column a second time. The column was then washed with 1 mL of TEN100 twice, followed by 0.5 mL of TEN300 (TENlOO buffer with 300 mM NaCl). Oligonucleotide was eluted with 400 uL of TEN600 (TEN100 with 600 mM NaCl). 128 RNA was analyzed by primer extension analysis. Each RNA sample (10 uL) was hybridized with radiolabeled Oligonucleotide ST-136 (3.5 uL; prepared as above) in a 15 uL final volume of 10 mM Tris-HCl, pH 8.3, containing 150 mM KCl, and 1 mM EDTA. These samples were immersed in a 65°C water bath. After 1 hour, samples were removed and allowed to cool to room temperature (30 minutes-1 hour). Samples were extended with AMV Reverse Transcriptase (Life Sciences, Inc., St. Petersburg, FL) in a 45 uL reaction containing (final concentrations) 20 mM Tris-HCl, pH 8.3, 10 mM MgC12, 6 mM DTT, 150 mg/mL actinomycin D (Boehringer—Mannheim), 0.3 mM each of dATP, dCTP, dTTP and dGTP, and 5 units of AMV reverse transcriptase. Extensions were performed for 1 hour at 42°C. Following extension, 100 uL of an RNAse solution (20 ug/mL RNAse A, 0.1 mg/mL herring sperm DNA) was added to each extension. RNAse incubations were continued at 37°C for 15 minutes. A 15 uL volume of 3M NaCH3COOH, pH 5.2 was added to each sample, followed by phenol/chloroform extraction and ethanol precipitation of the final products. Precipitated DNA was washed with 70% ethanol, air dried ten minutes and dissolved in 10 uL of loading bufl‘er (0.05% (w/v) bromophenol blue, 0.05% (w/v) xylene cyanol, 20 mM EDTA, pH 8 dissolved in formamide). Samples were separated on a 9% PAGE gel containing 7M Urea for 1 hour and 50 minutes at a constant power of 60 Watts. Gels were dried and visualized by autoradiography. Preparation of Gal4-Fusion Proteins Gal4-VP16 filsion proteins (VP16N, VP16C, and VP16NAD/L fiisions) were prepared by selective precipitation and ion exchange chromatographies. Cultures (4 x IL 129 cultures) of E. coli transformants of pJGRl 10-1 (Gal4-VP16N), pJGR81-ASma (Gal4- VP16C), or pVP16NXAD/L (Gal4-VP16NAD/L) expression vectors were grown on LB medium containing 80 ug/mL ampicillin at 37°C to an OD500 of 0.7. Expression was induced by addition of 240 mg of IPTG per liter of culture. One mL of 0.1 M Zn(CH3COO)2 was also added to supply sufficient Zn” for Gal4 DNA binding domain coordination. After an additional 3 hours of grth at 37°C, cells were harvested by centrifugation at 5,000 rpm in a Sorvall GSA rotor for 10 minutes. Supernatant was decanted and cells were either immediately lysed or stored at -80°C for subsequent use. Lysis was performed by sonication. Cells were resuspended in 160 mL of Buffer A (20 mM HEPES, pH 7.5, 10 DM Zn(CH3COO)2, 20 mM B-mercaptoethanol) containing 200 mM NaCl, and protease inhibitors (2 ug/mL Pepstatin A, 2 ug/mL Leupeptin, 1 mM benzamidine and 1 mM phenylmethylsulfonylfluoride). Cells were sonicated at 4°C at fiill power with a flat tip probe in 30 second bursts with 1 minute rest between each cycle. Sonication was repeated until greater than 90% of all cells were lysed as determined by comparison of OD600 before and after sonication. The extract was clarified by centrifiigation at 10,000 rpm in a Sorvall GSA rotor for 30 minutes (all subsequent centrifugation steps were performed at 4°C unless noted). The supernatant was retained and clarification was repeated if necessary. The volume was measured and polyethyleneimine (PEI) was added to a final concentration of 0.3% (v/v) while stirring at 4°C. The extract was stirred 15 minutes at 4°C. The PEI and associated material was precipitated by centrifugation at 10,000 rpm in a GSA rotor for 15 minutes. The supernatant was discarded and the PEI pellet was resuspended in 100 mL of Buffer A 130 containing 200 mM NaCl and protease inhibitors using a rubber policeman. The washed PEI pellet was reprecipitated from this suspension by centrifugation at 10,000 rpm in a GSA rotor for 15 minutes. This pellet was then resuspended in 100 mL of Bufi‘er A containing 750 mM NaCl and protease inhibitors to elute the fusion protein. The PEI was pelleted at 10,000 rpm in a GSA rotor, and the supernatant containing the Gal4-VP16 fusions was retained. Ammonium sulfate was added to 35% saturation while stirring at 4°C. This solution was stirred one hour to overnight. Protein was precipitated in a Sorvall SS-34 rotor at 10,000 rpm for 30 minutes. Precipitated protein was dissolved in approximately 30 mL of Buffer A containing 200 mM NaCl and protease inhibitors. The dissolved protein fraction was dialyzed against SCB (20 mM HEPES, pH 7.5, 10 M Zn(CH3COO)2, 1 mM DTT) containing 0.1 M NaCl. The protein fiaction was loaded at a flow rate of 10 mL/hr onto a Whatman P-ll phosphocellulose column (bed volume approximately 10 mL) equilibrated in SCB containing 0.1 M NaCl. Following loading, the column was washed overnight at 20 mL/hr with SCB + 0.1 M NaCl. The protein was eluted at a flow rate of 30 mL/hr with a linear gradient (4x column volume) from 0.1 to 1.0 M NaCl in SCB. Fractions containing Gal4-fusions were identified by SDS-PAGE and Western blot using a Gal4-VP16 reactive rabbit antiserum, LA2-3. Gal4-RelA fUSiOIIS were further purified by Heparin-afinity and DEAE-ion exchange chromatography. Protein fi'actions eluted from the phosphocellulose column were dialyzed against Buffer SCB containing 0.1 M NaCl. Protein was loaded at a flow rate of 15 mL/hr onto a 4 mL Heparin-agarose column equilibrated with SCB containing 0.1 M NaCl, and washed with the same buffer. Protein was eluted with a linear NaCl gradient (4x column volume) from 0.1 M to 1.0 M NaCl in Buffer SCB. Samples were 131 analyzed by SDS-PAGE. Samples that contained Gal4-RelA were pooled for subsequent chromatography. The pool was dialyzed against Buffer SzCB (20 mM HEPES, pH 7.9, 10 M Zn(CH3COO)2, 1 mM DTT) containing 0.1 M NaCl. The protein was loaded at a flow rate of 10 mL/hr onto a 3 mL Whatman DE-52 column equilibrated with SzCB containing 0.1 M NaCl. The column was washed with the same buffer and eluted using a linear gradient (4x column volume) of 0.1 to 1.0M NaCl in Bufi‘er $2CB. Samples were analyzed by SDS-PAGE. Aliquots were frozen in liquid nitrogen and stored at -80°C. Preparation of eTFIID The TFIID was epitope purified by irnmunopurification from a stable cell line prepared for this purpose. The cell line (LTRoc3) expresses a TBP gene with an N- terminal fiision of the influenza virus Hemagluttinin (HA) epitope. This small synthetic tag allows recognition of the TBP fusion by the monoclonal antibody 12CA5. The immunopurified TFIID contains an apparently full compliment of TAFs (Zhou, Q. et al., 1993) The LTRor3 cell line was grown in SMEM (Gibco/BRL) + 1% glucose, supplemented with 1x Pen/ Strep (Gibco/BRL) in spinner flasks. Cells were counted using a hemacytometer and daily diluted to 4 x 107 cells/ mL of media. LTRa3 cells were harvested at a volume of 15-30 liters by centrifugation at 2,000 rpm in a swinging bucket rotor for 10 minutes. Following harvest, cell pellets were pooled and pelleted in a conical bottom flask to measure cell pellet volume. The LTRor3 pellet was suspended in five times this pellet volume of phosphate buffered saline (PBS). PBS washed cells were centrifuged at 2,000 rpm and the PBS was discarded. The cell pellet was resuspended in 132 five volumes of Buffer A (10 mM HEPES pH 7.9, 1.5 mM MgC12, 10 mM KCl, 10 mM B-mercaptoethanol). The cell suspension was incubated at 4°C for 10 minutes, followed by centrifugation at 2,000 rpm for 10 minutes. At this stage the cell pellet swelled approximately 2-fold in volume. The Buffer A was aspirated and LTRor3 cells were suspended in 2 pellet volumes (Qflginal pellet volume) of fresh Buffer A. The cell pellet was lysed by Dounce homogenization using a Kontes all glass homogenizer with a B type pestle. Lysis was confirmed by microscopy. Crude nuclei were harvested by centrifiigation at 25,000 x g in an SS-34 rotor for 20 min. at 4°C. A 5/3 volume (original cell pellet volume) of Buffer C (20 mM HEPES pH 7.9, 25% glycerol, 0.42M NaCl, 1.5 mM MgC12, 0.2 mM EDTA, 0.5 mM PMSF, 10 mM B-mercaptoethanol) was added to the crude nuclei pellet. The pellet (at this stage gelatinous) was resuspended by Dounce homogenization 10 times. Following resuspension, the sample was nutated 30 min. at 4°C, followed by clarification at 25,000 x g in an SS-34 rotor for 30 min. at 4°C. This extract was loaded onto a phosphocellulose P-ll column (~1 mL bed volume/ 20 mg of total protein) equilibrated with buffer D100 (20 mM HEPES pH 7.9, 0.2 mM EDTA, 100 mM KCl, 20% glycerol, 10 mM B-mercaptoethanol). The P-ll column was washed with 2 column bed volumes of buffer D300 (D100 with 300 mM KCl), followed by 5 column volumes of D500 (D100 with 500 mM KCl). The phosphocellulose “D fraction” was eluted with ~1.5 column volumes of D1000 (D100 with 1 M KCl). Two milliliter fractions of eluate were collected. The eTFIID containing fractions were identified by Western blot using the Hemaglutinin (HA)-epitope monoclonal 12CA5. Monoclonal 12CA5 coupled Protein A-Sepharose beads for immunoprecipitation were prepared as follows. Dry Protein A- Sepharose CL—4B (Pharrnacia; 1.5 g) was 133 swollen by addition of 15 mL of water. The CL—4B media was washed four times by exchange of ~50 mL of fresh water. Swollen Protein A- Sepharose CL-4B was stored as a 50% slurry in PBS plus 20% (v/v) ethanol until subsequent use. For coupling, two milliliters of the 50% slurry was removed and decanted into a 15 mL conical plastic tube. The Sepharose was pelleted at 4000 rpm in a table top centrifiige for 5 minutes followed by removal of supernatant. The Protein A-Sepharose then was washed with 14 mL of fresh PBS, followed by centrifiigation at 4,000 rpm for 5 minutes. The PBS was aspirated, and the wash procedure was repeated four times. A 3 .5 mL volume of monoclonal antibody 12CA5 (Hemagluttinin antibody; 2 mg/mL in PBS) was added to the Protein A-Sepharose and the slurry was incubated at room temperature for two hours with gentle rocking. The Sepharose was pelleted at 4,000 rpm for 5 minutes and the supernatant was removed. Beads were washed with 20 mL of 0.2 M sodium borate, pH 9.0, followed by centrifugation at 4,000 rpm for 10 minutes. The supernatant was decanted and the beads were resuspended in 20 mL of 0.2 M sodium borate, pH 9.0 containing 0.104 g of dimethylpimelimidate (20 mM final concentration). This slurry was rocked at room temperature for 30 minutes to allow crosslinking of the 12CA5 antibody to Protein A. Beads were pelleted at 4,000 rpm for 10 minutes and washed with 20 mL of 0.2 M ethanolamine, pH 8.0 followed by a repeat of the centrifugation. The supernatant was discarded and the 12CA5-Protein A-Sepharose beads were resuspended in 20 mL of 0.2 M ethanolamine and gently rocked at room temperature for 2 hours to stop the cross- linking reaction. Crosslinked beads were pelleted at 4,000 rpm and washed with 20 mL of 0.1 M glycine-HCI pH 3.0 followed by exchange with fresh PBS. Coupled, washed beads were suspended in 2 mL of PBS containing 0.01% (w/v) merthiolate for storage until use. 134 Coupled beads were pelleted at 4,000 rpm and washed twice with 15 mL of buffer D500 (D100 with 500 mM KCl) containing only 5 mM B-mercaptoethanol to prevent light chain dissociation. The phosphocellulose D fraction was split into two, six milliliter pools. A seventy microliter bed volume of 12CA5 beads was aliquoted to each of 12 tubes and one milliliter aliquots of the P-1 1 TFIID fraction were added to each tube. These samples were nutated overnight at 4°C. Beads were precipitated by centrifugation for 5 minutes at 14,000 rpm in an Eppendorf fixed-angle microfuge, and consolidated into six tubes. Pellets were washed three times with buffer D500 containing 0.05% NP-40 and 5 mM [3- mercaptoethanol to dissociate endogenous TFHA. The supernatant was discarded. Beads were washed twice with 1 mL of D100 containing 5 mM B-mercaptoethanol, and eTFIID was eluted with buffer D100 containing 1 mg/mL HA peptide and 5 mM [3- mercaptoethanol. After a one hour incubation with gentle rocking at room temperature, holes were pierced in the bottom of each tube with a 20 gauge needle followed by centrifugation of all liquid from the tube at 14,000 rpm until the 12CA5 Sepharose beads were completely dried. Eluted eTFIID was frozen in liquid nitrogen and stored at -80°C. Preparation of His-tagged TFIIA E. coli strain M15 transformants of thTFHAor, thTFIIAB, or thTFIIAy, were grown at 37°C in LB media (1 liter) containing 100 ug/mL ampicillin and 25 ug/mL kanamycin to an OD600 of 0.7 (E. coli strain and plasmids a kind gift of Arnold Berk, UCLA). Cultures were induced with 0.5 mM IPTG and grown an additional 1.5 hours at 30°C. Cells were harvested at 5,000 rpm in a Sorvall GS-3 rotor for 5 minutes. The cell 135 pellets were used immediately or stored at -80°C. Pellets from 1 liter cultures were resuspended by manual swirling in 12 mL of Qiabuffer A (10 mM Tris base, 0.1 M NaH2P04, 6 M guanidine HCl, pH 8.0) containing 7 mM