'4. : Aux—u. add a ., 35"" 37} :r I:r{_: : ‘I‘ m”, «w. .v man-a MW: 1-41.. 41,, I ”ml“: .‘1. . :I,"r‘\ .‘r. v“;- ,. ..,,,,,,,.. II," ".N' .. “.3 0's “-1" I r. ‘( ‘.. '7‘: V ,..., .. 31,“: ' -' K» r} .‘l .97. J; .1 k _ u. in. want-r; «M .4 v4. “mg,“ .21.?“ u. r 4.4 w . ”A“ x: -.n.r~ F I»: u 1 R4. 21m: w. .u‘c‘ ‘- I ... .1. A». a. n ._l r. 13;“ x " ‘. . .V\I!'<}~:, u: mun-.u ”aim: .uu . e- . (wan “—1.1 my“. L a v‘ 71 “L :u. yaw“ r x xfl'm 2,, ~ . r e ' thflC“ _ ”.54.“: ' ~ 1‘ - . J 1 Jr: 1 ' 4: ’x" I‘ll H“, H”, ' {IV-157$ 4. ‘ . ‘ ' I “fry-f, 1'11“: A}. « J-X / 1‘; I ’ P f.-‘.';;,!,w.".€ ‘ .u - v , . , .. ”H A ‘ .. . , ..‘ . , . :, 5,-1.3 “ > :H r- ump”. U‘ u. . ‘ .., , N _ x I ‘ wt- ,.,.. ‘w‘a p x, ”qr“! . , ,, .. , .,.,,..'.,.‘ .. 1. .-. .. 1-; . p.,,. "7.511.- r 1': . " W4 “"17 I‘ ""YN‘{‘I-'.~':""\ "5"“ I . ‘.., ,. ..,. u . . ‘ .. ‘ I ' "r _ _ :u‘ , j! 1. .J,.,',V_:‘J, y _'-.~ I , {/11 'v’v 7‘ ' _',l L‘L‘hi"! 1.1:, . punk}; - i": , (i. uiyr K, At... . ~_ ‘ v I ,L... 41.1... "I "- ‘I ‘ -, V All" n n , l u. ,m.. mun...” “m “u. .I'Lh , w; -. ... l 2. . ,‘U ,‘ ~' 7 ¢ . {5 V f ,l“““.’.§=;" i": .r .4 .. r -_ e..- \4-‘1 :' ‘. 4,..1n "'" ‘65??? SITY LIBRAR E \llllllllllll x All This is to certify that the dissertation entitled CRITICAL STRUCTURAL ELEMENTS OF THE VP16 TRANSCRIPT IONAL ACT IVAT ION DOMAIN presented by William Douglas Cress, Jr. has been accepted towards fulfillment of the requirements for Ph . D . degree in Biochemistry 4 ’ , /Major§ofessor fl Date W? MS U is an Affirmative Action/Equal Opportunity Insriturion 0-12771 t. i . _ m i ; LEERAfiV i gMichigsn Rica, ‘1 i univa r$éfigz i PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE MSU Is An Affirmative Action/Equal Opportunity Institution c:\circ\detedue.pm3-p.1 CRITICAL STRUCTURAL ELENIENTS OF THE VP16 TRANSCRIPTIONAL ACTIVATION DOMAIN By William Douglas Cress, Jr. A DISSERTATION Submitted to Michigan State University ‘ in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Biochemistry 1991 ABSTRACT CRITICAL STRUCTURAL ELENIENTS OF THE VP16 TRANSCRIPTIONAL ACTIVATION DOMAIN By William Douglas Cress, Jr. VP16 is a virion protein of herpes simplex virus type 1 which enters the host cell upon infection and potently and specifically activates transcription of the viral immediate early genes. The domain of VP16 which confers the ability to activate transcription is one of the strongest transcriptional activation domains yet identified. The activation domain of VP16, rich in acidic amino acids, has been modeled as a random coil and alternatively as an amphipathic alpha-helix. However, neither of these models has previously been experimentally tested. In the present work, I have used a combination of site-directed mutagenesis and biophysical analysis in an attempt to determine which elements of structure are critical to the function of the VP16 transcriptional activation domain. Mutational analysis demonstrated that: 1) the VP16 activation domain activates transcription overall as a function of its net negative charge, 2) substitution of helix-breaking proline residues within the predicted amphipathic alpha-helix of the VP16 activation domain does not affect its activity, and 3) a phenylalanine residue at VP16 position 442 (Phe 442) is critical to its function. These observations are inconsistent with proposed random coil and amphipathic alpha-helix models. Ultraviolet spectroscopy revealed that VP16 Phe 442 is buried within the activation domain structure. Furthermore, comparison of the amino acid sequences of a variety of transcriptional activation domains revealed that a pattern of bulky hydrophobic residues is somewhat conserved. These observations suggest that intramolecular hydrophobic interactions may be critical to the structure and function of the VP16 activation domain and of transcriptional activation domains in general. The elements of secondary structure of the VP16 activation domain were directly measured using Fourier-transform infrared and circular dichroism spectroscopies. Spectral analysis indicates that the VP16 activation domain is composed primarily of coil and beta-sheet structures and possesses very little alpha-helix. Whatever you do, whether in word or deed, do it in the name of the Lord Jesus, giving thanks to God the Father through Him. The Apostle Paul in Col. 3:17 iv ACKNOWLEDGNIENTS I thank Steve Triezenberg for being an excellent graduate advisor and role model and I acknowledge his great contribution to the work described herein. I also thank my wife, Andrea Cress, for her tolerance as fellow student and collaborator and I acknowledge her input. I thank Jeff Regier for painstakingly correcting many versions of this dissertation. I also thank Drs. Lee Velicer, Lee McIntosh, Zach Burton and Bob Hausinger for their service as members of my guidance committee. I acknowledge the important contributions of my collaborators, including Drs. Shelley Berger and Leonard Guarente (MIT), Drs. Jim Ingles, Michael Shales and Jack Greenblatt (Univ. of Toronto) and Mark Prairie and Dr. Bill Kreuger (Upjohn Co. , Kalamazoo, MI). Those who have given advice or have loaned equipment have been explicitly referenced within the text. I thank those who took the time to help me learn molecular biology early in my graduate career: Jianli Cao, Don Lorimer, Matthew Morrel, Mark Bloom and Dave Schwab. I thank Steve Triezenberg, Dave Pepper], Andrea Cress, Jeff Regier, Rath Pichyangkura, and Lisa Ortquist for making 522 a pleasant place to work. I acknowledge the Westside Deli, the Korea House and the Thai Kitchen for their contributions to intellectual and not-so-intellectual communication during Thursday Noon. Finally, I thank my parents for their years of support. TABLE OF CONTENTS PAGE LIST OF TABLES ...................................... viii LIST OF FIGURES ...................................... ix LIST OF ABBREVIATIONS ................................ xi CHAPTER I: INTRODUCTION ............................. 1 Eukaryotic Transcription ............................... l Trans-activator Proteins ................................ 3 Mechanisms of Action of Trans-activator Proteins ................ 6 VP16 ........................................... 9 Acidic Transcriptional Activation Domains .................... 11 Overview ......................................... 18 CHAPTER II: ANALYSIS OF THE VP16 ACTIVATION DOMAIN BY SITE-SPECIFIC MUTAGENESIS ............... 20 Introduction ....................................... 20 Construction of VP16 Derivatives .......................... 20 Correlation Between Net Negative Charge and Activation ........... 28 Correlation Between Predicted Amphipathy and Activation .......... 28 Role of Putative Alpha-helical Structure in Activation .............. 33 A Critical Role for Phe 442 in Activation ..................... 36 Conservation of Bulky Hydrophobic Residues Among Various Activation Domains ......................... 43 Conclusions ....................................... 48 vi TABLE OF CONTENTS (cont’d) PAGE CHAPTER III: ANALYSIS OF THE VP16 ACTIVATION DOMAIN BY SPECTROSCOPIC AND BIOPHYSICAL METHODS . . 50 Introduction ...... , ................................. 50 GAL4-VP16 Fusion Proteins ............................. 50 Composition of GAL4-VP16 Fusion Proteins ................... 52 Purification of GAL4-VP16 Fusion Proteins ................... 55 Purification of VP16 Activation Domain Peptides ................ 5 8 Ultraviolet Spectroscopy ............................... 5 9 Fourier-Transform Infrared Spectroscopy ..................... 73 Circular Dichroism ................................... 81 Native Gel Electrophoresis .............................. 83 Differential Scanning Calorimetry .......................... 86 Conclusions ....................................... 90 CHAPTER IV: DISCUSSION ............................... 91 Introduction ....................................... 91 Mechanism of Action of the VP16 Activation Domain ............. 91 Critical Structural Elements of the VP16 Activation Domain ......... 96 Future Studies ...................................... 99 LIST OF REFERENCES .................................. 100 vii LIST OF TABLES PAGE Chapter II Table 1. Mutagenic oligonucleotides ............................ 25 viii LIST OF FIGURES Chapter I Figure l. A schematic representation of a pattern purported to be diagnostic of acidic transcriptional activation domains ........ Chapter 11 Figure 1. Nucleotide and amino acid sequences of the minimal activation domain of VP16 ......................... Figure 2. Relationship between net negative charge and the relative activities of VP16 derivatives .................. Figure 3. Schematic models depicting the predicted amphipathic alpha-helix of the minimal VP16 transcriptional activation domain Figure 4. Schematic helical-wheel model representing the putative amphipathic alpha-helical structure of the minimal VP16 activation domain ..................................... Figure 5. The VP16 minimal activation domain does not activate transcription as a function of its predicted amphipathy ........ Figure 6. VP16 Phe 442 is critical in transcriptional activation . . Figure 7. VP16 derivatives having substitutions at Phe 442 are expressed at levels indistinguishable from the parental protein . . . Figure 8. Substitution of Phe 442 does not affect the binding of VP16 to host DNA-binding proteins ................. Figure 9. Amino acids proximal to Phe 442 are critical in transcriptional activation .......................... ix PAGE .......... 17 .......... 22 .......... 3O .......... 31 .......... 32 .......... 35 .......... 37 .......... 38 .......... 40 .......... 42 LIST OF FIGURES (cont’d) PAGE Chapter II (cont’d) Figure 10. Conservation of bulky hydrophobic residues among activation domains ........................................ 46 Chapter HI Figure 1. GAL4 1-147 and GAL4—VP16 fusion proteins ................. 54 Figure 2. UV spectra of free tyrosine at different pHs .................. 62 Figure 3. UV spectra of Tyr 442 at different pHs .................... 65 Figure 4. Difference spectra of GAL4-del456 and GAL4-FY 442 ........... 68 Figure 5. Difference spectrum of GAL4-VP16 ...................... 71 Figure 6. FTIR spectra of model proteins ......................... 76 Figure 7. FTIR spectra of GAL4-VP16 proteins .................... 78 Figure 8. CD analysis of the full length VP16 activation domain ........... 82 Figure 9. Native gel electrophoresis of GAL4-VP16 proteins .............. 85 Figure 10. DSC scans of GAL4 1-147 and GAL4-VP16 ................ 89 BSA CAPS CAPSO CD DE DSC FPLC HEPES HSV-l ICP LIST OF ABBREVIATIONS absorbance at 230 nm absorbance at 280 nm absorbance at 600 nm amino acid amphipathic alpha-helix adenosine triphosphate basepair bovine serum albumin 3-[cyclohexylamino]- l-propanesulfonic acid 3-[cyclohexylamino]-2-hydroxyl-1-propane sulfonic acid circular dichroism delayed early differential scanning calorimetry fast protein liquid chromatography Fourier-transform infrared N —2-hydroxyethylpiperiazine-N ’-2-ethanesulfonic acid herpes simplex virus type-1 infected cell protein immediate early xi LIST OF ABBREVIATIONS (Cont’d) kb kilobases kd kilodaltons L late LTR long terminal repeat MSV murine sarcoma virus NMR nuclear magnetic resonance NTP nucleotide triphosphate psi pounds per square inch SDS sodium dodecyl sulfate SDS-PAGE sodium dodecyl sulfate-polyacrylamide gel electrophoresis TFII transcription factor of RNA polymerase II tk thymidine kinase TRIS tris(hydroxymethyl)aminomethane hydrochloride UAS upstream activating sequence UV ultraviolet VP16 virion protein number 16 Single letter abbreviations for the amino acids: A, Ala; C, Cys; D, Asp; E, Glu; F, Phe; G, Gly; H, His; 1, Ile; K, Lys; L, Leu; M, Met; N, Asn; P, Pro; Q, Gln; R, Arg; 8, Ser; T, Thr; V, Val; W, Trp; and Y, Tyr. xii CHAPTERI INTRODUCTION Eukaryotic Transcription An important question in biology is how eukaryotic cells regulate the expression of their protein coding genes during development and in response to environmental stimuli. A key regulatory step in the expression of protein coding genes is the initiation of transcription by RNA polymerase II (RNA pol II) which transcribes cellular mRNAs. Unlike the RNA polymerase holoenzyme of eubacteria, catalytic RNA pol H does not recognize promoters by itself (reviewed by Roeder, 1976), but rather requires additional factors (Matsui et a1. , 1980; Davidson et a1. , 1983; Samuels et a1., 1982; Dynan and Tjian, 1983; Dignam et al., 1983). These factors can be classified into two categories; the basal transcription factors which by definition are required for recognition of all class II promoters, and the specific transcription factors which operate at promoters containing specific regulatory sequences. The RNA pol II transcriptional regulatory elements can be divided (somewhat artificially) into two classes; namely the basal elements which are directly recognized by the basal transcription factors and the promoter-specific regulatory elements which are recognized by promoter-specific regulatory proteins. Two basal regulatory elements are the initiator element which includes or lies very close to the transcription 2 start site (Corden et al., 1980; Smale and Baltimore, 1989; Nakatani et al., 1990) and the TATA-box which is generally positioned around 30 nucleotides 5’ to the start of transcription in mammalian cells (reviewed by Breathnach and Chambon, 1981). The initiator and the TATA box element both appear to interact with the basal transcription factor TFIID (see below) and apparently function to establish the start site (Smale e_tjlé, 1990; Nakatani ML, 1990) and direction (Carcamo et_aL, 1990) of transcription. RNA pol II promoters may not require both of these basal elements, since functional promoters lacking each have been identified or constructed (Azizkhan fl, 1986; Myers et_a_l,_, 1986; Smale fl, 1990; Nakatani gal, 1990). In cases where both an initiator sequence and a TATA-box are present, they work in cooperation to yield a stronger basal promoter than is observed with either site alone (Smale and Baltimore, 1989; Nakatani et al., 1990; Conaway et al., 1990). The first step in the assembly of an initiation complex is the step of template commitment (reviewed by Sawadogo and Sentenac, 1990) during which the basal transcription factor TFIID binds weakly to the TATA-box (and / or initiator element) (Fire et_al,, 1984; Reinberg gal” 1987; Buratowski fl, 1989; Smale and Baltimore, 1989; Smale e_t_al._, 1990). Stable binding of TFIID appears to be facilitated by TFIIA (Davidson e_tA” 1983; Fire fl, 1984; Reinberg gal, 1987). Subsequent to stable complex formation, the remaining basal transcription factors and RNA pol II assemble to form the complete preinitiation complex (Davison 1g, 1983; Reinberg etal, 1987; Reinberg and Roeder, 1987; Buratowski e131,, 1989) which is thought to consist of TFIID, TFIIA, TFIIB, RNA Pol II, TFIIE (Ohkuma e_t a, 1990), TFIIF (Flores et al., 1989) or its RNA pol II associated equivalent 3 RAP30/74 (Burton et al., 1988) and the recently discovered TFIIG (Sumimoto et al., 1990). Once the preinitiation complex has formed, there is a requisite hydrolysis of the beta-gamma bond of ATP (Bunick eLaL, 1982) followed by the formation of the first phosphodiester bond. Following initiation, TFIID appears to remain bound at the transcription start site, whereas the other transcription factors must reassemble to initiate transcription a second time (Lin and Green, 1991). In addition to the basal promoter elements, many RNA pol II promoters possess specific regulatory sequences. Many of these promoter-specific elements, such as the CCAAT or GC elements, exert their control within a few hundred nucleotides from the transcription start site, thus they are generally considered part of the proximal promoter (McKnight and Kingsbury, 1982; Myers et al., 1986; Maniatis egg; 1987). Other promoter-specific regulatory sequences can exert control from many thousands of bases away from the start site of transcription. Such promoter elements which increase the rate of transcriptional initiation are termed enhancers (reviewed by Khoury and Gruss, 1983) and the proteins which bind to them are termed Ens-activators (McKnight and Tjian, 1986; Mitchell and Tjian, 1989; Johnson and McKnight, 1989). Promoter-specific regulatory elements which decrease the rate of transcriptional initiation are termed repressors (reviewed by Levine and Manley, 1989). Trans-activator Proteins Many tran s—activator proteins appear to be modular, in that they are composed of at least two separable domains, one of which binds to specific DNA elements and 4 the other of which confers the ability to activate transcription (Brent and Ptashne, 1985; Hope and Struhl, 1986). There are several well defined structural classes of DNA-binding domains, as well as several more poorly defined classes (reviewed by Mitchell and Tjian, 1989; Johnson and McKnight, 1989). The well-characterized classes of DNA-binding domains have recognizable conserved sequence motifs which are often given imaginative names. One such motif is the "Zinc finger" which was originally identified as the DNA—binding structure of TFIIIA (Miller et al., 1985). The Zn-binding domains of TFIIIA are composed of approximately 30 amino acids with two invariant cysteines and two invariant histidines which stabilize the domain by tetrahedrally coordinating a single Zn2+ ion. A second type of zinc binding domain, present in the DNA-binding domains of the steroid hormone receptors, utilizes two pairs of cysteines instead of the cysteine/histidine combination (Freedman gm, 1988). A third type of Zn-containing DNA-binding domain is the binuclear cluster exemplified by the DNA-binding domain of GAL4 (Pan and Coleman, 1990), in which four cysteine residues coordinate two Zn2+ ions. A second type of DNA-binding domain is the "bZip" motif which was first identified as the dimerization/DNA-binding domain of C/EBP (Johnson et al., 1987; Landscultz et_al., 1988). The bZip motif comprises two regions; a basic region (the b in bZip) composed of thirty or so amino acids characterized by net positive charge, and the "leucine zipper" region (the Zip in bZip) which is composed of a stretch of amino acids having four leucine residues positioned at intervals of seven amino acids. Biophysical and genetic characterizations clearly suggest that the leucine zipper region forms a coiled-coiled intermolecular dimer of helices (O’Shea et al., 1989; Hu et al., 5 1990). In complex with DNA, the basic regions of the bZip dimer are thought to surround the DNA with each basic helix malcing direct contact with the major groove of DNA (Vinson et al., 1989). A DNA—binding motif which appears similar to the bZip domain is the helix-loop-helix domain (Murre et al., 1989). The helix-loop-helix domain contains a basic region which appears to be involved in DNA binding and a dimerization region composed of two short putative amphipathic helices separated by a short intervening region which might form a loop (Davis et al., 1987; Voronova and Baltimore, 1990). A third well defined class of DNA-binding domains is the homeodomain which encompasses about 60 amino acids (Scott et al., 1989). The crystal structure of the D. melanogaster protein engrailed homeodomainzDNA complex has recently been solved (Kissinger et al., 1990). The structure of this domain is composed of three helices (1-3). Helices l and 2 pack against each other in an antiparallel arrangement; helix 3 is perpendicular to l and 2. The hydrophobic face of helix 3 makes contact with l and 2, while the exposed hydrophilic face makes direct contact with the major groove of the DNA. The DNA-binding domains which have been extensively characterized appear to form very highly ordered three-dimensional structures and have been found to have invariant sequence elements. In contrast, the activation domains of gaps-activator proteins do not appear to possess clear sequence homology to each other, but rather have been classified into three general categories; acidic, glutamine-rich, and proline—rich, based upon richness of particular amino acids (Mitchell and Tjian, 1989). Activation domains have typically been identified by deletion analysis and 6 have been somewhat characterized using genetics and site-directed mutagenesis. The best studied of these activation domains are those of the acidic class which will be reviewed in detail in a following section. In no case has the three-dimensional structure of a transcriptional activation domain been reported. Mechanisms of Action of Trams-activator Proteins When enhancer regulatory elements were first identified they posed something of a dilemma. How could a regulatory protein bound thousands of nucleotides away from a transcription start site affect the rate of productive transcriptional initiation by RNA pol II? A number of possible mechanisms were proposed to explain the activation phenomenon, which included models termed twisting, sliding, oozing and looping (reviewed by Ptashne, 1986). Presently, the looping model is generally favored. In this model the tran_s-activator binds to its particular fi-regulatory element through its DNA-binding domain, the transcriptional activation domain makes contact with some target protein or proteins involved in transcription and the long stretch of intervening DNA simply loops out (Ptashne, 1988; Ptashne and Gann, 1990). An important and unanswered question in regard to the looping model is what is the target (or targets) of the activation domains of trans—activator proteins. Numerous targets have been postulated (reviewed by Ptashne, 1988; Ptashne and Gann, 1990) and it is becoming evident that indeed a number of mechanisms of tr_an_s-activation may occur. Perhaps the simplest mechanism of activation demonstrated so far is termed antirepression (Croston et al., 1991). This mechanism of transcriptional activation 7 involves the fact that transcriptional regulation occurs in the context of genomic DNA packaged into chromatin structure. It is clear that such packaging is inhibitory of transcriptional initiation in vitrg (Workman et al., 1991; Croston et a1. , 1991) and presumably in vivo (reviewed by Elgin, 1988). It has recently been shown that mans-activator proteins from the glutamine-rich and acidic classes of activators are able to overcome the inhibition of initiation by nucleosomes or histone H1 Mtg; (Workman et_al., 1991; Croston et_al” 1991). This observation indicates that at least one mechanism of activation utilized by gag-activators involves overcoming inhibition by histones, presumably by making a promoter template more accessible to transcription factors. A second demonstrated