B-mercaptoethanol and 1 mM phenylmethylsulfonylfluoride (PMSF). Suspended cells were rocked at 4°C for 1 hour in Teflon Oakridge tubes. Ni-NTA resin (Qiagen) was prepared as follows. Two milliliters of a 50% slurry of resin was precipitated at 1,500 rpm for 5 minutes in a 15 mL Corex tube in a Sorvall table top centrifuge. This resin (1 mL resin bed volume) was resuspended in 12 mL of Qiabutfer A. The resin was precipitated with a second 5 minute spin. The supernatant was discarded and the resin was suspended in 12 mL of Qiabuffer F (6 M guanidine HCl, 0.2 M acetic acid). Following this wash, the resin was pelleted by a 5 minute spin. The acid washed resin was resuspended in 12 mL of Qiabufl‘er A to wash out residual acetic acid and the resin was precipitated by centrifugation for an additional 5 minutes. This wash step was performed an additional two times. The E. coli lysate was transferred to a 15 mL Corex tube containing the equilibrated Ni-NTA resin. The lysate was rocked at 4°C for 1 hour to allow Ni-His6 tag binding. Following this incubation, the resin was precipitated by centrifiigation at 1,5 00 rpm in a tabletop centrifiige. The supernatant was discarded. The resin was washed by rocking 15 minutes at 4°C, with 10 mL of Qiabuffer A containing 7 mM B- mercaptoethanol and 1 mM PMSF. The resin was then precipitated by centrifirgation and the process repeated, followed by washing of the resin by rocking in 10 mL of Qiabuffer B (10 mM Tris base, 0.1 M NaHzPO4, 8 M urea, pH 8.0) containing B-mercaptoethanol and PMSF. The resin was precipitated by centrifugation and the wash step was repeated. The 136 resin was then suspended in 5 mL of Qiabuffer B containing B-mercaptoethanol and PMSF and transferred to a disposable mini-column. Following transfer, the column was washed with an additional 10 mL of Qiabuffer B. All subsequently used chromatography bufi‘ers contained 7 mM B-mercaptoethanol and 1 mM PMSF. The column was washed with 30 mL of Qiabuffer B, adjusted to pH 6.3, followed by 12 mL of Qiabuffer B containing 10 mM imidazole. Following these wash steps, the proteins were eluted with Qiabuffer B containing 200 mM imidazole. One milliliter fractions were collected and analyzed by SDS-PAGE. TFIIA was renatured by step dialysis to remove urea. Protein was estimated by Bradford assay and the subunits were mixed in approximately equimolar amounts (1 mg of each of the large subunits and 0.5 mg of the small subunit, in an undiluted volume of approximately 5 mL). This mixture was dialyzed overnight against one liter of Bufl‘er D100 (20 mM HEPES-KOH, pH 7.9, 20% glycerol, 0.2 mM EDTA, 100 mM KCl, 7 mM B-mercaptoethanol, 1 mM PMSF) containing 2M urea. The mixture was then dialyzed against one liter of D100 containing 1M urea for 4 hours, followed by a change of dialysate to D100 plus 0.5M urea. Following an additional 4 hour dialysis against the 0.5M urea buffer, the sample was dialyzed against 3 x 1 liter of Buffer D100. Insoluble material was removed by centrifirgation at 14,000 rpm in an Eppendorf microfirge. Soluble TFIIA protein concentration was determined by a Bradford protein assay. Aliquots were frozen in liquid nitrogen and stored at -80°C. Preparation of EMSA Probes 137 The TBP-TFHA EMSA probe used was a 452 bp XhaI-HinDIII AdML promoter fragment, excised from the vector pML (obtained from Dr. Zachary Burton, Michigan State University). The fragment was gel purified and end-labeled with 32P-y-ATP by the exchange reaction using T4 DNA kinase. The agarose EMSA probe used was a 225 bp EcaRI-HinDIII fragment excised from the plasmid pG5E4TCAT (obtained from Arnold Berk, UCLA). This fragment was incubated with calf alkaline phosphatase for 30 minutes, agarose gel purified, and end-labeled with 32P-y-ATP by the forward reaction of T4 DNA kinase to Optimize specific activity of the probe. Probes were extracted with phenol/chloroform, ethanol precipitated and purified by gel filtration on a ~6cm Bio-Gel P6-DG column poured in a Pasteur pipette. One microliter samples of the 100 DL fractions were quantitated by Cerenkov counting in a scintillation counter. Specific activity was determined assuming 100% recovery of the DNA. Gal4 EMSA Assay Binding reactions were assembled in 10 mM HEPES-KOH, pH 7.8, 60 mM KCl, 12.5% glycerol, 5 mM MgC12, 10 mM B-mercaptoethanol, 1 mg/mL bovine serum albumin, 40 ug/mL polde:dC, and 500 cpm of probe (final concentrations). The probe used is a 240 bp fragment of pG5E4TCAT (see probe preparation section above) encompassing 5 upstream Gal4 sites, the adenovirus E4 TATA box and downstream promoter sequences, as well as a segment of the bacterial chloramphenicol acetyltransferase gene. Gal4-fusion proteins were diluted in buffer D400 (20mM HEPES- KOI-I, pH 7.9, 20% (v/v) glycerol, 0.2 mM EDTA, and 400 mM KCl) as indicated in text and/or legends. One microliter of each dilution was used per reaction. Incubations were 138 performed at 30°C for 30 minutes, followed by electrophoresis on a 4% PAGE gel in 1x TBE (90 mM Tris base, 90 mM boric acid, 2 mM EDTA) at 150 V for 1 hour and 45 minutes. Gels were dried under vacuum at 80°C onto Whatman 3MM paper. Results were visualized by autoradiography. TFHA-TBP EMSA Assay TFHA-TBP gel shifts were performed as described by Maldonado and Reinberg (Maldonado, E. et al., 1990). Yeast TBP (0.6 or 1.1 pmol) was incubated in binding buffer with hTFIIA (0.07-1.49 pmol) at 30°C for 30 minutes. The binding buffer contained 10 mM HEPES, pH 7.9, 4 mM MgC12, 5 mM (NH4)2SO4, 8% (v/v) glycerol, 2% (w/v) polyethyleneglycol (average molecular weight = 8000 Daltons), 50 mM KCl, 5 mM B-mercaptoethanol, 0.2 mM EDTA, 25 ug/mL poldede and l frnol of the 3’2P labeled AdML promoter DNA probe. Following the incubations, samples were loaded onto a 0.5x TBE gel and electrophoresed at 150 Volts for 1 hour and 45 minutes. Gels were dried onto Whatman 3MM paper and visualized by autoradiography. DA Complex EMSA Assay Binding reactions were assembled in 10 mM HEPES-KOH, pH 7.8, 60 mM KCl, 12.5% glycerol, 5 mM MgC12, 10 mM B-mercaptoethanol, 1 mg/mL bovine serum albumin, 40 ug/mL polde1dC, and 500 cpm of probe (final concentrations). The probe is identical to that described above (Gal4 EMSA methods section). As indicated in text and legends, reactions contained 1-2 uL of TFHD, 20 ng of TFIIA and/or Gal4-fi1sion protein (0.5x, 1.0x, 1.5x protein required to saturate all five Gal4 binding sites as determined 139 empirically by titration of Gal4-fiisions in a 4% PAGE EMSA. The Gal4-fusion proteins were dialyzed against buffer D400 (20mM HEPES-KOH, pH 7 .9, 20% (v/v) glycerol, 0.2 mM EDTA, and 400 mM KCl) prior to use. Reactions were assembled on ice, and incubated at 30°C for 30 minutes prior to electrophoresis. Samples were loaded directly without added loading dye to avoid shadowing by xylene cyanol. Electrophoresis was performed in 1x Buffer G (45 mM Tris base, 45 mM Boric acid, 0.5 mM EDTA, 5mM MgAcetate) on a 1.4% agarose (low EEO, Boehringer-Mannheim Biochemical) gel at 70V for 3 to 3.5 hours. After electrophoresis, gels were dried by vacuum at 80°C onto DEAE-cellulose paper reinforced with Whatman 3MM paper. Results were visualized by autoradiography. 140 Chapter Four Activation domains are composed of multiple, small modular elements The results described in the previous chapters provide evidence that activation domains are composite in nature, composed of repeated, perhaps related functional motifs. These essential elements are relatively small (approximately 10 residues) hydrophobic sequences. Elements fi'om different activators (e. g. VP16 and RelA) are relatively similar in terms of their common requirement for hydrophobic amino acids for function. However, such elements also display some structural distinctions; both VP16C subdomains contain a critical dipeptide (a phenylalanine followed by an acidic residue), whereas neither RelA subdomain examined contains an identical motif. Despite similar structural requirements, not all activation domains are functionally equal. This is apparent by comparison of a number of functional characteristics of VP16N and VP16C. Their variable potency in different species, their distinct capacity to derepress chromatin templates and their differential basal promoter specificities all suggest that VP16N and VP16C firnction through distinct mechanisms in viva. Perhaps more directly, in vitro studies establish that VP16C, but not VP16N, is capable of stimulating the formation of a TFIID-TFIIA-DNA (DA) complex. These results present a paradox. How do two VP16 motifs of such similar structure impart distinct functional attributes to each subdomain? One possibility is that the limited structural differences within these core motifs are responsible. To test this possibility, we constructed a VP16N mutant (VP16NAD/L) which contained a core primary structure similar to VP16C. However, this VP16NAD/L mutant did not stimulate 141 DA complex formation, arguing that the spacing of hydrophobic residues within an activation motif is not the determinant of this specific function. The RelA results also support this conclusion. We identified two motifs within the RelA activation domain that appear similar to the VP16C core motif. An alanine scan of each RelA subdomain verified these predictions in part. The fiinction of each RelA subdomain was critically dependent upon a similar cluster of hydrophobic residues. Interestingly, VP16C and RelAIII display identical promoter specificity in viva. However, the RelA activation domain is incapable of stimulating DA complex formation. These results argue that while some common mechanistic features of the RelA and VP16C activation domains may exist, DA recruitment is not responsible for the observed promoter specificity. These results illustrate that prediction of function through such structural analyses is unlikely to be fi'uitful. Clearly, the pattern of hydrophobic residues within the subdomain is not solely responsible for fimctional specificity. Yet, the alanine scan of VP16C demonstrates that the hydrophobic residues within the core FxexMF motif are primarily responsible for in viva activity. What is the role(s) of these hydrophobic residues? At least two formal possibilities exist. Either the hydrophobic residues are essential for nucleation of a critical tertiary structure within the activation domain, or the residues are involved directly in formation of a protein-protein interface between the domain and its target protein(s). The latter hypothesis seems likely given the results of stmctural studies of a number of complexes of activation domains and their proposed cellular targets. Activator-target protein-protein complexes appear predominantly hydrophobic 142 The crystal structure of a peptide from the p53 activation domain bound to MDM2 illustrates how an activation domain might complex with its cellular target (Kussie, P.H. et al., 1996). The MDM2 protein is the product of a p53 inducible gene. MDM2 binds the activation domain of p53 and masks its fimction. This interaction appears to involve the same residues of p53 essential for transactivation (Lin, J. et al., 1994), suggesting that the mode of binding of p53 to its transcriptional target may be similar to that observed in the MDM2 co-crystal. In the crystal structure, the activation domain peptide forms a short two-tum helix. Three hydrophobic residues (F19, W23 and L26) insert into a deep cleft in the MDM2 interaction domain. Each of these residues lies on the same face of the helix. Only two intermolecular hydrogen bonds are observed in the entire complex. One involves MDM2 hydrogen bonding to a p53 main-chain amide at the entrance to the cleft, and a second MDM2 hydrogen bond forms with the tryptophan indole group at the bottom of the cleft. This paucity of hydrogen bonding suggests that specific recognition must rely principally upon recognition of the hydrophobic residues. Consistent with this observation, the surface complementarity observed between p53 and MDM2 is very precise. The three critical hydrophobic residues in p53 pack against each other, producing a surface exactly complementary to the MDM2 binding pocket (Kussie, P.H. et al., 1996). The structure of the CREB-CBP complex has recently been solved by NMR (Radhakrishnan, I. et al., 1997). The CREB activator requires the coactivator CBP for full in viva function (Kwok, R.P.S. et al., 1994). Consistent with this observation, a domain of CBP (the KIX domain) specifically interacts with a segment of the CREB activation domain (pKID). The pKID-KIX interaction resembles the p53-MDM2 interaction in that the activation domain-target interface is predominantly composed of the 143 hydrophobic face of an amphipathic alpha helix, inserted into a hydrophobic channel on the surface of the KIX target protein. However, this structure also differs in some significant aspects. The pKID domain consists of two helices connected by a short flexible loop. This 100p contains a Protein Kinase A (PKA) phosphorylation site. Phosphorylation of this site is necessary for the pKID-KD( interaction (Zhong, H. et al., 1998). The N-terminal helix makes only minor contacts with the KIX domain. The C- terrninal helix is predominantly responsible for formation of the interface. This helix of pKID lies in a relatively shallow hydrophobic groove across the surface of the KDC domain packing at an angle across the surface of two KIX helices. This contrasts with the p53-MDM2 interaction, in which the p53 helix inserts into a very deep cleft between two helices. In addition to this distinction, the pKID-KD( interaction is significantly augmented by electrostatic interactions. Three ionic interactions are observed in the NMR structure. Two involve aspartic acid residues in the second helix of pKID and lysine residues in KIX, while a third occurs between an arginine residue of the first helix and a glutamate residue of the KIX domain. The PKA-phosphorylated serine residue of the central loop of pKID also hydrogen bonds to a tyrosine residue of the KIX domain. Thus, while the principal mode of interaction remains similar to the p53-MDM2 interaction, the CREB-CBP interface also contains multiple ionic interactions that contribute significantly to the specificity of the CREB-CBP interface. A third relevant structure has also been solved. The steroid receptors activate transcription from their target genes in a ligand dependent fashion. This activation is thought to occur via a ligand-induced conformational change that results in the recruitment of a family of related coactivator molecules (the p160 family of coactivators, 144 including SRC-l, GRIP-1, and ACTR). This interaction is thought to be dependent upon a common motif in the p160 proteins. The motif consists of a short LxxLL sequence. The crystal structure of the PPAR-y receptor (a steroid receptor superfamily member) complexed with a segment of SRC-l which contains two of these LxxLL motifs was recently reported (N olte, R.T. et al., 1998). The two SRC-l motifs contact each of the ligand binding domains of a PPAR-y homodimer. Interestingly, in many ways the PPAR- SRC-l interaction resembles the reverse of the p53-MDM2 and CREB-CBP interactions. In this case, the ligand-binding domain of the activator (PPAR-y) forms a globular domain that recognizes a short two-tum amphipathic alpha helix in the target (SRC-l). The remainder of SRC-l is disordered and appears to form a flexible tether between the LxxLL motifs. The SRC-l helices pack into a hydrophobic interface at the junction of four helices in the PPAR receptor ligand-binding domain. The interaction is predominantly hydrophobic in nature and is supplemented by the hydrogen bonding of two invariant receptor charged residues with the main-chain of the LxxLL motif of SRC-l at the beginning and end of the helical motif. This structure resembles a similar interaction between the thyroid hormone receptor (TRB) and the LxxLL motif of GRIP-l (another p160 family member) (Darimont, B.D. et al., 1998). Four residues of the LiorLL helix insert into a hydrophobic patch constructed of the interface of four helices of the TRB receptor. Again, the interface of the two proteins consists of a globular TRB domain that recognizes the hydrophobic face of an amphipathic helix in GRIP-l. In contrast to the CREB-CBP and p53-MDM2 complexes, the role of activator and target is reversed. The target, rather I45 than the activation domain consists of a small helical motif which is specifically recognized by a more complex globular domain (the receptor ligand-binding domain). All four of these complexes are relatively simple molecular interactions, with small interfacial areas, dominated by one of two partners. One partner (more commonly the activation domain) folds upon target recognition to form a short (2-3 turn) amphipathic helix, the hydrophobic face of which inserts into a complimentary pocket of the relatively more complex target protein. The target protein is principally responsible for specificity of recognition, providing a significantly more complex interface that is dependent upon the tertiary structure of the target protein. Induced folding of the activation domain is observed in both the p53-MDM2 (Kussie, P.H. et al., 1996) and CREB-CBP complexes (Radhakrishnan, I. et al., 1997). Specificity is achieved principally through complementary of the interacting surfaces, although additional electrostatic interactions also provide specificity to the CREB-CBP and receptor-p160 coactivator complexes. The predominance of one partner in these activator-target interactions is more reminiscent of a protein-peptide interaction than a typical protein-protein interaction (such as those observed in multisubunit enzymes and oligomeric proteins). Like the structures discussed above, protein-protein interfaces of oligomeric proteins are almost exclusively hydrophobic and contain large interfacial areas ranging in size from 1,400 to greater than 6,000 A2 in size (Janin, J. et al., 1988). In contrast, protein-peptide interactions are characteristically less hydrophobic, augmented by some additional ionic interactions (such as is observed in the KIX-pKID interaction), and contain significantly smaller interaction surfaces (Janin, J. and Chothia, C., 1990). These characteristics are necessary for transient protein-protein interactions such as antibody-antigen and protease-substrate 146 interactions, as these proteins must be stable in the absence of their partners. This requires a smaller, relatively less hydrophobic surface area that is not energetically unfavorable when exposed to solvent (Jones, S. and Thornton, J.M., 1996). These same properties would be expected of activator and target proteins. Consistent with these requirements, the interfaces observed in all three of the activator-target complexes is relatively small, ranging from 1,200 A2 for the CREB-CBP interaction to 1,500 A2 for the p53-MDM2 interface. These figures are actually slightly smaller than the mean figure reported for heterocomplexes (~1,900 A2) in a 1996 Brookhaven PDB survey performed by Jones and Thornton (Jones, S. and J .M. Thornton, 1996). However, even smaller interfaces have been observed. The SH3 domains of signal transduction proteins, which recognize a short polyproline helix, typically bury ~800 A2 of surface area upon complex formation (see review, Kuriyan, J. and Cowbum, D., 1997). The number of transient protein-protein complexes with known structure has been increasing steadily. While protease-inhibitor complexes and antibody Fab fiagments provided initial examples of such structures, advances in signal transduction research have provided a recent wealth of structural information. The structures of SH2, SH3, and PTB domains complexed with their targets have been solved (for review see Kuriyan, J. and D. Cowbum, 1997). No common theme emerges with respect to the structural requirements for these proteins. While SH3 domains recognize the hydrophobic face of a short polyproline helix (similar to the activation domain-target interfaces), the SH2 domains recognize an extended polypeptide chain. This extended chain is bound across an SH2- domain surface groove. Phosphotyrosine binds in a deep pocket at one end of the groove, and specificity is determined by the geometry and chemical composition of the groove. 147 ,Jv These properties variably favor the presence of hydrophobic or polar residues at positions C-terminal to the phosphotyrosine. This results in a binding pocket that is in essence capable of "reading" the primary structure of a short peptide motif. The PTB domains provide yet one more distinct method of protein-peptide interaction. The peptide ligand forms a [3 sheet that is incorporated into an antiparallel sheet of the PTB domain. The interaction is extensively stabilized by intermolecular hydrogen bonding between these two antiparallel B sheets. Nature has clearly found many approaches suitable for the formation of such transient protein-protein complexes. Modularity of activation domain structure In both VP16 and RelA, the activation domain appears to be modular in structure, composed of repeated motifs with similar structural requirements. Yet, these subdomains do appear to be functionally discemable. How is this "specificity" achieved? Two possibilities exist. The specificity may result from specific recognition or "readout" of the primary structure, as is observed with SH2 domains. Alternatively, the core motifs may provide the driving force of the molecular interaction, without directly contributing to specificity. The determinants of specificity may reside within surrounding sequences. Our studies seem to favor the second model. The RelA activation domain contains two amino acid segments that contain a F xx‘I’T motif that is also found in VP16C. Despite this structural similarity, the RelA activation domain does not share the capacity to stimulate formation of a promoter DA complex. Furthermore, a VP16N mutant engineered to contain a hydrophobic core with a core motif similar to that observed in VP16C does not 148 display this property either. These results argue that the precise primary structure of the Fxx‘I’T core motif is not a sufiicient determinant for specific function. Given these results it seems likely that the primary determinants of specificity reside outside of the core motif. Studies performed on the steroid receptor coactivators SRC-l and GRIPl provide a precedent that could be applicable to other activation domains. Darimont, et al. describe a set of biochemical and crystallographic studies of the GRIPl-TRB interaction (Darimont, B.D. et al., 1998). Sequences both N-terminal and C- terminal to the LxxLL core are required for high affinity binding of GRIPl to TRB. Despite this contribution, crystallographic studies reveal no direct interaction between these regions and the ligand-binding domain. Rather, the authors attribute the effects of flanking sequences to favorable electrostatics between GRIP] residues and the ligand- binding domain. 1 A role for core-flanking sequences in specific steroid receptor recognition has also been observed with SRC-l. McInerney, et al. tested the capacity of rnicroinjected SRC-l mutants to mediate receptor fimction in viva (McInemey, E.M. et al., 1998). Both the LxxLL core and flanking sequences were found critical for function in viva. Interestingly, mutations affecting specific residues C-terminal of one core motif of SRC-l differentially affected coactivation of the estrogen and retinoic acid receptors. Apparently, distinct residues within these C-terminal sequences differentially contribute to coactivation of specific receptors. Similar to the situation observed with the TRB receptor, the PPAR}!- SRC-l crystal structure failed to discern any direct contact between these flanking sequences and PPARy. 149 Signal transduction proteins provide another example of a strategy employed by nature to provide specificity to these protein-protein interactions. Like the activation domain-target complexes, SH2 and SH3 domains recognize small peptide motifs. Because of the small size of these interfaces, the strength of these complexes is limited, with typical complexes displaying high nanomolar to micromolar dissociation constants. These properties result in weak, transient interactions, which these proteins appear to have adapted to their advantage (reviewed in Kuriyan, J. and D. Cowbum, 1997). Analysis of the kinetics of SH2-phosphopeptide interactions reveals rapid on and off rates, allowing the rapid and repetitive sampling of alternate sites to facilitate the process of selective pairing of SHZ domains and phosphorylated targets. These signaling proteins share another property with the transcriptional activators; they are modular in structure. This modularity is thought to enhance both affinity and specificity, through cooperative function of individual domains. One example of this cooperativity is seen in T cell receptor signaling. The T-cell receptor is composed of several polypeptide chains, which contain specific phosphotyrosine motifs (ITAM motifs). The ITAM motifs are recognized by ZAP-70, a tyrosine kinase that contains two SHZ domains that are connected by a rigid linker. The juxtapositioning of the two domains provides a combinatorial ITAM/ZAP-70 interface that is specific for a very precise head to tail repeat of two phosphotyrosine motifs. This leads to cooperativity between the two protein-protein contacts (Hatada, M.H. et al., 1995). A similar example is provided by the SH-PTP2 tyrosine phosphatase. Binding of peptides to the SH2 domains of this protein leads to activation of the phosphatase. ,While single phosphotyrosine peptides can activate the phosphatase, activation is dramatically enhanced by peptides that contain tandem 150 phosphotyrosine motifs. The structure of SH2-PTP2 reveals that the two SH2 domains are rigidly constrained, once more selecting for the precise positioning of two phosphotyrosine motifs for optimal effect (Eck, M.J. et al., 1996). The GRB2 adapter provides yet one more example of the repetitive, modular structure of these proteins. GRB2 contains one SH2 and two SH3 motifs. In contrast to the previous two examples, the domains are flexibly linked, providing a molecule suited to more promiscuous interactions (Maignan, S. et al., 1995). Thus, these signaling molecules have evolved to combine multiple peptide recognition domains in specific structural combinations to facilitate either the flexibility or specificity of these protein-protein interactions. The structure of activation domains suggests that activation domain targets may exploit a similar strategy. RelA provides a good example of the modularity of activation domains. The domain is separable into three, independently functional subdomains. Each subdomain is similar in terms of required structural properties (all are highly dependent upon critical hydrophobic residues) and each is well conserved evolutionarily. Perhaps the activation domain has evolved by duplication of a common ancestral motif followed by specialization of these duplicated sequences into individual, distinct subdomains. These subdomains may cooperate to recognize a single target with enhanced affinity, or may have evolved to recognize independent targets. The VP16 activation domain is also composed of repeated, fiinctionally distinct subdomains. Multimerization of these motifs may enhance activation domain potency by simultaneously allowing the activator to recruit multiple targets to regulated promoters. This model could explain the greater than additive efi‘ect, or synergy, of multiple activators to transcription (Chi, T.H. et al., 1995). Consistent with this model, He and Weintraub 151 report that multirnerization of two activation domains of distinct fimction (VP16N and an Spl activation domain fragment) results in synergistic increases in activator potency (He, S. and SJ. Weintraub, 1998). The fimctional differences between the VP16N and VP16C subdomains seem to argue that the VP16N and VP16C subdomains may fiinction in a similar, complementary fashion. However, Emami et al. demonstrated that a similar efi‘ect is observed upon multirnerization of two or more VP16N subdomains on a Gal4 chimeric protein (Emami, K.H. and M. Carey, 1992). Thus, strictly speaking, synergy does not require two distinct subdomains. Perhaps, the VP16N (and other subdomains) may have evolved to fillfill several mechanistic roles. The cooperative nature of activator-target interactions is underscored by studies performed with the CREB activator. CREB provides an excellent model for activator fimctional studies. ' A functionally relevant target (CBP) has been identified and the mechanism of CREB-CBP interaction is understood in molecular detail (see above, Radhakrishnan, I. et al., 1997). Despite the well-characterized nature of the CREB-CBP interaction, it is clear that CREB-CBP interaction alone is insufficient for CREB activation. Phosphorylation dependent, CBP-dependent CREB activation can be reconstituted in vitro in HeLa extracts. The CREB activation domain is also bipartite. The pKID domain (discussed above) is responsible for CBP interaction. A second glutamine-rich domain, Q2, was found to bind dTAFn110 (F erreri, K. et al., 1994). Both interactions appear to required for activation in vitro, as antibodies against hTAFn130 blocked CREB activation (Nakajima, T. et al., 1997). The coactivator CBP provides another relevant observation illustrating that cooperativity of distinct activation domains results in transcriptional synergy. The Spl 152 and SREBP-la (Sterol-Response-Element binding protein) activators synergize to activate the LDLR (low density lipapratein receptor) promoter in vitro. Synergistic activation is observed in a crude nuclear extract, but is not observed in a transcription reaction reconstituted with purified basal factors. Robust synergy can be reproduced using a chromatin template in the presence of a partially purified CBP complex. This synergistic activity is dependent upon CBP, as CBP peptides can block this activation. Furthermore, TAFS are required for this synergy as well, since TBP cannot replace TFIID in the transcription reactions (N aar, AM. et al., 1998). These results imply that efficient synergy of these two activators requires multiple coactivator components (CBP and TAFS). The VP16C results suggest a potential mechanism for VP16N and VP16C synergy. The VP16N subdomain efiiciently derepresses an integrated reporter. In contrast, VP16C is inemcient when assayed from an integrated, chromosomal reporter. In contrast, VP16C stimulates TFIID recruitment via TFHA in vitro. This activity is not present in VP16N. VP16C also displays promoter specific function, requiring an Inr element for function in viva. Conceptually, these two observations could be related. The primary defect expected for a TATA-less promoter would be a reduced ability to recruit TFHD to the promoter. Thus, an activator that stimulates TFIID recruitment would be expected to enhance TATA-less promoter function. If this activity was unnecessary due to the presence of a consensus TATA sequences, a TFHA-dependent activator (VP16C) would be expected to provide little activation. Given this model, the VP16C Inr- preference can be rationalized by its TFHA-dependent TFIID (DA) recruitment activity. 153 The two subdomains of VP16 may synergize to produce a highly potent activation domain due to these two complementary activities. This model suffers from a significant deficiency. The Adenovirus E4 promoter utilized for the DA recruitment assays does not contain an Inr element. Furthermore, Emami, et al. failed to detect stable TFIID binding to an Inr promoter in vitro, even in the presence of TFHA (Emami, K.H. et al., 1997). This is perhaps not surprising. While TFIIA does bind DNA in the TBP-TFIIA-DNA complex, these DNA-interactions are relatively limited (Geiger, J .H. et al., 1996). Perhaps TFIIA, being unable to specifically recognize promoter sequences, requires recruitment via a sequence-specific activator. It may prove interesting to test the ability of Gal4-VP16C to enhance TFIID recognition of an Inr promoter in the DA agarose EMSA. Results from experiments performed by Hansen and Tjian also bolster this hypothesis (Hansen, SK. and R. Tjian, 1995). Hansen found that TFHA, in conjunction with TAFS, mediate core-promoter specificity of Adh promoter usage in Drosophila. The TFIIA preferred promoter contains an Inr element critical for the observed specificity. These results imply that TFIIA recruitment could conceivably mediate Inr-specificity. Recent developments with RelA are also of potential interest. RelA is a NF-KB family member. These transcription factors are regulated by multiple mechanisms. Primary control is maintained through a mechanism of cellular localization. A family of IKB proteins masks the nuclear localization signal and tethers the NF-KB proteins in the cytoplasm. In response to a wide variety of signal transduction pathways, the IKB proteins are degraded by proteolysis allowing nuclear translocation of the NF-KB transcription factors. In addition to this mechanism, NF-KB is also regulated by 154 phosphorylation. A subset of cellular protein kinase A (PKA) is found in a stable complex with NF-KB and IKB (Zhong, H. et al., 1997). Activation of this PKA species occurs by a mechanism independent of CAMP levels. The cellular coactivator CBP also potentiates RelA transcriptional activity. The CBP protein was originally described as a necessary adapter for CREB transactivation. The structure of the CREB-CBP interface has been described above. Binding of pKID to KIX is dependent upon phosphorylation of a critical serine residue. Similarly, transactivation by RelA is stimulated by phosphorylation of Ser 276 within the rel homology domain (Zhong, H. et al., 1998). The crystal structure of the rel homology domains of a p50/RelA heterodimer has been solved (Chen, F.E. et al., 1998). The structure of this region of the rel homology domain displays no structural homology to the pKID structure observed in the NMR studies of CBP and CREB. The phosphorylation site resides in a loop between two antiparallel beta strands at the N-terminus of the last beta sheet in the dimerization domain. However, S276A mutations seriously debilitate transcription in cooperation with CBP in viva (Zhong, H. et al., 1998). Interestingly, Zhong et al. identified via sequence alignment a region of similarity between RelA and CREB immediately C-terminal to this phosphorylation site. This sequence includes a second, potential phosphorylation site (Thr 308). Perhaps phosphorylation of Ser 276 induces a conformational change exposing these C-terminal sequences to PKA, another kinase, or CBP itself. Phosphorylation of this Thr 308 residue might then allow interaction of this pKID-like sequence with the KIX domain of CBP. For this reason, mutagenesis of these C-terminal sequences, particularly Thr 308 would have been of interest. 155 These studies also produced evidence of the potential for cooperative interactions between multiple RelA domains and separate CBP domains. Zhong et al. assayed CBP coactivator activity using a cotransfection assay in viva. Cotransfection of a CBP expression vector enhances NF-KB transactivation. This enhancement correlates with the ability of CBP to bind RelA in vitro. The authors found that two RelA activation domain segments bind adjacent domains of the CBP adapter. A segment of RelAH binds the N- terminus of CBP, while the Ser 276 dependent segment of the rel homology domain binds the KD( domain of CBP immediately adjacent to this N-terminal segment. This modularity of both signaling partner (the activator) and signal transduction target (the CBP adapter) parallels the strategies employed by SH2 and SH3 domain containing signal transduction components. Whether these properties also result in cooperativity of activator-target protein-protein interactions remains to be addressed. These observations also suggest that the investigation of the effects of our RelA mutants in the natural context of the RelA activator may be worthwhile. Perhaps the RelAH region we have targeted is involved in CBP binding. The least understood aspects of these protein-protein interactions are perhaps the most fundamental. Traditionally, studies of protein-protein interactions in transcriptional regulation have been primarily phenomenological. Studies have been biased toward identification of potential targets. Although target recruitment through protein-protein interactions has long been suggested as the mode of action of transcriptional activators, the lack of identified targets has severely hampered our understanding of the nature of these binding events. The recent identification of the CBP/p300 family as coactivators of a large number of activators has allowed a more detailed study of these protein complexes. 156 More quantitative studies may allow a more detailed understanding of the molecular details of these events. More detailed quantitative studies of VP16-target interactions may also provide insight into VP16 structure and function. The VP16 activation domain binds multiple proteins of potential transcriptional relevance. Interactions with TBP (Stringer, K.F. et al., 1990), TFIIB (Roberts, S.G.E. et al., 1993), hTAFn32 (Klemrn, R.D. et al., 1995), TFHA (Kobayashi, N. et al., 1998), TFIIH (Xiao, H. et al., 1994) have all been reported. The activation domain may have evolved to produce a flexible interface capable of interaction with a variety of transcriptional targets. This flexibility would be advantageous to a viral protein, allowing strong activation within a genetically compact molecule. Alternatively, the complications of in vitro biochemistry using a protein that is both highly charged and highly hydrophobic may have contributed to the generation of artifacts. Typically, studies have utilized VP16 mutants to correlate in vitro and in viva phenotypes to minimize these concerns. Rarely have the energetics of these protein-protein interactions been investigated. Quantitation could provide a means to compare and contrast these interactions. The mechanism of DA complex formation may be another aspect that could benefit from more quantitative studies. The inability of VP16NAD/L or RelA to stabilize the DA complex suggests that the VP16C Fxx‘P‘I’ motif is not sufficient for this firnction. Additional residues within the subdomain must be necessary for this activity. A set of alanine substitutions throughout the entire VP16C subdomain has been constructed (Sullivan, SM. et al., 1998). Comparison of the effects of these mutations on DA complex formation may be informative. In addition, the component within the DA 157 complex that is recognized by VP16C remains unclear. Kobayashi, et al. established a correlation between VP16C mutant in viva potency and their capability to bind TFHA and stimulate DA complex formation. However, other components may be involved in this process, most notably TAFn32, which has been shown to bind VP16C (Klemm, RD. et al., 1995). Comparison of TAF binding, TFHA binding and DA complex formation with the alanine scan set of mutations could provide a detailed data set capable of resolving these issues. More quantitative studies using the alanine scan mutants may help identify the structural elements that specify DA complex stabilization. Perhaps sequences external to the FxxW motif of VP16C participate in TFIIA binding in a way similar to that observed in the SRC-l and GRIP] LxxLL motifs. 