mechanism of action of a m-activator protein also relies on relieving a repression; in this case not a repression of initiation, but a repression of elongation. This mechanism of activation is exemplified by the hsp 70 promoter of D. melanogaster. Prior to the induction of heat shock, a single RNA pol II molecule is transcriptionally engaged at the hsp 70 promoter, it has formed a nascent RNA chain of approximately 25 nucleotides, but it is apparently arrested at that point (Gilmour and Lis, 1986; Rougvie and Lis, 1988). Upon heat shock, the heat shock transcription factor (I-ISF) , which previously was not bound to DNA, is activated such that it binds to the hsp 70 heat shock response element and through an unknown mechanism frees the arrested RNA pol 11. Activation of transcription by gang-activator proteins, however, is not limited to relieving repression, since transcriptional activation 'mv_itr_o has been observed on naked DNA in the absence of histone proteins using relatively pure preparations of 8 basal transcription factors (Peterson et a1. , 1990). Thus it is possible that some m-activator proteins may activate transcription by recruiting basal factors to the promoter site. For instance, genetic experiments (Allison and Ingles, 1989) and in m experiments using ms-activator proteins as affinity column ligands (Bde and Struhl, 1989) suggest that yeast Ens-activators may interact directly with RNA pol II. Similarly, the TATA element-binding protein TFIID is clearly a candidate for the target of transcriptional activation domains. Experimental evidence that TFIID could be a target of transcriptional activators came from the observation that the transcription factor ATF interacts with the TFHD protein to facilitate the establishment of a preinitiation complex (Horikoshi et a1. , 1988; Hai et al., 1988). Further evidence that TFIID might be a direct target of m-activators came from an experiment using cloned recombinant yeast TFHD (Stringer et_al., 1990). In this experiment the acidic activation domain of VP16 (see below) was used as a column ligand in affinity chromatography. They found that the VP16 affinity column bound very selectively to human TFIID and purified recombinant yeast TFIID. Lin and Green (1991) have recently discovered evidence that the VP16 activation domain specifically interacts with yet another basal transcription factor, TFIIB. In this work, Lin and Green used the VP16 activation domain as a chromatography affinity matrix in a manner similar to Stringer _et_al_. (1990). They passed Hela cell nuclear extracts over the VP16 affinity column and observed binding of TFIID to the column. However, Lin and Green went on to demonstrate that TFIID is not the only basal transcription factor bound to the VP16 column. TFIIB was also bound and appeared to be bound with greater affinity than is TFIID (with 8 basal transcription factors (Peterson et a1. , 1990). Thus it is possible that some m—activator proteins may activate transcription by recruiting basal factors to the promoter site. For instance, genetic experiments (Allison and Ingles, 1989) and 'm \_'_i_tr_'Q experiments using m-activator proteins as affinity column ligands (Bde and Struhl, 1989) suggest that yeast m—activators may interact directly with RNA pol 11. Similarly, the TATA element-binding protein TFIID is clearly a candidate for the target of transcriptional activation domains. Experimental evidence that TFIID could be a target of transcriptional activators came from the observation that the transcription factor ATF interacts with the TFIID protein to facilitate the establishment of a preinitiation complex (Horikoshi et al. , 1988; Hai et al., 1988). Further evidence that TFIID might be a direct target of trans-activators came from an experiment using cloned recombinant yeast TFIID (Stringer et al. , 1990). In this experiment the acidic activation domain of VP16 (see below) was used as a column ligand in affinity chromatography. They found that the VP16 affinity column bound very selectively to human TFIID and purified recombinant yeast TFIID. Lin and Green (1991) have recently discovered evidence that the VP16 activation domain specifically interacts with yet another basal transcription factor, TFIIB. In this work, Lin and Green used the VP16 activation domain as a chromatography affinity matrix in a manner similar to Stringer gal. (1990). They passed Hela cell nuclear extracts over the VP16 affinity column and observed binding of TFIID to the column. However, Lin and Green went on to demonstrate that TFIID is not the only basal transcription factor bound to the VP16 column. TFIIB was also bound and appeared to be bound with greater affinity than is TFIID (with 9 respect to the concentration of KCl used as eluant). Stringer et_g, (1990) did not eliminate the possibility that TFHB was bound to their columns. VP16 Herpes simplex virus type I is a widespread human pathogen (reviewed by Whitley, 1990) which has been used as a model system for the study of gene expression in mammalian cells. HSV-l possesses a large double-stranded DNA genome encoding approximately 70 genes which are transcribed by host RNA pol H (reviewed by Roizman and Sears, 1990). During lytic HSV-l infection there is temporal expression of three major classes of coordinately regulated genes known as immediate-early (IE), delayed early (DE), and late (L) genes (reviewed by Wagner, 1990). Temporal HSV-l gene expression is regulated by a mechanism whereby accumulated IE gene products suppress IE expression and activate the expression of DE genes. DE gene products result in the expression of L genes by initiating replica- tion of the viral genome upon which L gene expression is dependent. A late gene product and component of the HSV-l virion, termed VP16 (also known as Vmw65, alpha-TIF and ICP25), then specifically and potently activates IE gene expression during the next round of infection (Batterson and Roizman, 1983; Campell et_g, 1984; reviewed by Roizman and Spector, 1991). Activation of IE gene expression by VP16 depends upon specific sequence elements present (often in several copies) in all IE 9i; regulatory regions (Mackem and Roizman, 1982; Cordingley et_aL, 1983; Preston fl, 1984; Bzik and Preston, 1986). The sequence motif 5’-TAATGARAT (where R indicates a purine) is the 10 major sis-regulatory factor in VP16-mediated tans-activation (Kristie and Roizman, 1984; Triezenberg et al., 1988a). VP16 is an atypical member of the class of gang-activator proteins since, apparently, VP16 does not have a high affinity for DNA (Marsden fl, 1987 ; Kristie and Sharp, 1990), but rather is tethered to the TAATGARAT element by binding to the ubiquitous octamer-binding protein Oct-l which in turn binds the TAATGARAT (or the overlapping sequence, 5 ’-ATGCTAAT) sequence (Gerster and Roeder, 1988; Preston iii, 1988; O’Hare and Goding, 1988). At least one other host factor appears to be involved in the VP16:Oct-1zDNA complex (McKnight et_al., 1987 ; Kristie and Sharp, 1990; Katan e_tA, 1990), but its role and identity are unknown. A second motif, rich in G and A residues, is also important for maximal VP16-mediated m-activation (Triezenberg et_al., 1988a; Spector et al. , 1990). The host protein, termed IEF that binds to this element has ga’ been characterized (LaMarco and McKnight, 1989), but a direct interaction with VP16 has not been demonstrated. VP16 is present at an estimated 500-1000 copies per virion (Spear and Roizman, 1972; Hiene fl, 1974), but it is unknown how VP16 accompanies the viral DNA to the cell nucleus. The VP16 polypeptide is composed of 490 amino acid residues as inferred from DNA sequence (Pellett et_al, 1985; Dalrymple girl, 1985; Triezenberg et_al., 1988b) with a calculated (Devereux fl, 1984) molecular weight and isoelectric point of 54,316 daltons and 4.7, respectively. VP16 purified from virus particles is a phosphoprotein with an apparent molecular weight of approximately 65 kd as measured by SDS-PAGE (Gibson and Roizman, 1974). 11 Despite the apparent discrepancy between calculated and predicted molecular weights the VP16 polypeptide does not appear to be glycosylated (Hiene et_al_., 1974). Molecular genetic studies of VP16 have divided the polypeptide into two functional domains relevant to its function as a tr__a_n_§-activator. The IE specificity function of VP16 (i.e. the ability of VP16 to interact with Oct-1) is conferred by a poorly defined domain (or domains) contained within the N-terminal 400 amino acids of VP16 (Triezenberg et al. 1988b, Ace et al., 1988, Greaves and O’ Hare 1989, Werstuck fl” 1990). The transcriptional activation function is carried within a domain composed of the C-terminal 80 amino acids of VP16 (Triezenberg et_al_., 1988b; Sadowski et_al., 1988; Greaves and O’ Hare, 1989; Cousens et_al., 1989; Werstuck and Capone, 1989). The 80 amino acid C—terminal domain of VP16 has been shown to be sufficient to function as an activation domain (independent of other VP16 polypeptide sequences) by attaching it to an unrelated DNA—binding domain from the yeast GAL4 protein. This chimeric protein has been found to be among the strongest transcriptional activators tested both in vivo (Sadowski et al., 1988; Cousens et al., 1989) and in vitro (Chasman et al., 1989; Carey et al., 1990a). Acidic Transcriptional Activation Domains Inspection of the amino acid sequence of the 80 amino acid VP16 activation domain reveals that this domain is rich in acidic residues. Yeast transcriptional activator proteins GAL4 (Ma and Ptashne, 1987a and b), GCN4 (Hope and Struhl, 1986), and HAP4 (Forsburg and Guarente, 1989), and a variety of mammalian hormone receptors (Hollenberg and Evans, 1988; Godowski et al., 1988; reviewed by 12 Evans, 1988; Zhu fl, 1990) also contain activation domains which possess net negative charge, suggesting that negative charge is a critical component of at least one class of eukaryotic transcriptional activators (reviewed by Mitchell and Tjian, 1989) . GAL4 is a yeast protein of 881 amino acids which is required for the expression of the Saccharomyces oerevisiae genes needed for galactose metabolism (Douglas and Hawthorne, 1964). GAL4 binds to upstream activating sequence (U AS G) of galactose metabolizing genes and activates their transcription. GAL4 binds to UASG DNA through an N-terminal domain (amino acids 1-147) and possesses two acidic transcriptional activation domains I and II, encompassed by amino acids 148-196 and 768-881, respectively (Ma and Ptashne, 1987a). In an effort to study the critical elements of an acidic activation domain of GAL4, Gill and Ptashne (1987) chemically mutagenized the GAL4 DNA region I activation domain. The authors characterized seven mutations that increased the activity of the GAL4 region 1 activation domain, five which retained partial activity, and twenty which were completely inactive. All of the mutations which increased the activity of the mutated GAL4 protein increased the net negative charge of the domain by replacing positively charged residues with neutral counterparts. However, no improved activators were obtained resulting from the addition of negative charge into the domain, suggesting that indiscriminant placement of negative charge into the domain was not an easy way to increase its strength as an activator. Of the five partially active mutants characterized, two