158 BIBLIOGRAPHY 159 BIBLIOGRAPHY Apone, L.M., C.-M.A. Virbasius, J .C. Reese and MR Green (1996). “Yeast TAFII90 is required for cell-cycle progression through G2/M but not for general transcription activation.” Genes and Dev. 10: 2368-23 80. Arany, Z., W.R. Sellers, D.M. Livingston and R. Eckner (1994). “ElA-associated p300 and CREB—associate CBP belong to a conserved family of coactivators.” Cgfl 77: 799- 800. Arias, J., A.S. Alberts, P. Brindle, F.X. Claret, T. Smeal, M. Karin, J. Feramisco and M. Montminy (1994). “Activation of CAMP and mitogen responsive genes relies on a common nuclear factor.” Nature 370: 226-229. Amdt, K.M., S.L. Ricupero, D.M. Eisenmann and F. Winston (1992). “Biochemical and genetic characterization of a yeast TFIID mutant that alters transcription in vivo and DNA binding in vitro.” Mol. Cell. Biol. 12: 2372-2382. Amdt, K.M., S. Ricupero-Hovasse and F. Winston (1995). “TBP mutants defective in activated transcription in viva.” EMBO J. 14(7): 1490-1497. Amdt, K.M., C.R. Wobbe, S. Ricupero-Hovasse, K. Struhl and F. Winston (1994). “Equivalent mutations in the two repeats of yeast TATA-binding protein confer distinct TATA recognition specificities.” Mol. Cell. Biol. 14(6): 3719-3728. Aso, T., J .W. Conaway and RC. Conaway (1995). “The RNA polymerase elongation complex.” FASEB J. 9: 1419-1428. Auble, D.T., Hansen, K. E., Mueller, C. G. F., Lane, W. S., Thomer, J., Hahn, S. (1994). “Motl, a global repressor of RNA polymerase II transcription, inhibits TBP binding to DNA by an ATP-dependent mechanism.” Genes and Development 8(16): 1920-1934. Bagby, 8., S]. Kim, E. Maldonado, K.I. Tong, D. Reinberg and M. Ikura (1995). “Solution structure of the C-temrinal core domain of human TFIIB: Similarity to cyclin A and interaction with TATA binding protein.” C_el_l 82: 857-867. Bannister, A.J., T. Oehler, D. Wilhelm, P. Angel and T. Kouzarides (1995). “Stimulation of c-Jun activity by CBP: c-Jun residues Ser63/73 are required for CBP induced stimulation of in viva and CBP binding in vitro.” Oncogene 11: 2509-2514. Bentley, D.L. (1995). “Regulation of transcriptional elongation by RNA polymerase II.” Curr. Opin. Genetics Dev. 5(2): 210-216. 160 Berger, S.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.” Cel_l 70: 25 1-265. Blair, W.S., H. Bogerd and RR. Cullen (1994a). “Genetic analysis indicates that the human foamy virus bel-l protein contains a transcriptional activation domain of the acidic class.” J Virol. 68: 3803-3808. Blair, W.S., H.P. Bogerd, S.J. Madore and BK Cullen (1994b). “Mutational analysis of the transcription activation domain of RelA: Identification of a highly synergistic minimal acidic activation module.” Mol. Cell. Biol. 14(11): 7226-7234. Blair, W.S. and ER. Cullen (1997). “A yeast TATA-binding protein mutant that selectively enhances gene expression from weak RNA polymerase II promoters.” Mol. Cell. Biol. 17: 2888-2896. Blair, W.S., RA. Fridell and ER. Cullen (1996). “Synergistic enhancement of both initiation and elongation by acidic transcription activation domains.” EMBO J. 15: 1658- 1665. Blau, J., H. Xiao, S. McCracken, P. O'Hare, J. Greenblatt and D. Bentley (1996). “Three functional classes of transcriptional activation domains.” Mol. Cell. Biol. 16: 2044-2055. Brown, S.A., C.S. Weirich, EM. Newton and RE. Kingston (1997). “Transcriptional activation domains stimulate initiation and elongation at different times and via different residues.” EMBO J 17: 3146-3154. Brownell, J .E., J. Zhou, T. Ranalli, R. Kobayashi, D.G. Edmondson, S.Y. Roth and CD. Allis (1996). “Tetrahymena histone acetyltransferase A: A homolog to yeast Gcn5p linking histone acetylation to gene activation.” C_efl 84: 843-851. Bryant, G.O., L.S. Martel, S.K. Burley and A.J. Berk (1996). “Radical mutations reveal TATA-box binding protein surfaces required for activated transcription in viva.” Genes Dev. 10: 2491-2504. Bunick, D., R. Zandomeni, S. Ackerman and R. Weinmann (1982). “Mechanism of RNA polymerase II-specific initiation of transcription in vitro: ATP requirement and uncapped run-off transcripts.” Cell 29: 877-886. Buratowski, S., S. Hahn, L. Guarente and PA. Sharp (1989). “Five intermediate complexes in transcription initiation by RNA polymerase H.” Cell 56: 549-561. Buratowski, S. and H. Zhou (1992). “Transcription factor IID mutants defective for interaction with transcription factor IIA.” Science 255: 1130-1132. 161 Burke, T.W. and J .T. Kadonaga (1996). “Drosophila TFIID binds to a conserved downstream basal promoter element that is present in many TATA-box—deficient promoters.” Genes Dev. 10: 711-724. Burley, SK. and RC. Roeder (1996). “Biochemistry and structural biology of transcription factor 11]) (TFIID).” Annu. Rev. Biochem. 65: 769-799. Cairns, B.R., Y]. Kim, M.H. Sayre, B.C. Laurent and RD. Kornberg (1994). “A multisubunit complex containing the SWIl/ADR6, SWIZ/SNFZ, SW13, SNFS, and SNF6 gene products isolated from yeast.” Proc. Natl. Acad. Sci. USA 91: 1950-1954. Candau, R. and S.L. Berger (1996). “Structural and functional analysis of yeast putative adaptors - evidence for an adaptor complex in viva.” J. Biol. Chem. 271: 5237-5245. Carcamo, J ., L. Buckbinder and D. Reinberg (1991). “The initiator directs the assembly of a transcription factor-IID-dependent transcription complex.” Proc. Natl. Acad. Sci. USA 88: 8052—8056. Carey, M., J. Leathervvood and M. Ptashne (1990). “A potent GAL4 derivative activates transcription at a distance in vitro.” Science 247 : 710-711. Carruth, L.M., Hardwick, J. M., Morse, B. A., Clements, J. E. (1994). “Visna virus tat protein: a potent transcription factor with both activator and suppressor domains.” Journal - of Virology 68(10): 6137-6146. Cavallini, B., I. Faus, 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 substitute for the human TATA box-binding factor.” Proc. Natl. Acad. Sci. USA 86: 9803—9807. Chang, C. and J .D. Gralla (1993). “Properties of initiator -associated transcription mediated by GAL4-VP16.” Mol. Cell. Biol. 13(12): 7469-7475. Chang, C., C.F. Kostrub and Z.F. Burton (1993). “RAP30/74 (transcription factor IIF) is required for promoter escape by RNA polymerase II.” J. Biol. Chem. 268: 20482-20489. Chang, J ., D.H. Kim, S.W. Lee, K.Y. Choi and Y.C. Sung (1995). “Transactivation ability of p53 transcriptional activation domain is directly related to the binding affinity to TATA- binding protein.” J. Biol. Chem. 270: 25014-25019. Chasman, D.I., J. Leatherwood, M. Carey, M. Ptashne and RD. Kornberg (1989). “Activation of yeast polymerase H transcription by herpesvirus VP16 and GAL4 derivatives in vitro.” Mol. Cell. Biol. 9: 4746-4749. 162 Chatterjee, and K. Struhl. (1995). “Connecting a promoter-bound protein to TBP bypasses the need for a transcriptional activation domain.” Nature 347: 820-822. Chen, F .E., D.-B. Huang, Y.-Q. Chen and G. Ghosh (1998). “Crystal structure of p50/p65 heterodimer of transcription factor NF-kappaB bound to DNA.” Nature 391: 410-413. Chen, J ., L.D. Attardi, C.P. Verrijzer, K. Yokomori and R. Tjian (1994). “Assembly of recombinant TFIID reveals differential coactivator requirements for distinct transcriptional activators.” QB 79(1): 93-105. Chi, T. and M. Carey (1996). “Assembly of the isomerized TFHA-TFIID-TATA ternary complex is necessary and sufficient fo gene activation.” Genes Dev. 10: 2540-2550. Chi, T.H., P. Lieberman, K. Ellwood and M. Carey (1995). “A general mechanism for transcriptional synergy by eukaryotic activators.” Nature 377: 254-257. Choi, B., M. Bando, S. Hasegawa and M. Horikoshi (1996). “Isolation and characterization of a cDNA encoding a novel human TFHD subunit containing similarities with histones H2B and H3.” Gene 169: 263-267. Choy, B. and MR. Green (1994). “Eukaryotic activators fimction during multiple steps of preinitiation complex assembly.” Nature 366: 531-536. Coleman, RA. and BF. 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 BF. Pugh (1995). “Dimerization of the TATA binding protein.” J. Biol. Chem. 270: 13842-13849. Colgan, J. and J .L. Manley (1992). “TFHD can be rate limiting in viva for TATA- containing, but not TATA-lacking, RNA polymerase II promoters.” Genes Devel. 6: 304- 315. Colgan, J ., S. Wampler and J .L. Manley (1993). “Interaction between a transcriptional activator and transcription factor-IIB in viva.” Nature 362(6420): 549-553. Conaway, RC. and J .W. Conaway (1993). “General initiation factors for RNA polymerase H.” Ann. Rev. Biochem. 62: 161-190. 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. 163 Cote, J ., J. Quinn, J .L. Workman and CL. Peterson (1994). “Stimulation of GAL4 derivative binding to nucleosomal DNA by the yeast SWI/SNF complex.” Science 265: 53-60. Cress, W.D. and SJ. Triezenberg (1991). “Critical structural elements of the VP16 transcriptional activation domain.” Science 251: 87-90. Crowley, T.E., T. Hoey, J .K. Liu, Y.N. Jan, L.Y. Jan and R. Tjian (1993). “A new factor related to TATA-binding protein has highly restricted expression patterns in Drosophila.” Nature 361: 557-561. Csordas, A. (1990). “On the biological role of histone acetylation.” Biochem J. 265: 23- 28. Dai, P., H. Akimaru, Y. Tanaka, D.X. Hou, T. Yasukawa, C. Kanei-Ishii, T. Takahashi and S. Ishii (1996). “CBP as a transcriptional coactivator of c-Myb.” Genes and Dev. 10: 528-540. Darimont, B.D., R.L. Wagner, J .W. Apriletti, M.R. Stallcup, P.J. Kushner, J .D. Baxter, R.J. Fletterick and KR. Yamamoto (1998). “Structure and specificity of nuclear receptor- coactivator interactions.” Genes Dev. 12: 3343-3356. De Felice, M., G. Damante, M. Zannini, H. Francis-Lang and R. Di Lauro (1995). “Redundant domains contribute to the transcriptional activity of the thyroid transcription factor 1.” J. Biol. Chem. 270: 26649-26656. Dikstein, R., S. Ruppert and R. Tjian (1996). “TAFIIZSO is a bipartite protein kinase that phosphorylates the basal transcription factor RAP74.” Cgfl 84: 7 8 1-790. Drapkin, R. and D. Reinberg (1994). “The multifunctional TFHH complex and transcriptional control.” Trends Biochem. Sci. 19: 504-508. Drysdale, C.M., E. Duenas, B.M. Jackson, U. Reusser, G.H. Braus and AG. Hinnebusch (1995). “The transcriptional activator GCN4 contains multiple activation domains that are dependent on hydrophobic amino acids.” MCB 15(3): 1220-1233. Dynlacht, B.D., T. Hoey and R. Tjian (1991). “Isolation of coactivators associated with the TATA-binding protein that mediate transcriptional activation.” Q11 66: 563-5 76. Eck, M.J., S. Pluskey, T. Trueb, S.C. Harrison and SE. Shoelson (1996). “Spatial constraints on the recognition of phosphoproteins by the tandem SH2 domains of the phosphatase SH-PTP2.” Nature 379: 277-280. Emami, K.H. and M. Carey (1992). “A synergistic increase in potency of a multimerized VP16 transcriptional activation domain.” EMBO J. 11(13): 5005-5012. 164 Emami, K.H., A. Jain and ST Smale (1997). “Mechanism of synergy between TATA and initiator: synergistic binding of TFHD following a putative TFHA-induced isomerization.” Genes and Dev. 11: 3007-3019. Emami, K.H., W.W. Navarre and ST. Smale (1995). “Core promoter specificities of the Spl and VP16 transcriptional activation domains.” Mol. Cell. Biol. 15: 5906-5916. 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(3): 1582-1593. Emili, A. and OJ. Ingles (1995). “Promoter-dependent photocross-linking of the acidic transcriptional activator E2F-1 to the TATA-binding protein.” J. Biol. Chem. 270(23): 13674-13680. F elsenfeld, G. (1992). “Chromatin as an essential part of the transcriptional mechanism.” Nature 355: 219-224. Ferreri, K., G. Gill and M. Montminy (1994). “The CAMP-regulated transcription factor CREB interacts with a component of the TFIID complex.” Proc. Natl. Acad. Sci. USA 91: 1210-1213. Fletcher, TM. and J .C. Hansen (1995). “Core histone tail domains mediate oligonucleosome folding and nucleosomal DNA organization through distinct mechanisms.” J. Biol. Chem. 43: 25359-25362. Folkers, G.E., E.C. Vanheerde and PT. Vandersaag (1995). “Activation function 1 of retinoic acid receptor beta 2 is an acidic activator resembling VP16.” J. Biol. Chem. 270: 23552-23559. Gasch, A., A. Hoffinann, M. Horikoshi, R.G. Roeder and N.-H. Chua (1990). “Arabidopsis thaliana contains two genes for TFIID.” Nature 346: 390-3 94. Ge, H. and RC. Roeder (1994). “Purification, cloning, and characterization of a human coactivator, PC4, that mediates transcriptional activation of class H genes.” C§l_l 78(3): 5 13-523 . Geiger, J .H., S. Hahn, S. Lee and PB. Sigler (1996). “Crystal structure of the yeast TFHA/TBP/DNA complex.” Science 272: 830-836. Gerster, T. and R. G. Roeder (198 8). “A herpesvirus trans-activating protein interacts with transcription factor OTF-l and other cellular proteins.” Proc. Natl. Acad. Sci. USA 85: 63 47 -63 5 1 . . 165 Gill, G., E. Pascal, Z.H. Tseng and R. Tjian (1994). “A glutamine-rich hydrophobic patch in transcription factor Spl contacts the dTAF H1 10 component of the Drosophila TFHD complex and mediates transcriptional activation.” Proc. Natl. Acad. Sci. USA 91: 192- 196. Gill, G. and M. Ptashne (1987). “Mutants of GAL4 protein altered in an activation fiinction.” Call 51: 121-126. Goodrich, J .A., T. Hoey, C.J. Thut, A. Admon and R. Tjian (1993). “Drosophila TAF(H)40 interacts with both a VP16 activation domain and the basal transcription factor TFIIB.” Lell 75: 519-530. Goodrich, J.A. and R. Tjian (1994). “Transcription factors IIE and IHI and ATP hydrolysis direct promoter clearance by RNA polymerase H.” C_ell 7 7: 145-156. 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 firnctions 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 ID and J.L. Workman (1998). “A subset of TAFHs are integral components of the SAGA complex rquired for nucleosome acetylation and transcriptional stimulation.” Ce_ll 94: 45-53. Gu, W. and R. G. Roeder (1997). “Activation of p53 sequence-specific DNA binding by acetylation of the p53 C-terrninal domain.” Ccfl 90: 595-606. Guarente, L., RR. Yocum and P. Gifford (1982). “A GALIO-CYCl hybrid yeast promoter identifies the GAL4 regulatory region as an upstream site.” Proc. Natl. Acad Sci. USA 79: 7410-7414. Hahn, S. (1993). “Structure(?) and function of acidic transcription activators.” Cgfl 72(4): 481-483. Hahn, S., S. Buratowski, P.A. Sharp and L. Guarente (1989). “Isolation of the gene encoding the yeast TATA binding protein TFIID: A gene identical to the SPT15 suppressor of Ty element insertions.” Ce_ll 58: 1173-1181. Hampsey, M. (1998). “Molecular genetics of the RNA polymerase H general transcription machinery.” Microbiology and Molecufirr Biolagy Reviews 62: 465-503. Hanna-Rose, W. and U. Hansen (1996). “Active Repression mechanisms of eukaryotic transcription repressors.” Trends Genet. 12: 229-234. 166 Hansen, S.K., S. Takada, RH. Jacobson, J .T. Lis and R. Tjian (1997). “Transcription properties of a cell type-specific TATA-binding protein.” Ce_ll 91: 71-83. Hansen, SK. and R. Tjian (1995). “TAFS and TFHA mediate differential utilization of the tandem Adh promoters.” Len 82: 565-575. Hatada, M.H., X. Lu, E.R. Laird, J. Green and J .P. Morgenstem (1995). “Molecular basis for the interaction of the protein tyrosine kinase ZAP-70 with the T-cell receptor.” Nature 377 : 32-38. Hayward, GS. (1993). “Immediate-early gene regulation in herpes simplex virus.” Sem. Virol. 4(1): 239-251. He, S. and SJ. Weintraub (1998). “Stepwise recruitment of components of the preinitiation complex by upstream activators in viva.” Mol. Cell. Biol. 18: 2876-2883. Hernandez, N. (1993). “TBP, a universal eukaryotic transcription factor?” Genes and Development 7(7b): 1291-1308. Hisatake, K., S. Hasegawa, R. Takada, Y. Nakatani, M. Horikoshi and RC. Roeder (1993). “The p250 subunit of native TATA box-binding factor-TFHD is the cell-cycle regulatory protein-CCGI .” Nature 362(6416): 179-181. Hisatake, K., T. Ohta, R. Takada, M. Guerrnah, M. Horikoshi, Y. Nakatani and RC. Roeder (1995). “Evolutionary conservation of human TAT A-binding-polypeptide- associated factors TAF(H)31 and TAF(H)80 and interactions of TAF(H)80 with other T AF 3 and with general transcription factors.” Proc. Natl. Acad. Sci. USA 92: 8195-8199. Hoey, T., B.D. Dynlacht, MD. Peterson, B.F. Pugh and R. Tjian (1990). “Isolation and characterization of the drosophila gene encoding the TATA box binding protein, TFIID.’ Ce_ll 61: 1179-1186. ‘) Hoey, T., R.O.J. Weinzierl, G. Gill, J .L. Chen, B.D. Dynlacht and R. Tjian (1993). “Molecular cloning and functional analysis of drosophila TAF l 10 reveal properties expected of coactivators.” Call 72(2): 247-260. Hofiinann, A., C.M. Chiang, T. Oelgeschlager, X.L. Xie, S.K. Burley, Y. Nakatani and R. G. Roeder (1996). “A histone octamer-like structure within TFIID.” Nature 380: 356- 359. Hoffmann, A., E. Sinn, T. Yamamoto, J. Wang, A. Roy, M. Horikoshi and RC. Roeder (1990). “Highly conserved core domain and unique N terminus with presumptive regulatory motifs in a human TATA factor (TFIID).” Nature 346: 387-3 90. 167 Hope, IA. and K. Struhl (1986). “Functional dissection of a eukaryotic transcriptional activator protein, GCN4 of yeast.” Lell 46: 885-894. Horikoshi, M., M.F. Carey, H. Kakidani and RC. Roeder (1988a). “Mechanism of action of a yeast activator: direct effect of GAL4 derivatives on mammalian TFIID-promoter interactions.” Ce_ll 54: 665-669. Horikoshi, M., T. Hai, Y.S. Lin, M.R. Green and RC. Roeder (1988b). “Transcription factor ATF interacts with the TATA factor to facilitate establishment of a preinitiation complex.” C_ell 54: 1033-1042. Horikoshi, M., C.K. Wang, H. Fujii, J .A. Cromlish, P.A. Weil and R.G. Roeder (1989). “Cloning and structure of a yeast gene encoding a general transcription initiation factor TFIID that binds to the TATA box.” Nature 341: 299-303. 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. Huang, M., N. Amheim and M.F. Goodman (1992). “Extension of base mispairs by Taq DNA polymerase: implications for single nucleotide discrimination in PCR.” Nucleic Acids M 20: 4567-4573. Imbalzano, A.N., H. Kwon, M.R. Green and RE. Kingston (1994a). “Facilitated binding of TATA-binding protein to nucleosomal DNA.” Nature 370: 481-485. Imbalzano, A.N., K.S. Zaret and RE. Kingston (1994b). “Transcription factor TFIIB and TFIIA can independently increase the affinity of the TATA-binding protein for DNA.” L Biol. Chem. 269(11): 8280-8286. Imhof, A., X. Yang, V.V. Ogryzko, Y. Nakatani, A.P. Wolffe and H. Ge (1997). “Acetylation of general transcription factors by histone acetyltransferases.” Current Biol. 7: 689-692. Ing, N.H., J .M. Beekman, S.Y. Tsai, M.J. Tsai and B.W. O'Malley (1992). “Members of the steroid hormone receptor superfamily interact with TFIIB (S300-II).” J. Biol. Chem. 267(25): 17617-17623. Ingles, C.J., M. Shales, W.D. Cress, S.J. Triezenberg and J. Greenblatt (1991). “Reduced binding of TFIID to transcriptionally compromised mutants of VP16.” Nature 351: 588- 590. Inostroza, J .A., F.H. Merrnelstein, 1. Ha, W.S. Lane and D. Reinberg (1992). “Drl, a TATA-binding protein-associated phosphoprotein and inhibitor of class H gene transcription.” C_ell 70: 477-489. 168 Inoue, M., M. Isomura, S. Ikegawa, T. Fujiwara, S. Shin, H. Moriya and Y. Nakamura (1996). “Isolation and characterization of a human cDNA clone (GCN5L1) homologous to GCN5, a yeast transcription activator.” Cytogenetics Cell Genetics 73: 134-136. 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. Jacq, X., C. Brou, Y. Lutz, I. Davidson, P. Chambon and L. Tora (1994). “Human TAFII3O is present in a distinct TFIID complex and is required for transcriptional activation by the estrogen receptor.” C_ell 7 9(1): 107-117. Janin, J. and C. Chothia (1990). “The structure of protein-protein recognition sites.” L Biol. Chem. 265: 16027-16030. J anin, J., S. Miller and C. Chothia (1988). “Surface, subunit interfaces and interior of oligomeric proteins.” J. Mol. Biol. 204: 155-164. J eong, S. and A. Stein (1994). “Micrococcal nuclease digestion of nuclei reveals extended nucleosome ladders having anomalous DNA lengths for chromatin assembled on non- replicating plasmids in transfected cells.” Nucleic Acids Res. 22: 370-3 75. Jones, S. and J .M. Thornton (1996). “Principles of protein-protein interactions.” Proc. lfirfl. Acad. Sci. USA 93: 13-20. Kamei, Y., L. Xu, T. Heinzel, J. Torchia, R. Kurokawa, B. Gloss, S.C. Lin, RA. Heyman, D.W. Rose, C.K. Glass and MG. Rosenfeld (1996). “A CBP integrator complex mediates transcriptional activation and AP-l inhibition by nuclear receptors.” Call 85: 403-414. Kao, C.C., P.M. Lieberman, M.C. Schmidt, Q. Zhou, R. Pei and A.J. Berk (1990). “Cloning of a transcriptionally active human TATA binding factor.” Science 248: 1646- 1650. Kaufrnann, J. and ST. Smale (1994). “Direct recognition of initiator elements by a component of the transcription factor HD complex.” Genes Dev. 8(7): 821-829. Kaufinann, J ., C.P. Verrijzer, J. Shao and ST Smale (1996). “CIF, an essential cofactor for TFIID-dependent initiator function.” Genes Dev. 10: 873-886. Kelleher HI, R.J., P.M. Flanagan and RD. Kornberg (1990). “A novel mediator between activator proteins and the RNA polymerase H transcription apparatus.” Call 61: 1209- 121 5. 169 Kim, T.K., S. Hashimoto, RJ. Kelleher, HI, P.M. Flanagan, RD. Kornberg, M. Horikoshi and R. G. Roeder (1994). “Effects of activation-defective TBP mutations on transcription initiation in yeast.” Nature 369: 252-255. Kim, Y.C., J.H. Geiger, S. Hahn and PB. Sigler (1993). “Crystal structure of a yeast TBP/TATA-Box complex.” Nature 365: 512-520. Kim, Y.-J., S. Bjorklund, Y. Li, M.H. Sayre and RD. Kornberg (1994). “A multiprotein mediator of transcriptional activation and its interaction with the c-terrninal repeat domain of RNA polymerase H.” C_ell 77(4): 599-608. Kingston, R.E., C.A. Bunker and AN. Irnbalzano (1996). “Repression and activation by multiprotein complexes that alter chromatin structure.” Genes Dev. 10: 905-920. Klages, N.a.S., M. (1995). “Stimulation of RNA polymerase H transcription initiation by recruitment of TBP in viva.” Nature 374: 822-823. Klein, C., Struhl, K. (1994). “Increased recruitment of TATA-binding protein to the promoter by transcriptional activation domains in viva.” Science 266: 280-282. Klemm, R, J. Goodrich, S. Zhou and R. Tjian (1995). “Molecular cloning and expression of the 32-kDa subunit of human TFIID reveals interactions with VP16 and TFIIB that mediate transcriptional activation.” Proc. Natl. Acad. Sci. USA. 92(13): 5788-5792. Klemm, R.D., J .A. Goodrich, S.L. Zhou and R. Tjian (1995). “Molecular cloning and expression of the 32-kDa subunit of human TFHD reveals interactions with VP16 and TFHB that mediate transcriptional activation.” Proc. Natl. Acafd. Sci. USA. 92(13): 5788-5792. Kobayashi, N., T.G. Boyer and A.J. Berk (1995). “A class of activation domains interacts directly with TFHA and stimulates TFHA-TFIID-promoter complex assembly.” Mol. Cell. Biol. 15: 6465-6473. Kobayashi, N., P]. 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. Kokubo, T., D.W. Gong, S. Yamashita, M. Horikoshi, RG. Roeder and Y. Nakatani (1993a). “Drosophila 230-kD TFIID subunit, a functional homolog of the human cell cycle gene product, negatively regulates DNA binding of the TATA box-binding subunit ofTFIID.” _Ge_nesm 7(6): 1033-1046. Kokubo, T., D.-W. Gong, J .C. Wooton, M. Horikoshi, RG. Roeder and Y. Nakatani (1994). “Molecular cloning of the Drosophila TFHD subunits.” Nature 367(3): 484-487. 170 Kokubo, T., S. Yamashita, M. Horikoshi, RG. Roeder and Y. Nakatani (1993b). “Interaction between the N-terminal domain of the 23 O-kDa subunit and the TATA-box binding subunit of TFIID negatively regulates TATA-box binding.” Proc. Natl. Acad. Sci. LISA 91: 3520-3524. Koleske, A.J. and RA. Young (1994). “An RNA polymerase H holoenzyme responsive to activators.” Nature 368: 466-469. Koleske, A.J. and RA. Young (1995). “The RNA polymerase H holoenzyme and its implications for gene regulation.” Trends Biochem. Sci. 20: 113-116. Kretzschmar, M., K. Kaiser, F. Lottspeich and M. Meisteremst (1994). “A novel mediator of class H gene transcription with homology to viral immediate-early transcriptional regulators.” Ce_ll 78: 525-534. Krumm, A, L.B. Hickey and M. Groudine (1995). “Protomer—proximal pausing of RNA polymerase H defines a general rate-limiting step after transcriptional initiation.” Genes and Dev. 9: 559-572. Krumm, A., T. Meulia, M. Brunvand and M. Groudine (1992). “The block to transcriptional elongation within the human c-myc gene is determined in the promoter - proximal region.” Genes and Development 6(11): 2201-2213. Kumar, K.P., S. Akouliychev and D. Reinberg (1998). “Promoter-proximal stalling results from the inability to recruit transcription factor III-I to the transcription complex and is a regulated event.” Proc. Natl. Acad. Sci. USA 95: 9767-9772. Kunkel, TA. (1985). “Rapid and efiicient site-specific mutagenesis without phenotypic selection.” Proc. Natl. Acad. Sci. USA 82: 488-492. Kunzler, M., G.H. Braus, O. Georgiev, K. Seipel and W. Schaffirer (1994). “Functional differences beteen mammalian transcription activation domains at the yeast Gall promoter.” EMBO J. 13(3): 641-645. Kuriyan, J. and D. Cowbum (1997). “Modular peptide recognition domains in eukaryotic signaling.” Annu. Rev. Biophys. Biomol. Struct. 