neutralized acidic residues, two replaced hydrophilic residues with hydrophobic counterparts and one introduced a stop codon which severely truncated the domain. The observation that two mutations at noncharged 13 residues decreased the activation function of GAL4 suggests that some structural requirement in addition to net negative charge is important to the function of the domain. Of the 20 completely inactive GAL4 derivatives which were isolated none encoded the entire GAL4 gene protein. The authors suggest that their inability to isolate single (or even multiple) amino acid substitutions that completely inactivated the domain was evidence of its functional redundancy. In an accompanying paper, workers from the same lab reported a novel type of genetic experiment aimed at understanding what elements are critical to the function of acidic activation domains (Ma and Ptashne, 1987b). In this experiment, the authors fused "random" 8J1; 3A fragments of M genomic DNA to the gene encoding the GAL4 DNA-binding domain (GAL4 amino acids 1-147). They then screened for fusion proteins that functioned as transcriptional activators in yeast. The strength of the activating hybrids roughly correlated with net acidity of the amino acids coded by the random m DNA fragments. The activating E43111 sequences showed no extended homology to each other (particularly in relation to distribution of negative charge) or to the native GAIA activation domains. Hope, Mahadevan and Struhl (1988) characterized an acidic activation domain from the yeast GCN4 protein (Hope and Struhl, 1986) by high resolution deletion mutagenesis. In this experiment the authors fused twenty-six carboxyl-terminal deletions of the GCN4 activation domain to the 100-residue GCN4 DNA-binding domain and measured the transcriptional activation function of each derivative in an i_n yi_vo assay. The activities of these deletion derivatives fit a simple pattern; a series of small stepwise decreases in activity were observed as amino acids (and net negative 14 charge) were truncated. The authors argue from this data that the activation domain of GCN4 must be unlike rigid structural elements, such as active sites of enzymes, where one would expect some point where there would be a sudden loss of activity. The authors suggest that the GCN4 activation domain represents a "repeated structure composed of small units acting additively" and that it might possibly function as "a dimer of alpha-helices from two GCN4 monomers ". The results of the previously described work on GAL4 and GCN4 suggest that yeast transcriptional activation domains require some particular distribution of negative charge (Gill and Ptashne, 1987 ; Ma and Ptashne, 1987b), but are functionally redundant (Hope gal, 1988). Unfortunately these studies did not identify any sequence homologies that would suggest structural similarities among the various activating peptides. This observation is not a totally novel one in that, as pointed out by Ma and Ptashne (1987b), there are other examples of protein sequences which bear little sequence homology, but which possess identical biochemical activities. For instance, the sequences which identify proteins for secretion bear little sequence homology (reviewed by Briggs and Gierasch, 1986), nor do the signal sequences of mitochondrial proteins (von Heijne, 1986; Schatz, 1987). Indeed, it has been demonstrated that both the signal sequences of secreted (Kaiser e_t al, 1987) and of mitochrondrial proteins (Baker and Schatz, 1987) can be functionally replaced by a large fraction of random peptide sequences. The apparent lack of sequence homology among various acidic transcriptional activation domains led X-ray crystallographer Paul Sigler (1986) to propose that acidic activation domains may function, more or less, as random coils and that a 15 specific secondary structure is not a critical requirement. This model is not inconsistent with mest of the results just described. However, the "acid blob" or "negative noodle" model (as it has been dubbed) is difficult to reconcile with the observation that mutations not affecting charge of the GAL4 region I activation domain can nonetheless have significant effects on activity. A second model (Gininger and Ptashne, 1987; Ptashne, 1988; Hope et_all, 1988) proposed that acidic activation domains form amphipathic alpha-helices (AAHs) possessing opposing hydrophobic and acidic "faces" (a tendency toward a third hydrophilic face was also suggested, as well as dimerization mediated by the hydrophobic faces). This model stems from the observation that many of the acidic activation domains identified would indeed form such structures if they folded into alpha-helices. To test this model Gininger and Ptashne (1987) constructed a DNA fragment encoding a peptide which could potentially form a four-tum, negatively charged, alpha-helix having alternating hydrophobic, hydrophilic and acidic faces. In addition, a second DNA fragment encoding a peptide which might also form a four—tum, negatively charged, AAH, but in which the hydrophilic face of the first peptide (glutamine residues) was replaced by a hydrophobic face (valine residues). As controls, two peptides of identical composition were constructed which would not form amphipathic structures if folded into alpha-helices. Each of these peptides were fused to the DNA-binding domain of GAL4 (amino acids 1-147), the fusion proteins were expressed in yeast and assayed for activation of a GALl—ng fusion gene. Only the two peptides which could form AAHs were functional. The peptide proposed to have alternating hydrophobic, hydrophilic and acidic faces (termed AH) functioned 16 more than half as well as the region I activation domain of GAL4 while the peptide which could potentially have only acidic and hydrophobic faces functioned only 5 % as well. The control peptides which could not form amphipathic structures were not functional. The suggestion that acidic activation domains might function as negatively charged AAHs prompted a group led by Temple Smith and James Figge to use a supercomputer system [ARIEL (Lathrop mg, in press)] to compare the sequences of various known acidic activation domains in order to determine if a common sequence/ structural motif pattern among these activator proteins could be identified if predictions of secondary structure were taken into account rather than simple sequence homology (Zhu ML, 1990). Using such a procedure the group identified a pattern diagnostic of acidic transcriptional activation domains. This pattern (represented schematically in Figure 1) consisted of nine components: (N to C-terminal) a local hydrophobic minimum (often matched to a region containing proline), a spacer of -1 to 7 residues, followed by a predicted alpha-helix overlapped with the following charged template; charge-charge—any residue-any residue-charge—any residue-any residue-charge—charge, a spacer of —l to 8 residues, a local hydrophobic maximum containing a valine, leucine or isoleucine, a spacer of 4-8 amino acids and finally a predicted beta-turn with an aspartic acid residue between 5 and 0 amino acids from the N-terminal end of the beta-tum. The element most heavily weighted in this analysis was the predicted alpha-helix having a negatively charged face. Various acidic activation domains were found to fit this pattern including activation domains from viral activators (such as the VP16 activation domain), steroid hormone (4.8) B-lum (5.010 [v1.11 ’1 Local Hydrophobic profile maximum wuh ('1-8) Schematic Elements alpha-helix with acid faced segment (~1.7)u Local Hydrophobic profile minimum Pattern figure 1. A schematic representation of a pattern purported to be diagnostic of acidic transcriptional activation domains. A. Ribbon model representation of an amphipathic alpha helix. B. The nine elements of the diagnostic pattern refined by ARIEL (Lathrop et al., in press). This figure was taken from Zhu et al., 1990. l... l . 18 receptors, yeast transcriptional activator proteins and some of the acidic activation domains isolated from E. coli (Ma and Ptashne, 1987b). However, five of the nine strongest activating sequences isolated from E. coli did not fit this pattern (B17, B7, Bl, B32, and B3; Ma and Ptashne, 1987b). Overview The observations of Zhu et al. (1990) suggest that various acidic transcriptional activation domains may have some structural similarities including an AAH. Unfortunately the structural similarities identified by Zhu et al. are based completely upon structural predictions which are known to be unreliable. Thus it becomes imperative to test this model experimentally. The VP16 acidic activation domain appears to be the strongest yet identified, thus, it is an ideal candidate for studying acidic activators in general and is the topic of this dissertation. In the next chapter of this dissertation I describe work in which I sought to test the various models regarding the structure of the VP16 activation domain using mutagenesis. In that work I made the following observations: 1) the VP16 activation domain activates transcription overall as a function of its net negative charge, 2) the predicted AAH of the VP16 activation domain did not appear to be very important, 3) Phe 442 of the VP16 activation domain was critical to its function, and 4) there is some conservation of bulky hydrophobic residues among the sequences of various activation domains from different classes. My observations support the suggestion of Zhu et al. (1990) that various activation domains do have structural similarities (similarities which probably center around a hydrophobic core), but demonstrate that an AAH is probably 19 not a necessary component of their structure. In chapter 111, I describe work which was aimed at directly examining the structural elements of the VP16 activation domain. In that work, I used ultraviolet spectroscopy to demonstrate that the critical phenylalanine, Phe 442, of the VP16 activation domain is buried within the activation domain structure. This observation supports a model that hydrophobic interactions are very important in maintaining the structure of the VP16 activation domain. I then utilized infrared and circular dichroism spectroscopies to demonstrate that the VP16 activation domain contains virtually no helical structure; a result consistent with the observation from chapter 11 that alpha-helix was not a critical element of the activation domain structure. In chapter IV, I discuss alternative models regarding the structure of the VP16 activation domain and propose future studies. Furthermore, I discuss collaborations and other investigations in which a number of the VP16 mutants described in this dissertation have been used as reagents to address the question of how the VP16 activation domain may interact with other transcription factors to increase the rate of productive initiation. This section ties the relevance of my work to the larger arena of transcriptional control. CHAPTER II ANALYSIS OF THE VP16 ACTIVATION DOMAIN BY SITE-SPECIFIC MUTAGENESIS Introduction The eighty amino acid VP16 activation domain is rich in acidic amino acid residues, as are the activation domains of the yeast proteins GCN4 (Hope and Struhl, 1986), GAL4 (Ma and Ptashne, 1987a), and HAP4 (Forsburg and Guarente, 1989), and numerous mammalian hormone receptors (reviewed by Evans, 1988). The preponderance of negative charge among these activation domains suggests that negative charge is a critical component of activation domain structure. I began my characterization of the VP16 activation domain by measuring the correlation between the ability of the VP16 activation domain to activate transcription and its total net negative charge. For this purpose, I generated a series of activation domain mutations in which the acidic amino acids aspartate and glutamate of the VP16 activation domain were replaced with uncharged asparagine and glutamine residues. Construction of VP16 Derivatives Figure 1 shows the region of the minimal VP16 activation domain targeted for mutagenesis (VP16 codons 427 to 451), which includes ten acidic amino acid residues. This minimal activation domain, previously identified by deletion 20 21 Figure 1. Nucleotide and amino acid sequences of the minimal activation domain of VP16. The amino acids targeted for mutagenesis are highlighted in bold letters; the acidic residues are numbered 1-10. This 135 basepair E I/ BlamH I DNA fragment [which originates from pMSVPl6 del456 (Triezenberg et al., 1988b)] was cloned into the phage vector M13 mp19 (Yanisch-Perron gal, 1985) to generate a single-stranded template for oligonucleotide—directed mutagenesis. The codons for the ten acidic amino acids were mutagenized in four clusters using mixed-sequence oligonucleotides (Table I) to generate all combinations of desired substitutions within each cluster. Aspartate codons (GAT or GAC) were altered to asparagine codons (AAT or AAC), and the glutamate codon GAG was altered to the glutamine codon CAG. The Ng I and m I restriction sites were introduced using translationally silent nucleotide changes, to facilitate construction of intercluster combinations. Altered activation domain DNA fragments were reintroduced into the parental VP16 expression plasmid, pMSVPl6 del456, and mutant plasmid identities were confirmed by sequence determination (Sanger et al., 1977). 22 EEmm 05.3000 sz m 0 m omv ovv H00 000 OO._. .20 000 O_ ”.9920 :_ Gaza _ 3:2 _ ooz omv Enadolqno 80 .20 80.94.30 90 98 .98 000 95 EL 05 ._ < o < I < 2 < > o m o a 4 z a m N F __ 530 final omv _ _mm 0 So O._.O O o ._. d n. < ._. m 4 Figure 1 23 mutagenesis, retains about one-half of the activity of the full-length domain both i_n_ m (Triezenberg et_al., 1988b) and mm; (Berger gal, 1990). Use of this minimal activation domain reduced the target size for mutagenesis, and in principle permitted the detection of variants with increased activity, since the truncated domain possessed only half the full-length activity. The minimal activation domain of Figure 1 was divided into four "clusters" which were mutagenized using mixed-sequence oligonucleotides (Table 1). These mixed-sequence oligonucleotides were synthesized by incorporating equal proportions of wild-type and mutagenic nucleotides at each of several positions selected to give desired mutations. In addition these oligonucleotides generated an N_c_o I site between clusters I and II and an _1V_11_u I site between clusters H and III to allow intercluster recombination to generate activation domain derivatives with as many as ten negative charges replaced. The details of the constructions of these derivatives are as follows. The plasmid pMSVP16 del456 was used as a source of VP16 DNA for mutagenesis and has been described (Triezenberg fig, 1988b). The VP16 minimal activation domain sequence (from Figure 1) was prepared for mutagenesis by cloning the 135 bp Sell 1/ M I fragment of pMSVPl6 del456 into the phage vector M13 mp19 to generate mpDCl. The system of nomenclature used to refer to mutants emerging from mutagenesis utilizes the numbers assigned to the ten negative residues highlighted in Figure 1 and the single letter codes for the amino acids aspartate (D) and asparagine (N). A mutant changing negative residues 1, 2 and 5 from acidic to neutral is termed DN 125, while a mutant changing all ten negative residues is termed DN 1-10. For 24 Table 1. Mutagenic oligonucleotides. The mixed oligonucleotides used to alter the VP16 gene are represented in this table. Parentheses indicate mixed positions. All oligonucleotides were purchased from the MSU Department of Biochemistry Macromolecular Facility and were used without further purification. Mutagenesis reactions were performed according to standard protocols (Zoller and Smith 1982; Kunkel, 1985; Bio-Rad Catalog #170-3571). Single-stranded uracil-containing template DNA was prepared by growing M13 phage in E, coli host CI 236 (gitgl, u_ngl, rm, relAl; pCJ105(Cm') which incorporates uracils at about 1% of the normal thymidine residues. Uracil containing phage were purified by equilibrium sedimentation in CsCl, phage DNA was extracted by treatment with phenol/chloro- form, and circular template was purified by electrophoresis. Synthesis of the second strand was performed i_n m using a mutagenic oligonucleotide as primer. Mutagenic primers were annealed to templates in 20 mM TRIS-HCI (pH 7.4), 2 mM MgCl, 5.0 mM NaCl by heating to 65 °C and cooling 1 C/minute to 4 °C using a standard primer/template ratio of 20:1 (Bio—Rad Catalog #170-3571). Synthesis of the complementary DNA strand was done in 23 mM TRIS-HCl (7 .4), 5 mM MgC12, 35 mM NaCl, 1.5 mM dithiothreitol, 0.4 mM deoxynucleotide triphosphates, 0.75 mM ATP, 0.2 Units/p1 T4 DNA ligase, and 0.1 Units/pl T4 DNA polymerase. The synthesis reactions were carried out five minutes at 4 C, five minutes at 25 °C and 90 minutes at 37 °C and were terminated by a 20:1 dilution in water followed by freezing. Mutagenic synthesis reactions were then used to transform competent dut+ ung” MV1193 cells {(Alac-pro AB), thi, supE, A(srl-rec A)306::Tn10(tet’)[F’:tra D36,pro.AB lac IqZAM151}. Transformation of this duplex DNA into dut+ ung+ E, poll results in selection against the uracil-containing wild-type strand due to the action of uracil glycosylase. Transformant MV1193 plaques were screened for mutations by dideoxy sequencing (Sanger gal, 1977). Controls in mutagenesis procedures included uracil-containing template without added primer, uracil-containing template plus a control primer having no mismatches, and uracil-containing template with both control and mutagenic primers. Control reactions without added primer indicated that less than 2% of mutagenic transformants resulted from endogenous priming and from template molecules not inactivated by incorporation of uracil. 25 Table 1 Mutagenic oligonucleotides ST 1a ST lb ST 2 ST 3 ST 4 ST 9 ST 10 ST 11 ST 14 ST 15 ST 16 5’-CATGGCCACGT(t/c)CT(g/c)G(t/c)CGT(t/c)TAAGTG-3’ 5’-CATGGCCACGT(t/c)CTGG(t/c)CGT(t/c)TAAGTG-3’ 5’~GTCTAACGCGT(t/c)G(t/g)CA(t/c)(c/g)CGC-3’ 5’-CATGT(t/c)CAGAT(t/c)GAAAT(t/c)GT(t/c)TAACGCG-3’ 5’-GGAAT(t/c)C(t/c)CGT(t/c)C(t/c)CCAA-3’ 5’—GCCGTCTAAGGGGAGCTCGT-3’ 5’-CGCGTCGGGATGCGCCATGGGCACGTC-3’ 5’-GTCCAGATCGGGATCGTCTAA-3’ 5’-GTCCAGATCG(t/g)(a/c)ATCGTCTAA-3’ 5’-CAACATGGCCAGAGCGAAAGCGGCTAACGC-3’ 5’-CAACATITCCAGTTCGAATTC'I'TCTAACGC-3 p" ’3 26 simplicity this nomenclature ignores the fact that acidic residue number two is glutamate not aspartate. Mutagenesis of cluster I was done using mpDCl as single-stranded template and the mixed—oligonucleotide primers ST 1a and b (see Table 1). Cluster I mutagenesis resulted in the isolation of the pseudo-wildtype mutant mpDC2 (which contains an artificial Nco I site between clusters I and II) and the mutants mpDC DN 1, 2, 3, 12, 13, 23 and 123. Mutagenesis of mpDC2, using ST 2 as mutagenic primer, generated the mutant mpDC DN 4 and a second pseudo-wildtype mutant, mpDC3, which possessed artificial Ng I and m I sites, between clusters I and II, and, II and 111, respectively (see Figure l). The pseudo-wildtype mutant mpDC3 was used as the template for all subsequent mutagenesis. The cluster III sequence was altered using ST 3 which generated the mutants mpDC DN 5, 6, 7, 8, 56, 57, 58, 67, 68, 78, 567, 568, 578, 678, and 5678. Mutagenesis of cluster IV was accomplished using the oligonucleotide ST 4 which generated the mutants mpDC DN 9, 10 and 910. Combinations between HI and IV were made by taking the two cluster IV single mutants, DN 9 and DN 10, and performing ST 3-mediated mutagenesis upon them. All fifteen possible combinations from each template were isolated. Thus, I was able to construct all necessary combinations of the ten targeted acidic residues simply using the Ngo I site between clusters I and II, the ltflu I site between clusters II and III and a unique Bgl 11 site within the M13 mp19 vector. Minimal activation domain derivatives were reintroduced into the pMSVPl6 de145 6 vector which drives the expression of VP16 del45 6 using the MSV-LTR (which includes a transcriptional enhancer, promoter and mRN A cap site) within the 27 vector 5’ to the VP16 gene (Graves gl 1985). Termination codons and polyadenylation signals were also provided by the vector (Triezenberg et_al, 1988b). Plasmid pMSVP16 C +119bp (which has two S_a_l_ I sites and two BgmH I sites) was partially digested with Slal I and single-cut, linearized plasmid was purified by agarose gel electrophoresis. The E I linearized vector was methylated using Msp I methylase (New England Biolabs) to protect one BgmH I site of the pMSVPl6 C +119 bp vector while the other EmH I site was digested to completion with M I. The resulting 6.1 kb pMSVP16 B_arr_1H I/ E I vector was purified by agarose gel electrophoresis. Mutant 135 bp MH I/S_al_ I activation domain fragments were excised from the various mpDC clones by digestion with Ba_mH I and fl 1, electro- phoretically purified and cloned into the 6.1 kb pMSVPl6 BamH I/fl I fragment. The ability of VP16 derivatives to activate IE gene transcription was measured by a transient transfection assay (Triezenberg gal, 1988a) in which a plasmid expressing the wildtype or a mutant VP16 protein was cotransfected with both an indicator and an internal control plasmid. The indicator plasmid pSJT-703 contained the IE upstream regulatory sequences of the HSV~1 ICP4 gene, fused to the body of the HSV-l tk gene. The second plasmid, pMSV-tk, served as an internal control for transfection efficiency, RNA recovery and primer extension assays. Transcription of tk-specific RNA from this plasmid is not activated by VP16 (Triezenberg gal, 1988a). The IE—tk and MSV-tk RNAs yielded primer extension products of 81 and 55 bases, respectively, which were separated by electrophoresis, detected by autoradiography and quantitated by scintillation spectroscopy. 