26: 259-288. Kussie, P.H., S. Gorina, V. Marechal, B. Elenbaas, J. Moreau, A.J. Levine and NP. Pavletich (1996). “Structure of the MDM2 oncoprotein bound to the p53 tumor suppressor transactivation domain.” Science 274: 948-953. Kwok, RP.S., J .R. Lundblad, J.C. Chrivia, J .P. Richards, H.P. Bachinger, RG. Brennan, S.G.E. Roberts, M.R Green and RH. Goodman (1994). “Nuclear protein CBP is a coactivator for the transcription factor CREB.” Nature 370: 223 -226. 171 Lagrange, T., A.N. Kapanidis, H. Tang, D. Reinberg and RH. 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. Layboum, RJ. and ME. Dahmus (1990). “Phosphorylation of RNA polymerase IIA occurs subsequent to interaction with the promoter and before the initiation of transcription.” J. Biol. Chem. 265: 13165-13173. Lee, D.Y., J .J . Hayes, D. Pruss and AP. Wolffe (1993). “A positive role for histone acetylation in transcription factor access to nucleosomal DNA.” Call 72(1): 73-84. 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 viva.” Mol. Cell. Biol. 15: 5461-5469. Lee, M. and K. Struhl (1997). “A severely defective TATA-binding protein-TFHB interaction does not preclude transcriptional activation in viva.” Mol. Cell. Biol. 17: 1336- 1345. Lee, TI. and RA. Young (1998). “Regulation of gene expression by TBP-associated proteins.” Genes Dev. 12: 1398-1408. Lemesle-Varloot, L., B. Henrissat, C. Gaboriaud, V. Bissery, A. Morgat and J .-P. Mornon (1990). “Hydrophobic cluster analysis: procedures to derive structural and functional information fi'om 2-D-representation of protein sequences.” Biochimie 72: 555-574. Leung, D.W., B. Chen and D.V. Goeddel (1989). “A method for random mutagenesis of a defined DNA segment using a modified polymerase chain reaction.” Technique 1: 11-15. Leuther, K.K., J.M. Salmeron and SA. Johnston (1993). “Genetic evidence that an activation domain of GAL4 does not require acidity and may form a beta sheet.” Cell 72(4): 575-585. Lieberman, P. (1994). “Identification of functional targets of the Zta transcriptional activator by formation of stable preinition complex intermediates.” Mol. Cell. Biol. 14(12): 8365-8375. Lieberman, RM. and A.J. Berk (1991). “The Zta trans-activator protein stabilizes TFIID association with promoter DNA by direct protein-protein interaction.” Genes Dev. 5: 2441-2454. Lieberman, PM. and A.J. Berk (1994). “A mechanism for TAFS in transcriptional activation: activation domain enhancement of TFIID-TFHA-promoter DNA complex formation.” Genes Dev. 8: 995-1006. 172 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, binding to mdm-2 and the adenovirus 5 ElB 55-kD protein.” Genes Dev. 8(10): 1235-1246. Lin, Y.S., M.F. Carey, M. Ptashne and MR. Green (1988). “GAL4 derivatives fimction alone and synergistically with mammalian activators in vitro.” C_ell 54: 659-664. Lin, Y.S. and MR. Green (1991). “Mechanism of action of an acidic transcriptional activator in vitro.” C_ell 64: 971-981. Lin, Y.S., 1. Ha, E. Maldonado, D. Reinberg and MR. Green (1991). “Binding of general transcription factor TFIIB to an acidic activating region.” Nature 353: 569-5 71. Liu, D.J., R. Ishima, K.I. Tong, S. Bagby, T. Kokubo, D.R. Muhandiram, L.E. Kay, Y. Nakatani and M. Ikura (1998). “Solution structure of a TBP-TAF(H)230 complex: Protein mimicry of the minor groove surface of the TATA box unwound by TBP.” Call 94: 573-583. Logie, C. and CL. Peterson (1997). “Catalytic activity of the yeast SWI/SNF complex on reconstituted nucleosome arrays.” EMBO J 16(22): 6772-6782. Logie, CL, C. Tse, J .C. Hansen and CL. Peterson (1998). “The core histone N-terrninal domains are required for catalytic chromatin remodeling by the SWI/SNF and RSC complexes.” : submitted. Loidl, P. (1994). “Histone acetylation. Facts and questions.” Chromosoma 103: 441-449. Lu, H. and A.J. Levine (1995). “Human TAF(II) 31 protein is a transcriptional coactivator of the p53 protein.” Proc. Neil. Acad. Sci. USA 92(11): 5154-5158. Luger, K., Maeder, A.W., Richmond, R.K., Sargent, DE, and Richmond, TJ. (1997). “Crystal structure of the nucleosome core particle at 2.8 angstrom resolution.” Nature 389: 251-260. Ma, D., H. Watanbe, F . Merrnelstein, A. Admon, K. Oguri, X. Sun, T. Wada, T. Imai, S. Toshifiimi, D. Reinbeg and H. Handa (1993). “Isolation of a cDNA encoding the largest subunit of TFHA reveals functions important for activated transcription.” Genes Devel 7: 2246-2257. Ma, J. and M. Ptashne (1987). “Deletion analysis of GAL4 defines two transcriptional activating segments.” C_ell 48: 847-853. Maignan, S., J .-P. Guilloteau, N. Fromage, B. Amoux, J. Becquart and A. Decruix (1995). “Crystal structure of the mammalian GRB2 adaptor.” Science 268: 291-293. 173 Maldonado, E., 1. Ha, P. Cortes, L. Weis and D. Reinberg (1990). “Factors involved in specific transcription by mammalian RNA polymerase H: role of transcription factors HA, HD, and IIB during formation of a transcription-competent complex.” Mol. Cell. Biol. 10: 633 5-6347. Maldonado, E., R. Shiekhattar, M. Sheldon, H. Cho, R. Drapkin, P. Rickert, E. Lees, A.C. W., S. Linn and D. Reinberg (1996). “A human RNA polymerase H complex associated with SRB and DNA-repair proteins.” Nature 381: 86-89. 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: 3 197-3 205. 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. Martinez, E., C. Chiang, H. Ge and RG. Roeder (1994). “TATA-binding protein- associated factor(s) in TFIID function through the initiator to direct basal transcription from a TATA-less class H promoter.” EMBO J 13: 3115-3126. Martinez, E., H. Ge, Y. Tao, C. Yuan, V. Palhan and R. g. Roeder (1998a). “Novel cofactors and TFHA mediate functional core promoter selectivity by the human TAFH150-containing TFIID complex.” Mol. Cell. Biol. 18: 6571-6583. Martinez, E., T.K. Kundu, J. Fu and RG. Roeder (1998b). “A human SPT3-TAF(H)31- GCNS-L acetylase complex distinct from transcription factor IID.” J. Biol. Chem. 273: 23781-23785. McInemey, B.M., D.W. Rose, S.E. Flynn, S. Westin, T. Mullen, A. Krones, J. Inostroza, J. Torchia, R. Nolte, N. Assa-Munt, M.V. Milbum, C.K. Glass and MG. Rosenfeld (1998). “Determinants of coactivator LXXLL motif specificity in nuclear receptor transcriptional activation.” Genes and Dev. 12: 3357-3368. 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. Komarnitsky and S. Buratowski (1998). “Histone-like TAFS are essential for transcription in viva.” Mol. Cell 2: 663-673. Miller, J .H. (1972). Experiments in Molecular Genetic; Cold Spring Harbor, New York. Mitchell, RJ. and R. Tjian (1989). “Transcriptional regulation in mammalian cells by sequence-specific DNA binding proteins.” Science 245: 371-378. 174 Mizzen, C.A., X. Yang, T. Kokubo, J .E. Brownell, A.J. Bannister, T. Owen-Hughes, J. Workman, L. Wang, S.L. Berger, T. Kouzarides, Y. Nakatani and CD. Allis (1996). “The TAFHZSO subunit of TFIID has histone acetyltransferase activity.” Cell 87: 1261-1270. Moore, P.A., S.M. Ruben and CA. Rosen (1993). “Conservation of transcriptional activation functions of the NF-kB p50 and p65 subunits in mammalian cells and Saccharomyces cerevisiae.” Mol. Cell. Biol. 13: 1666-1674. 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. Muhich, M.L., C.T. Iida, M. Horikoshi, RG. Roeder and CS. Parker (1990). “cDNA clone encoding drosophila transcription factor-TFIID.” Proc. hilt]. Acad. Sci. USA 87: 9148-9152. Naar, A.M., P.A. Beaurang, K.M. Robinson, J .D. Oliner, D. Avizonis, S. Scheek, J. Zwicker, J .T. Kadonaga and R Tjian (1998). “Chromatin, TAFS, and a novel multiprotein coactivator are required for synergistic activation by Spl and SREBP-la in vitro.” Genes and Dev. 12: 3020-3031. Nakajima, T., C. Uchida, S.F. Anderson, J .D. Parvin and M. Montminy (1997). “Analysis of a CAMP-responsive activator reveals a two-component mechanism for transcriptional induction via signal-dependent factors.” Genes and Dev. 11: 73 8-747. Nielsen, D.A., T. Chang and DJ. Shapiro (1989). “A highly sensitive mixed phase assay for chloramphenicol acetyltransferase activity in transfected cells.” Anal. Biochem. 179: 12-23. Nikolov, DB. and SK. Burley (1997). “RNA polymerase H transcription initiation: a structural view.” Proc. Natl. Acad. Sci. 94: 15-22. Nikolov, D.B., H. Chen, E.D. Halay, A. Hoffrnann, RG. Roeder and SK. Burley (1996). “Crystal structure of a human TATA box-binding protein/TATA element complex.” Proc. Nat. Acad. Sci. USA 93: 4862-4867. Nikolov, D.B., H. Chen, E.D. Halay, A.A. Usheva, K. Hisatake, D.K. Lee, RG. Roeder and SK. Burley (1995). “Crystal structure of a TFIIB-TBP-TATA-element ternary complex.” Nature 377 : 119-128. Nikolov, D.B., S. Hu, J. Lin, A. Gasch, A. Hoffrnann, M. Horikoshi, N. Chua, RG. Roeder and SK. Burley (1992). “Crystal structure of TFHD TATA-box binding protein.” Nature 360: 40-46. 175 Nishikawa, J ., T. Kokubo, M. Horikoshi, RG. Roeder and Y. Nakatani (1997). “Drosophila TAF(H)230 and the transcriptional activator VP16 bind competitively to the TATA box-binding domain of the TATA box-binding protien.” Proc. Natl. Acad. Sci. 94: 85-90. Nolte, R.T., G.B. Wisely, S. Westin, J .E. Cobbs, 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: 137-143. O'Brien, T. and J .T. Lis (1991). “RNA polymerase H pauses at the 5'end of the transcriptionally induced Drosophila hsp70 gene.” Mol. Cell. Biol. 11: 5285-5290. Oelgeschlager, T., C.M. Chiang and RG. Roeder (1996). “Topology and reorganization of a human TFIID-promoter complex.” Nature 382: 73 5-738. Ogryzko, V.V., R.L. Schiltz, V. Russanova, B.H. Howard and Y. Nakatani (1996). “The transcriptional coactivators p300 and CBP are histone acetyltransferases.” Ce_ll 87: 953- 959. Orphanides, G., T. Lagrange and D. Reinberg (1996). “The general transcription factors of RNA polymerase H.” Genes Dev. 10: 2657-2683. Oshea-Greenfield, A. and ST. Smale (1992). “Roles of TATA and initiator elements in determining the start site location and direction of RNA polymerase H transcription.” J_. Biol. Chem. 267: 1391-1402. Ozer, J ., A.H. Bolden and RM. Lieberman (1996). “Transcription factor HA mutations show activator-specific defects and reveal a HA function distinct fi'om stimulation of TBA-DNA binding.” J. Biol. Chem. 271: 11182-1 1 190. Ozer, J ., K. Mitsouras, D. Zerby, M. Carey and RM. Lieberman (1998). “Transcription factor HA derepresses TATA-binding protein (TBP)-associated factor inhibition of TBP- DNA binding.” J. Biol. Chem. 273: 14293-14300. Ozer, J ., P.A. Moore, A.H. Bolden, A. Lee, C.A. Rosen and PM. Lieberman (1994). “Molecular cloning of the small (gamma) subunit of human TFHA reveals fimctions critical for activated transcription.” Genes Dev. 8(19): 2324-2335. Paal, K., P.A. Baeuerle and ML. Schmitz (1997). “Basal transcription factors TBP and TFIIB and the viral coactivator ElA 13S bind with distinct affinities and kinetics to the transactivation domain of NF-kappa B p65.” Nucleic Acids Res. 25: 1050-1055. 176 Payne, J .M., R]. Laybourn and ME. Dahmus (1989). “The transition of RNA polymerase H from initiation to elongation is associated with phosphorylation of the carboxyl-terminal domain of subunit Ha.” J. Biol. Chem. 264: 19621-19629. Peterson, CL, A. Dingwall and MP. Scott (1994). “Five SWI/SNF gene products are components of a large multisubunit complex required for transcriptional enhancement.” Proc. Neal. Acad. Sci. USA 91: 2905-2908. Peterson, CL. and I. Herskowitz (1992). “Characterization of the yeast SW11, SW12 and SW13 genes, which encode a global activator of transcription.” C_ell 68: 573-583. Peterson, M.G., N. Tanese, B.F. Pugh and R. Tjian (1990). “Functional domains and upstream activation properties of cloned human TATA binding protein.” Science 248: 1625-1630. 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(10): 5981-5989. Pinto, 1., DE. Ware and M. Hampsey (1992). “The yeast SUA7 gene encodes a homolog of human transcription factor TFIIB and is required for normal start site selection in vivo.” C_efl 68: 977-988. Poon, D., Y. Bai, A.M. Campbell, S. Bjorklund, Y]. Kim, S. Zhou, RD. Kornberg and RA. Weil (1995). “Identification and characterization of a TFHD-like multiprotein complex from Saccharomyces cerevisiae.” Proc. Natl. Acad. Sci. USA 92: 8224-8228. 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.” C_ell 52: 425-434. Ptashne, M. (1988). “How eukaryotic transcriptional activators work.” Nature 335: 683- 689. Ptashne, M. and A. Gann (1997). “Transcriptional activation by recruitment.” Nature 386: 569-577. Pugh, BF. and R. Tjian (1990). “Mechanism of transcriptional activation by Spl: evidence for coactivators.” Q11 61: 1187-1197. Radhakrishnan, I., GO Perez-Alvarado, D. Parker, H.J. Dyson, M.R. Montminy and PE. Wright (1997). “Solution structure of the KIX domain of CBP bound to the transactivation domain of CREB: A model for activatorzcoactivator interactions.” C_efl 91: 741-752. 177 Ramboarina, S., N. Morellet, P. Petitjean, B.P. Roques and MC. Fournie-Zaluski (1998). “Model of the complex formed between the H2 domain of the TATA box binidng protein and the L(281-301) fragment of the human transcription factor TFIIA.” Protein Engng 11: 729-738. Ranish, J .A., W.S. Lane and S. Hahn (1992). “Isolation of two genes that encode subunits of the yeast transcription factor IIA.” Science 255: 1127-1129. Reddy, P. and S. Hahn (1991). “Dominant negative mutations in yeast TFIID define a bipartite DNA binding region.” Q 65: 349-357. Reese, J .C., Apone, L., Walker, S. S., Griffin, L. A., Green, M. R. (1994). “Yeast TAF(II)s in a multisubunit complex required for activated transcription .” Nature 371(6497): 523-527. Regier, J .L., F. Shen and SJ. 