28 Correlation between net negative charge and transcriptional activation. From the relative activities of a collection of 21 such mutants, summarized in Figure 2, I infer that negative charge is an important feature of the VP16 activation domain. Replacement of increasing numbers of acidic residues with uncharged residues led to a progressive decrease in transcriptional activation. Removal of seven or more negative charges was required to inactivate the protein beyond the level of detection of this assay. However, net negative charge is not the sole determinant of activity, since some derivatives with identical net negative charge had activities which were clearly distinguishable. For instance, the VP16 derivative DN 2 is less active than DN 6; DN 1234 is less active than DN 5678; DN 1-4,910 is less active than DN 5-10; and DN 123,5-8 is less active than DN 4-10. In each of these cases, mutations of negative residues 1, 2 and 3 tend to have a greater effect on activity than mutations at the other positions. These observations suggest that either charge distribution and/or secondary structure also contribute to activity. Correlation Between Predicted Amphipathy and Activation It has been proposed that AAHs (Gininger and Ptashne, 1987; Ptashne, 1988; Zhu et al. , 1990) are critical structural elements of acidic activation domains. Furthermore, secondary structure predictions for the VP16 amino acid sequence suggested that the domain could form an AAH (Cousens et_al, 1989). Schematic models representing the minimal VP16 activation domain, should it form an alpha-helix, are shown in Figures 3 and 4. I tested the role of predicted amphipathy by constructing a series of VP16 derivatives in which I removed blocks of L: 29 Figure 2. Relationship between net negative charge and the relative activities of VP16 derivatives. The relative activities of twenty-one VP16 mutants are plotted as a function of the number of remaining acidic amino acids. The derivative designations indicate the amino acid substitutions made (DN means Asp to Asn) and the amino acid positions affected (numbered 1 through 10; see Figure 1). For simplicity this nomenclature ignores the fact that acidic residue number two is glutamate not aspartate. Bars represent two standard deviations about the mean value obtained from at least six transient expression assays. Method. Transfections were performed essentiall as described by Triezenberg et al. 1988b. One day prior to transfection, 5 x 1 mouse L cells (tk', aprt') were plated onto 60 mm culture dishes in Dulbecco’s modified Eagle’s medium, supplemented with 10% fetal calf serum. CsCl-purified DNAs were transfected into the cells using the DEAE-dextran—mediated method (Lopata gal, 1984). Each plate received two pg of the ICP4—gt fusion plasmid (pSJT-703), two pg of the internal control plasmid (pMSV-tlg) and 50 ng of activating plasmid (pMSVP16 de145 6 or derivative thereof). Control plates did not receive inducing plasmid. Forty-two hours after transfection, total RNA was harvested by the proteinase K/ DNAse I method (Eisenberg gal, 1985). A 32P labeled synthetic oligonucleotide which hybridizes to 1k mRN A was extended by reverse transcriptase in presence of deoxynucleotides. Extension products were separated by electrophoresis, detected by autoradiography and quantitated by scintillation spectroscopy. The relative abilities of VP16 mutants to activate IE transcription was calculated from the ratio of indicator signal (pSJT-703) to internal control signal (MSV-tk), normalized to the parental clone, pMSVPl6 del456. 30 3:3 9.23 222 a .3532 O _ N n .v m o h m m 07:3 2.6.3.3 o-.zo 90.3.5 _ _ _ _ _ 94 2o 231.20 a-.. 20 07¢ 20 3&5 . o3 3:3 3 1. 2.»... 2o 8:3 aoeza 232° «.3 czo .L J 0320 nzo r .E oz o\oON $o¢ n3.0m $00 $00. eSON. Kumov alumnae Figure 2 31 Figure 3. Schematic models depicting the predicted amphipathic alpha-helix of the minimal VP16 transcriptional activation domain. A. The upper half of this flattened-helix model for VP16 amino acids 425 to 451 shows the preponderance of negatively charged residues [boldface, numbered (by superscript) as in Figure 1.] on one surface of the putative alpha-helix. The lower half displays a predominantly hydrophobic surface. B. A ribbon model of the VP16 minimal activation domain if it is helical. Negative residues are highlighted as circles and are numbered 1-10. 32 Figure 4. Schematic helical-wheel model representing the putative amphipathic alpha-helical structure of the minimal VP16 activation domain. Numbers within smaller circles identify positions of amino acids within the VP16 polypeptide chain. Acidic amino acids are represented by filled circles. 33 four negatively charged residues circularly permuted about the putative AAH. I reasoned that activation domains lacking charged residues in the center of the charged face of the AAH would likely be strongly affected in transcriptional activation, whereas mutants lacking charged residues on the periphery of the charged face might be only slightly affected. Helical wheel models and relative activities of this series of derivatives are shown in Figure 5 . Although the activities of these mutants display a four-fold range, no correlation between transcriptional activation and predicted amphipathy is observed. For instance, the mutants DN 34610 and DN 34510 each lack negatively charged residues in the center of the charged face of the putative AAH, yet their activities are higher than the mutants DN 1789 and DN 2789 which retain negative charges on the putatively exposed surface. Role of Putative Alpha-helical Structure in Activation Because I observed no relationship between predicted amphipathy and transcriptional activation, I questioned whether the function of VP16 depends on the predicted helicity between amino acids residues 421 and 444. The cyclic side-chain of proline cannot be accommodated into an alpha-helix; therefore, proline substitutions in the activation domain should alter local helical structure. Therefore, I replaced selected amino acids with proline residues to disrupt the predicted alpha-helical structure in three regions across the activation domain. This series included two constructs with single substitutions at His 425 (HP 425) or Phe 442 (PP 442), and one construct with substitutions at Ala 432 and Ala 436 (AP 432/6). This double mutation was constructed because this region possessed the greatest predicted 34 Figure 5. The VP16 minimal activation domain does not activate transcription as a function of its predicted amphipathy. In the schematic helical-wheel models that represent the predicted structures of ten VP16 derivatives, filled circles represent wildtype acidic amino acids and numbered circles represent altered sites. The mean relative activity of each derivative is presented along with standard deviations determined from at least six transient expression assays. Method. Transient expression assays were performed as described in the legend of Figure 2. These derivatives were constructed in the context of the M13 vector by intercluster recombination among appropriate derivatives utilizing the artificial N_co_ I and MI! I restriction sites generated between clusters I and II and H and HI and a unique Bgl 11 site within the M13 mpl9 vector. Following recombination activation domain fragments were reintroduced into the pMSVPl6 background as described in "Construction of VP16 Derivatives" section. 35 o\o w “NV O_mvN 20 fig o\om H8 89 20 Figure 5 36 helix-forming tendencies. Figure 6 shows the effects of these changes on activation of IE gene expression. Surprisingly, neither the single substitution HP 425 nor the double substitution AP 432/ 6 had a detectable effect on activation by VP16, suggesting that the predicted helical structure is not critical in activation. The substitution of proline at Phe 442 however, essentially abolished transcriptional activation. A Critical Role for Phe 442 in Activation To determine whether the effect of the proline substitution of Phe 442 was due to local disruption of helical structure or to the elimination of a critical side-group, I replaced Phe 442 with three helix-compatible amino acids, serine (F S 442), alanine (PA 442) and tyrosine (FY 442). The activation of IE gene expression by these three derivatives is shown in Figure 6. Only tyrosine was able to functionally replace Phe 442, producing an activation domain with about one-third (35 j; 15%) the wildtype activity. In contrast, substitution of Phe 442 with the non-aromatic residues alanine and serine potently inactivated VP16. These results demonstrate that the side-group of amino acid 442 has a critical role in VP16 transcriptional activation. To be certain that the effects of the Phe 442 substitutions were due to structural changes specific to the activation domain, it was important to demonstrate that these substitutions affected neither the stability of the derivative polypeptides nor the function of the VP16 IE specificity domain. Western blot analysis (Figure 7) indicated that the loss of activity of the PP 442 derivative could not be accounted for by diminished accumulation of the derivative polypeptides in transfected cells. 37 'E'“‘ In -- Control --~ HP 425 AP 432/6 FP 442 FS 442 FA 442 FY 442 no VP16 wt Figure 6. VP16 Phe 442 is critical in transcriptional activation. Autoradiogram of a transient expression assay of several VP16 derivatives designed to test the role of predicted helicity in VP16 activation. The positions of primer extension products corresponding to transcripts from the reporter (IE-g) and control plasmids are indicated. Method. This transient expression assay was performed as described in the legend of Figure 2. The mutants HP 425, AP 432/6 and FF 442 were generated using the oligonucleotides ST 9, ST 10 and ST 11 of Table 1, respectively. The mutants FS 442, FA 442 and FY 442 were generated using the mixed oligonucleotide ST 14 (Table l). 38 97kd—~ 68kd—\ 45kd—\ ‘59 =rwww 29kd—\ to a. a 8 (ommNNNN .ia‘WHH; 0.0. EggI 1 co <1- > V co <2- 0 o. a) E o n. a: m t: u. '5 '5 C U. '0 '0 not infected infected Figure 8. Substitution of Phe 442 does not affect the binding of VP16 to host DNA-binding proteins. In this activation interference assay mouse L cells were transfected with 3 pg of plasmid expressing VP16 derivatives as indicated below each lane along with the indicator and internal control plasmids described in the legend of Figure 2. Two days later cells were infected (the first four control lanes were not infected) with HSV-l at a multiplicity of infection of 10 plaque forming units per cell in the presence of 100 pg/ ml cycloheximide. 'vao hours postinfection RNAs were isolated and analyzed by primer extension. 41 probably other hydrophobic residues) is involved in a hydr0phobic interaction either in self-folding of the activation domain or in contact with the molecular target of VP16. The role of hydrophobic interactions in protein self-folding (Lesk and Chothia, 1980) and in protein: protein interactions such as antigen-antibody complexes (Amit et_al_._, 1986; Standfield gg, 1990) is well established. Disruption of that hydrophobic interaction, either by addition of a polar hydroxyl group onto the aromatic ring (as in the case of FY 442) or by removal of the bulky hydrophobic side-group (as for PA 442), resulted in a dramatic reduction of transcriptional activation by VP16. The environment around the critical residue Phe 442 may also be important for activation. I noted that the VP16 mutant DN 567 8, in which the four aspartic acid residues closest to Phe 442 (Figure 1) have been replaced by asparagine, is relatively active (62 i 11%). If secondary structure is important in that local region, then the substitution of asparagine for aspartic acid must conserve that structure. To determine whether making more radical changes in the vicinity of Phe 442 would have stronger effects on activity, I changed the four proximal aspartic acid residues to glutamate (DE 567 8) or alanine (DA 567 8). Glutamate substitution was chosen as a conservative change which is helix favorable, conserves negative charge, and has only a slightly different side-chain structure than aspartate. Alanine substitution was chosen to be compatible with alpha-helical structure, but conserving neither charge, polarity nor side-chain structure. The autoradiogram shown in Figure 9 reveals a hierarchy of relative activities; mutant DN 5678 (62 j; 11%) is stronger than DE 5678 (40 i 10%), which is 42 IE-tk .a. Control “m no VP16 DE 5678 DN 5678 DA 5678 wt Figure 9. Amino acids proximal to Phe 442 are critical in transcriptional activation. Autoradiogram of a transient expression assay of three VP16 derivatives designed to test the importance of the four acidic residues in the local environment of Phe 442. Method. This transient expression assay was performed as described in the legend of Figure 2. The VP16 derivatives DE 5678, and DA 5678 were generated using the oligonucleotides ST 15 and 16 of Table 1, respectively. 43 stronger than DA 5678 (15 j; 5%). Thus, asparagine is the best replacement for aspartate in the structure of the activation domain despite the fact that it does not have negative charge. Glutamate is a slightly poorer replacement than asparagine, even though it conserves negative charge and is very helix-favorable (Chou and Fasman, 1978). Finally, alanine, although compatible with helical structure, is a very poor substitute for aspartate. I conclude that polar or charged amino acids flanking Phe 442 are also important to the VP16 activation domain structure. Conservation of Bulky Hydrophobic Residues Among Various Activation Domains Two models have been previously put forth to describe the structure of acidic transcriptional activation domains. The first model proposes that such domains exist as poorly structured polypeptides whose strengths are simply related to the net charge (Sigler, 1988). This model is inconsistent with the results reported here, which indicate that the acidic amino acids are important but not sufficient for activation by VP16. The second model suggests that acidic activation domains form AAHs (Gininger and Ptashne, 1987; Ptashne, 1988; Zhu girl, 1990). This model, too, is inconsistent with my results. Disruption of predicted amphipathy or helicity did not generally correlate with activity. Furthermore, the failure of mutant FA 442 to activate transcription argues strongly that the predicted acidic AAH is not sufficient for the function of VP16. I have reexamined the primary sequences of VP16 and other transcriptional activators in an effort to identify some common features consistent with the results 44 presented in this report. In Figure 10, the amino acid sequences of several activation domains from three recognized classes of transcriptional activators (Mitchell and Tjian, 1989) are aligned using as a guide the six bulky hydrophobic residues of the minimal activation region of VP16. Remarkably, I observe great similarity between the acidic activation domain of VP16 and the glutamine-rich transcriptional activation domains of Spl (Courey and Tjian, 1988). For instance, all six bulky hydrophobic residues of VP16 can be aligned with bulky hydrophobic residues in the activation domains A and B of Spl. In addition, eight of the ten VP16 acidic amino acids can be directly aligned with glutamine residues in the Spl (A) domain. I have shown that uncharged, carbonyl-containing residues can functionally replace aspartate residues of the VP16 activation domain, with some decrease in transcriptional activation. VP16 is a much stronger activator than Spl; for instance, the IE upstream regulatory region has five Spl binding sites (Jones and Tjian, 1985), yet I barely detect IE-tk transcripts in absence of VP16. It is possible that the Spl activation domains are uncharged (and thus less active) analogs of the VP16 activation domain which nonetheless utilize similar hydrophobic contacts. The hydrophobic matches between VP16 and the other acidic or proline-rich activators shown in Figure 10 are not as striking as for Spl. However, shorter regions of similarity are observed. These short regions of similarity are characterized by bulky hydrophobic residues flanked by carbonyl-containing residues. When I began this project (1988) a critical role for hydrophobic contacts was not anticipated. The analysis of Zhu gg (1990), however, included a region of local hydrophobic maximum containing leucine, valine or isoleucine as one of nine 45 Figure 10. Conservation of bulky hydrophobic residues among activation domains. Alignments are based upon visual inspection aided by sequence comparison programs (Devereux M... 1984), but do not necessarily represent the most parsimonious alignment. Spl (Courey and Tjian, 1988) and CTF (Mermod et_al., 1989) are mammalian transcriptional activators of the glutamine-rich and proline-rich classes, respectively. GAL4 (Ma and Ptashne, 1987a), GCN4 (Hope and Struhl, 1886) and HAP4 (Forsburg and Guarente, 1989) are yeast acidic transcriptional activator proteins. B17 (Ma and Ptashne, 1987b) is encoded by a randomly-cloned ; 99h DNA fragment, which when fused to the GAL4 DNA binding domain activates transcription in yeast. AH (Gininger and Ptashne, 1988) is an artificial sequence designed to form an AAH. 46 QOZWZZ (3132401211112 oo>,amom< (DQaOiDUJQOaE-tHOi QHQDJZOQH>O Iqquqzqzqql a0. m> Zm D0100 [uHthNi—IAMBi—Il DZBOMMDZE—tm QOaODZEuEIJMDOiOl AEZI—‘lEi—Ii—JIEEAQ <>mmm9m>om QO>QE4ILDZUIOI «mommmommq mammogmmmm 40228 NZHr—I (1).-1&0 mmm mO BA 2d |:>:>:>.—itutu|:nom DOE-immbdmmz momHmquo OZE—tE—ttxJDNKCm oomm> AQHZ>AHU§< IEZHQKCBKch-I Figure 10 47 diagnostic components of acidic activation domains. Although Phe 442 does not lie within the region of hydrophobic maximum identified by Zhu et_al (1990) it is possible that the hydrophobic maximum identified by Zhu ml, (1990) may be a critical structural element of acidic activation domains. In all cases of activators aligned in Figure 10 the region of hydrophobic maximum identified by Zhu et_al, (1990) lies just C-terminal to the regions of activation domains similar to the region Phe 442 (for instance Leu 447 of VP16 was highlighted by Zhu e_t_al, 1990). The comparisons in Figure 10 predict that other bulky hydrophobic amino acids, such as leucine or tryptophan, may function at position 442 of VP16, and that other bulky hydrophobic residues may be important in transcriptional activation. Jeff Regier has characterized a number of VP16 mutants which substitute Phe 442 and flanking hydrophobic residues. Jeff has found that tryptophan, like tyrosine, can replace Phe 442 with retention of about one-third activity. On the other hand, leucine was dysfunctional at position 442. This observation strongly suggests that the residue at position 442 must be aromatic. Jeff has also altered two bulky hydrophobic residues flanking Phe 442, Leu 439 and Leu 444. Jeff has found that substitution of these leucine residues with other bulky hydrophobic residues such as valine and phenylalanine preserve activity, whereas substitutions of these residues with alanine or serine severely affect activation. Jeff’ s observations clearly support a critical role for hydrophobic residues in the activation by VP16. 48 Conclusions The results of studies described in this chapter suggest the following conclusions: 1) the VP16 activation domain activates transcription as a function of its net negative charge, but negative charge is not sufficient for activation, 2) the predicted AAH of the VP16 activation domain does not appear to be a critical structural component of the VP16 activation domain, 3) Phe 442 of the VP16 activation domain is critical to its function, and 4) there is some conservation of bulky hydrophobic residues among the sequences of various activation domains from different classes. These results suggest that both ionic and hydrophobic interactions are critical to the function of the VP16 activation domain. It seems reasonable to assume that this domain binds to a target protein which couples VP16 to some component of the basal transcriptional machinery. I suggest two possible models which may describe the structure of the VP16 activation domain and which may account for the importance of Phe 442 within that structure. In the first model I propose that Phe 442 is involved in hydrophobic contacts within the folded structure of the activation domain itself. Phe 442 and other hydrophobic residues interact hydrophobically to determine the internal structure of the domain, whereas hydrophilic and acidic residues seek the surface of the domain. Charged residues on the surface can then interact with the target of the VP16 activation domain by electrostatic interactions. If appropriate hydrophobic contacts are disrupted (as when Phe 442 is replaced) the internal structure of the activation domain is altered, which in turn alters the charged surface of the folded molecule which can no longer interact appropriately with a 49 charged surface of the presumed target. In the second model, I propose that Phe 442 may instead be involved in a hydrophobic contact with a target protein. For instance, charged residues could bring the VP16 activation domain into close proximity to its target (presumed to be positively charged) by long-range electrostatic attraction. Once the target is in close proximity, hydrophobic contact could prevail leading to intimate binding. Phe 442 for instance, might fit into an hydrophobic cleft within the structure of the target molecule. Such an interaction is not unprecedented. For example, the anticoagulant hirudin which binds to the exosite of thrombin possesses a critical phenylalanine residue which inserts itself into a hydrophobic cleft of the thrombin molecule upon binding (Rydel et al., 1990). In fact, the amino acid sequence of the hirudin polypeptide flanking its critical phenylalanine residue is similar to the amino acid sequences flanking Phe 442 of VP16. CHAPTER III ANALYSIS OF THE VP16 ACTIVATION DOMAIN BY SPECTROSCOPIC AND BIOPHYSICAL NIETHODS Introduction In the preceding chapter I described a mutational analysis aimed at understanding the critical structural elements of the VP16 activation domain. That analysis suggested that net charge and some structural component were critical to the function of the VP16 activation domain. Data emerging from mutagenesis further demonstrated that position 442 was important to the function (and presumably structure) of the activation domain. Unfortunately, mutagenesis cannot be used to define protein structure m. Thus, I initiated a biophysical characterization aimed at directly defining the structural elements of the VP16 activation domain. This characterization has utilized five methods of structural analysis: ultraviolet (UV) spectroscopy, Fourier-transform infrared (FTIR) spectroscopy, circular dichroism (CD) spectroscopy, native gel electrophoresis, and differential scanning calorimetry (DSC). GAL4-VP16 Fusion Proteins Many methods of biophysical analysis require milligram quantities of the sample polypeptide to be characterized. Unfortunately, it would be very difficult to 50 51 adequately purify the amounts of virion-borne VP16 necessary for such analysis (D. Lewis, S. Lui, R. Pichyangkura, J. Regier and A. Cress, unpublished experiments). Biophysical analysis of the VP16 activation domain in the context of the virion-born protein is further complicated in that the activation domain represents less than one-fifth of the entire VP16 polypeptide (the secondary structure of which is totally undefined) making it very difficult to define structural elements specific to the activation domain. To circumvent these problems I have utilized GAL4-VP16 fusion proteins for structural analysis. GAL4-VP16 fusion proteins are composed of the C-terminal activation domain of VP16 (or derivatives thereof) fused to the yeast GAL4 DNA-binding domain (Sadowski gg, 1988). There are a number of advantages to using these proteins for structural analysis of the VP16 activation domain: 1) milligram quantities of these proteins can be prepared from Egli at greater than 90% purity, 2) the GAL4 DNA-binding domain has been extensively studied and its structural elements are well defined, 3) VP16 activation domain peptides containing only a few non-VP16 residues at their N -termini could be separated from fusion proteins by proteolysis with trypsin, 4) the DNA-binding activity of the GAL4 component of the fusion protein offered a simple assay to monitor for the presence of the fusion protein, and 5) GAL4-VP16 fusion proteins can activate transcription in i_n 1mg assays (Chasman mil, 1988; Berger gal, 1990), thus offering a way to determine the affect of various treatments upon activity (there is no such in vitro assay for transcriptional activation by full-length VP16). 