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. Reinberg, D., M. Horikoshi and RG. Roeder (1987). “Factors involved in specific transcription by mammalian RNA polymerase 11. Functional analysis of initiation factors HA and HD and identification of a new factor operating at sequences downstream of the initiation site.” J. Biol. Chem. 262: 3322-3330. Roberts, S.G.E., 1. Ha, E. Maldonado, D. Reinberg and MR. Green (1993). “Interaction between an acidic activator and transcription factor IIB is required for transcriptional activation.” Nature 363(6431): 741-744. Roberts, SM. and F. Winston (1996). “SPT20/ADA5 encodes a novel protein firnctionally related to the TATA-binding protein and important for the transcription in Saccharomyces cerevisiae.” Mol. Cell. Biol. 16: 3206-3213. Roeder, RG. ( 1991). “The complexities of eukaryotic transcription initiation: regulation of preinitiation complex assembly.” TIBS 16: 402-407. Roeder, RG. (1996). “The role of general initiation factors in transcription by RNA polymerase 11.” Trends Biochem. Sci. 21: 327-335. Roy, A.L., S. Malik, M. Meisteremst and RG. Roeder (1993). “An alternative pathway for transcription initiation involving TFII-I.” Nature 365(6444): 355-3 59. Ruden, D.M., J. Ma, Y. Li, K. Wood and M. Ptashne (1991). “Generating yeast transcriptional activators containing no yeast protein sequences.” Nature 21: 250-252. 178 Ruppert, S., B.H. Wang and R Tjian (1993). “Cloning and expression of human TAF(H)250 - a TBP-associated factor implicated in cell-cycle regulation.” Nature 362(6416): 175-179. Sadowski, I., J. Ma, S. Triezenberg and M. Ptashne (1988). “GAL4-VP16 is an unusually potent transcriptional activator.” Nature 335: 563-564. Sakaguchi, K., J.E. Herrera, S. Saito, T. Miki, M. Bustin, A. Vassilev, C.W. Anderson and E. Appella (1998). “DNA damage activates p53 through a phosphorylation- acetylation cascade.” Genes and Dev. 12: 2831-2841. Sauer, F., S.K. Hansen and R. Tjian (1995). “Multiple TAFIIs direction synergistic activation of transcription.” Science 270: 1783-1788. Sauer, F., D.A. Wassarrnan, G.M. Rubin and R. Tjian (1996). “TAFIIs mediate activation of transcription in the Drosophila embryo.” Cell 87 : 1271-1284. Sawadogo, M. and R. G. Roeder (1985). “Interaction of a gene-specific transcription factor with the adenovirus major late promoter upstream of the TATA box region.” Cefl 43: 165-175. Schmidt, M.C., C.,C. Kao, R. Pei and A.J. Berk (1989). “Yeast TATA-box transcription factor gene.” Proc. Md. Acad. Sci. USA 86: 7785-7789. Schmitz, M., M. Silva and P. Baeuerle (1995). “Transactivation domain 2 (TA(2)) of p65 NF-kappa B-Sirnilarity to TA(1) and phorbol ester-stimulated activity and phosphorylation in intact cells.” J. of Bio. Chemistg 270(26): 15576-15584. Schmitz, M.L., M.A. dos Santos Silva, H. Altmann, M. Czisch, T.A. Holak and RA. Baeuerle (1994). “Structural and firnctional 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. Seipel, K., O. Geogriev and W. Schaffner (1994). “A minimal transcription activation domain consisting of a specific array of aspartic acid and leucine residues.” Biol. Chem. Hoppe—Seyler 375(7): 463-470. Seipel, K., O. Georgiev and W. Schaffner (1992). “Different activation domains stimulate transcription from remote (enhancer) and proximal (promoter) positions.” EMBO J. 11(13): 4961-4968. Sekiguchi, T., T. Miyata and T. Nishimoto (1988). “Molecular cloning of the cDNA of a human X chromosomal gene (CCGI) which complements the temperature-sensitive G1 mutants, tsBN462 and ts13, of the BHK cell line.” EMBO J 7: 1683-1687. 179 Sekiguchi, T., Y. Nohiro, Y. Nalamura, N. Hisamoto and T. Nishimoto (1991). “The human CCGI gene, essential for progression of the G1 phase, encodes a 210 kilodalton nuclear DNA-binding protein.” Mol. Cell. Biol. 11: 3317-3325. Seto, E., A. Usheva, G.P. Zambetti, J. Momand, N. Horikoshi, R. Weinmann, A.J. Levine and T. Shenk (1992). “Wild-type p53 binds to the TATA-binding protein and represses transcription.” Procflatl. Acad. Sci. USA 89(24): 12028-12032. Siebenlist, U., G. Franzoso and K. Brown (1994). “Structure, regulation and fiinction of NF-kB.” Annp. Rev. Cell Biol. 10: 405-455. Sigler, P. (1988). “Acid blobs and negative noodles.” Nature 333: 210-212. Smale, ST (1997). “Transcription initiation from TATA-less promoters within eukaryotic protein-coding genes.” Biochim. et Biophys. Acta 1351: 73-88. Smith, CL. and G.L. Hager (1997). “Transcriptional regulation of mammalian genes in viva.” J. Biol. Chem. 272: 27493-27496. Spencer, CA. and M. Groudine (1990). “Transcription elongation and eukaryotic gene regulation.” Oncogene 5: 777-786. Stargell, LA. and K. Struhl (1995). “The TBP-TFHA interaction in the response to acidic activators in viva.” Science 269: 75-78. Stern, M., Jensen, R, Herskowitz, I. (1984). “Five SWI genes are required for expression of the H0 gene in yeast.” J Mol. Biol. 178(4): 853-868. Stoscheck, CM. (1990). “Quantitation of protein.” Methods Enzymol 182: 50-68. 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. Struhl, K. (1998). “Histone acetylation and transcriptional regulatory mechanisms.” Genes and Dev. 12: 599-606. Sullivan, S.M., P.J. Horn, V.A. Olson, A.H. Koop, W. Nu, RH. Ebright and SJ. Triezenberg (1998). “Mutational analysis of a transcriptional activation region of the VP16 protein of herpes simplex virus.” Nucleic Acids Res.: submitted. Sun, X., D. Ma, M. Sheldon, K. Yeung and D. Reinberg (1994). “Reconstitution of human TFHA activity fi'om recombinant polypeptides: a role in TFHA-mediated transcription.” Genes Dev 8(19): 2336-2348. 180 Taggart, A.K.P. and BF. Pugh (1996). “Dimerization of TFHD when not bound to DNA.” Science 272: 1331-1333. Tan, S., Y. Hunziker, D.F. Sargent and TI. Richmond (1996). “Crystal structure of a yeast TFHA/TBP/DNA complex.” Nature 381: 127-134. Tanaka, M. and W. Herr (1994). “Reconstitution of transcriptional activation domains by reiteration of short peptide segments reveals the modular organization of a glutamine-rich activation domain.” Mol. Cell. Biol. 14(9): 6056-6067. Tang, H., X. Sun, D. Reinberg and RB. Ebright (1996). “Protein-protein interactions in eukaryotic transcription initiation: structure of the preinitiation complex.” Proc. Natl. Acad. Sci. USA 93: 1119-1124. Tansey, WP. and W. Herr (1997a). “Selective use of TBP and TFIIB revealed by a TATA-TBP-TFIIB array with altered specificity.” Science 7: 829-831. Tansey, WP. and W. Herr (1997b). “TAFS; Guilt by association.” C_ell 88: 729-732. Tasset, D., L. Tora, C. Fromental, E. Scheer and P. Chambon (1990). “Distinct classes of transcriptional activating domains function by different mechanisms.” Cgfl 62: 1177-1187. Thompson, C.M., A.J. Koleske, D.M. Chao and RA. Young (1993). “A multisubunit complex associated with the RNA polymerase H CTD and TATA-Binding protein in yeast.” Call 73: 1361-1375. Thut, C.J., J. Chen, R. Klemm and R. Tjian (1995). “p53 transcriptional activation mediated by coactivators TAF1140 and TAFII60.” Science 267: 100-104. Triezenberg, SJ. (1995). “Structure and fiinction of transcriptional activation domains.” Curr. Opin. Genetics Dev. 5: 190-196. Triezenberg, S.J., RC. 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. Turner, BM. (1991). “Histone acetylation and and control of gene expression.” J. Cell _S_cl 99: 13-20. Usheva, A. and T. Shenk (1994). “TATA-binding protein-independent initiation: YYI, TFHB, and RNA polymerase H direct basal transcription on supercoiled template DNA.” C_ell 76: 1115-1121. 181 Utley, RT., K. Ikeda, P.A. Grant, J. Cote, D]. Steger, A. Eberharter, S. John and J .L. Workman (1998). “Transcriptional activators direct histone acetyltransferase complexes to nucleosomes.” Nature 394: 498-502. 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.” Ce_ll 81(7): 1115-1125. Verrijzer, C.P., K. Yokomori, J .L. Chen and R. Tjian (1994). “Drosophila TAFH 150: similarity to yeast gene TSM-l and specific binding to core promoter DNA.” Science 264: 933-941. 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 TFIIB binding.” Mol. Cell. Biol. 13(9): 5233-5244. Walker, 8.8., J .C. Reese, L.M. Apone and MR. Green (1996). “Transcriptional activation in cells lacking TAF(H)s.” Nature 383: 185-188. 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. Weinheirner, S.P., B.A. Boyd, S.K. Durham, J .L. Resnick and DR. O'Boyle (1992). “Deletion of the VP16 open reading frame of herpes simplex virus type 1.” J. Virol. 66: 25 8-269. Weinzierl, R.O.J., S. Ruppert, B.D. Dynlacht, N. Tanese and R. Tjian (1993). “Cloning and expression of Drosophila TAF 1170 reveal conserved interactions with other subunits ofTFIID.” EMBO J 12: 5303-5309. White, J ., C. Brou, J. Wu, Y. Lutz, V. Moncollin and P. Chambon (1992). “The acidic transcriptional activator GAL-VP16 acts on preformed template-committed complexes.” EMBO J. 11: 2229-2240. Winston, F. and M. Carlson (1992). “Yeast SNF/SWI transcriptional activtors and the SPT/SIN chromatin connection.” Trends Genet. 8(11): 387-391. Wolffe, A. (1994). “Nucleosome positioning and modification: chromatin structures that potentiate transcription.” Trends Biol. Sci. 19: 240-244. Workman, J.L. and RE. Kingston (1998). “Alteration of nucleosome structure as a mechanism of transcriptional regulation.” Annp. Rev. Biochem. 67: 545-579. Wu, 8., E. Kershnar and C. Chiang (1998). “TAFH-independent activation mediated by human TBP in the presence of the positive cofactor PC4.” EMBO J 17: 4478-4490. 182 Wu, Y., RJ. Reece and M. Ptashne (1996). “Quantitation of putative activator-target affinities predicts transcriptional activating potentials.” EMBO J. 15: 3951-3963. 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(10): 7013-7024. Xie, X.L., T. Kokubo, S.L. Cohen, U.A. Mirza, A. Hoffrnann, B.T. Chait, RG. Roeder, Y. Nakatani and SK. Burley (1996). “Structural similarity between TAFS and the heterotetrameric core of the histone octamer.” Nature 380: 316-322. Xu, X., C. Prorock, H. Ishikawa, E. Maldonado, Y. Ito and C. Gelinas (1993). “Functional interaction of the v-Rel and c-Rel oncoproteins with the TATA-binding protein and association with transcription factor-HB.” Mol. Cell. Biol. 13(11): 6733-6741. Yamamoto, T., M. Horikoshi, J. Wang, S. Hasegawa, P.A. Weil and RG. Roeder (1992). “A bipartite DNA binding domain composed of direct repeats in the TATA box binding factor TFHD.” Proc. Natl. Acad. Sci. USA 89: 2844-2848. Yang, X.-J., V.V. Ogryzko, J.-I. Nishikawa, B.H. Howard and Y. Nakatani (1996). “A p3 00/CBP-associated factor that competes with the adenoviral oncoprotein ElA.” Nature 382: 319-324. - Yankulov, K., J. Blau, T. Purton, S. Roberts and D.L. Bentley (1994). “Transcriptional elongation by RNA polymerase II is stimulated by transactivators.” Ce_ll 77(5): 749-759. Yeung, K.C., Inostroza, J. A., Merrnelstein, F. H., Kannabiran, C., Reinberg, D. (1994). “Structure-firnction analysis of the TBP-binding protein Drl reveals a mechanism for repression of class H gene transcription.” Genes and Development 8(17): 2097-2109. Yokomori, K., A. Admon, J .A. Goodrich, J. Chen and R. Tjian (1993). “Drosophila TFIIA-L is processed into two subunits that are associated with the TBP/TAF 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.” M 8(19): 2313-2323. Yuan, W., G. Condorelli, M. Caruso, A. Felsani and A. Giordano (1996). “Human p300 protein is a coactivator for the transcription factor MyoD.” J. Biol. Chem. 271: 9009- 901 3 . 183 Zawel, L. and D. Reinberg (1993). “Initiation of transcription by RNA polymerase II: a multi-step process.” Prog. Nucleic Acids Res. Mol. Biol. 44: 67-108. Zhong, H., H. SuYang, H. Erdjument-Bromage, P. Tempst and S. Ghosh (1997). “The transcriptional activity of NF-kappaB is regulated by the IkappaB-associated PKAc subunit through a cyclic AMP-independent mechanism.” Ccfl 89: 413-424. Zhong, H., RE. V011 and S. Ghosh (1998). “Phosphorylation of NF-kB p65 by PKA stimulates transcriptional activity by promoting a novel bivalent interaction with with the coactivator CBP/p300.” Mol. Cell 1: 661-671. Zhou, R, P.M. Lieberman, T.G. Boyer and A.J. Berk (1992). “Holo-TFIID supports transcriptional stimulation by diverse activators and fiom a TATA-less promoter.” Genes Dev. 6: 1964-1974. 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(2): 180-187. Zhu, H., V. Joliet and R. Prywes (1994). “Role of transcription factor TFHF in serum response factor-activated transcription.” J. Biol. Chem. 269: 3489-3497. Zwilling, S., A. Annweiler and T. Wirth (1994). “The POU domains of the Octl and 0th transcription factors mediate specific interaction with TBP.” Nucleic Acids Res. 22: 1655- 1662. 184 UNIV 1'11'11111'. 3020 1 {H 11 1'1 3129 1' 1 NICHIGQN STQTE