52 Composition of GAL4~VP16 Ersion Proteins The first bar of Figure 1 represents the GAL4 DNA-binding domain (GAL4 1-147) which is composed of the 147 N-terminal amino acids of the yeast GAL4 protein. CD analysis suggests that the GAL4 1—147 protein is approximately 40% helical, 40% coil and less than 20% beta-sheet (Pan and Coleman, 1989). GAL4 1-147 has three important subregions, a Zn-binding region (Pan and Coleman, 1989), a DNA-specificity region (Corton and Johnston, 1989) and a dimerization region (Carey et al. , 1990a). The Zn-binding region forms a Zn(II)2Cys6 binuclear cluster as determined using 113Cd NMR (Pan and Coleman, 1990) and the complete secondary structure of the Zn-binding domain (residues 7-49) has been determined using 1H NMR (Gadhavi e_tg, 1990). The entire three-dimensional structure of GAL4 1-147 should be forthcoming in the near future. The second bar in Figure 1 represents GAL4-VP16 in which the 78 C-terminal amino acids of VP16 (the full-length activation domain) are fused to the GAL4 1-147 polypeptide (Sadowski e_t_g, 1988). This fusion protein is a powerful transcriptional activator from promoters containing GAL4 binding sites in mammalian cells in vivo (Sadowski g, 1988) and in yeast and mammalian cell extracts in_vm (Chasman et al., 1989; Carey e_t_al, 1990a) demonstrating that activation domain structural elements are preserved in these fusion proteins. Recent mutational studies by Jeff Regier (unpublished data) indicate that the full—length activation domain appears to encode two activation regions, termed regions I and II. The third bar in Figure 1 shows GAL4-del456 in which only activation region I is fused to GAL4 1-147. In the previous chapter, region I was referred to as the minimal (or truncated) activation 53 Figure 1. GAL4 1-147 and GAL4-VP16 fusion proteins. GAL4 1-147 is the DNA-binding domain of GAL4 (stippled). GAL4-VP16 is GAL4 1-147 fused in—frame to the full—length VP16 activation domain (V P16 amino acids 413-490). GAL4—del456 is GAL4 1-147 fused in-frame to the region I activation domain (VP16 amino acids 411-456). The various mutants described in this work are derivatives of GAL4—del456, as indicated above. The GAL4-VP16 fusion protein contains a seven amino acid linker (PEFPGIW) between GAL4 Ser 147 and VP16 Ala 413. The GAL4-del456 fusion protein contains a two amino acid linker (GS) between GAL4 Ser 147 and VP16 Ser 411. 54 mmnm 20-3420 9va $.35 mm mm «3. E35 Ti «3. $-35 N3. E420 gag / o _ 28mm 2 838-420 1 EVE: mm N W. :8 8&3 8a. 8... «5. E ............................................. o ___8_am Eosmm z oE>-3 O" U) A 3 O' m :3 O (D 0.08 " B 0.04 - C D . r ._L 1 r I E F5 I I 257 271 > 285 299 313 Wavelength (nm) Figure 4 69 difference spectrum of 0.4 mM GAL4-FY 442 with 0.4 mM free tyrosine as shown in Figure 4. The extremely weak difference spectrum of Tyr 442 relative to an equal concentration of free tyrosine strongly suggests, in agreement with the previous titration data, that Tyr 442 is buried. I extended UV analysis to the more complex fusion protein containing the full- length activation domain, GAL4—VP16. GAL4-VP16 has three polar aromatic residues in addition to GAL4 Trp 36 and Tyr 40: a tryptophan residue which is encoded by the cloning linker between the GAL4 and VP16 domains (Trp 154), a natural tyrosine at VP16 position 465 (Tyr 465) and another natural tyrosine residue at VP16 position 488 (Tyr 488) just two amino acids from the VP16 C-terminus. Figure 5 compares the 20% ethylene glycol difference spectra of 0.4 mM GAL4-VP16 with the spectrum of a mixture of 0.4 mM tryptophan and 0.8 mM tyrosine which would be the expected spectrum of 0.4 mM GAL4-VP16 if both tyrosines and the tryptophan were exposed. The similarity of the spectra compared in Figure 5 suggest that Trp 154, Tyr 465 and Tyr 488 are exposed to solvent. The fact that the real GAL4-VP16 difference spectrum is broader than the spectrum of equivalent concentrations of the free amino acids may reflect interactions with other groups within the protein. It is not surprising that Trp 154 is solvent exposed, since it is encoded by an artificial linker which is not likely to form a specific structure that would occlude solvent from a specific residue. Similarly it is not unexpected that Tyr 488 is exposed since it lies so close to the C-terminus of the protein. Tyr 465 lies within the activating region H of the full-length VP16 activation domain proximal to a phenylalanine residue thoughtto be homologous to Phe 442 of activating region I. 70 Figure 5. Difference spectrum of GAL4-VP16. A. 0.4 mM GAL4-VP16, B. 2:1 mixture of 0.8 mM tyrosine and 0.4 mM tryptophan (stippled). Method. These spectra were obtained as described in the legend of Figure 4. >cmoqcmzoo 0.16- 71 _ _ 2 8 4| 0 0 0 508...... 0 CI l0 “8:0"00'09 0' Chilleh .0 I‘I‘COIIO fiODe.‘ 0.04 - 313 299 (nm) Figure 5 285 Wavelength 271 257 72 Knowing that Tyr 465 is exposed to solvent will aid in future studies of the structure of the full-length VP16 activation domain. The results of UV analysis suggest strongly that Tyr 442 is buried within the structure of GAL4-FY 442, and from this I infer that Phe 442 is buried. There are at least three problems with applying conclusions from this data to the structure of the VP16 activation domain. First, I do not know that Tyr 442 being buried is functionally relevant. For instance, Tyr 442 may be buried in the context of the GAL4 fusion protein by a fairly nonspecific hydrophobic interaction with the GAL4 DNA-binding domain. One observation which argues against this criticism is that the GAL4-FY 442 protein activates transcription m (S. Berger and A. Cress, unpublished results) with approximately the same relative activity as pMSVP16 FY 442 activates transcription ill—Vllg (the previous chapter). Second, I cannot be certain that Tyr 442 accurately reflects the environment of Phe 442; however, it is very likely that it does. For instance, if Phe 442 were solvent exposed it would be unlikely that Tyr 442 (which is more polar than Phe 442) would seek a more nonpolar environment in the same context. A third problem with drawing conclusions from this UV data is a technical one, in that I do not know precisely the concentrations of proteins used in these experiments. The protein concentrations used in these experiments are estimated primarily from a commercially available protein assay (described in the legend of Figure 3), since attempts to accurately determine protein concentrations using quantitative amino acid analysis have been unsuccessful for an unknown reason (J. Leykam, unpublished experiments). Although I do not expect a large inaccuracy 73 based upon this protein assay, an independent method of accurately determining the protein concentration would be preferred and will be obtained prior to publication of these results. The finding that Phe 442 is likely buried is clearly in support of the model proposing that Phe 442 is critical in the folding of the VP16 activation domain upon itself. VP16 derivatives with substitutions such as proline, alanine and serine at position 442 probably can not fold properly. Substitution of Phe 442 with tyrosine allows folding of the domain into a structure with diminished activity. Diminished activity is likely due to the fact that polar tyrosine doesn’t fit as well as phenylalanine into the natural hydrophobic environment of Phe 442. Fourier-Transform Infrared Spectroscopy UV analysis suggests that the VP16 activation domain structure insulates Phe 442 from solvent, but doesn’t give information regarding secondary or tertiary structure. Infrared absorption spectroscopy, on the other hand, has been used to study the secondary structure of polypeptides. Protein structure analysis by FTIR is generally done in D20 solution, instead of water, since D20 is relatively transparent in the region of interest. In deuterium oxide solution the most useful absorption band for secondary structure studies is the amide I band (1620 to 1690 cm‘l) which corresponds to the C =0 stretching vibration of peptide groups. The C =0 stretch is sensitive to secondary conformation, thus different secondary substructures give rise to slightly different absorption maximum. Although the various C =0 stretching Components of typical proteins have inherently broad line shapes, the spectral 74 contributions of various substructures can often be resolved using mathematical techniques. For instance, Fourier self-deconvolved (Kauppinen gal, 1981) FTIR spectroscopy has been used to examine the structures of 21 globular proteins (Byler and Suzi, 1986). In that study, the amide I spectra of various proteins consisted of six to nine components which were empirically assigned to various elements of secondary structure such as alpha-helix, beta-sheets, coil regions and turns. The power of FTIR spectroscopy to resolve between alpha-helical and beta-sheet structural components of proteins in solution is particularly good, as demonstrated in Figure 6 which presents the deconvolved FTIR spectra of concanavalin A and hemoglobin. Hemoglobin is approximately 80% alpha-helical and is without beta-structure or turns. Accordingly, its IR spectrum shows a single large peak centered at 1651 cm‘1 which corresponds to the single helix-specific C=O stretch identified by Byler and Suzi (1986). As expected, hemoglobin does not have strong absorbances corresponding to beta-sheet or turn structures. Concanavalin A, on the other hand, is about 60% beta-sheet, thus its IR spectrum shows a principle peak at 1635 cm'1 and a weaker peak at 1623 cm‘1 corresponding to two beta-sheet specific C=O stretches identified by Byler and Suzi (1986). I have applied FTHI analysis in an attempt to measure the elements of secondary structure within the VP16 activation domain. Figure 7 shows the amide I region of the FTIR spectra of GAL4 1-147, GAL4-VP16, GAL4-del456, GAL4-PP 442 and the VP16 activation domains purified from GAL4-VP16 and GAL4-del456. Each of these polypeptides have FTIR maxima in the spectral region corresponding to alpha-helical (1650-1654 cm'l) and coil structures (1644-1648 cm‘l), but do not have 75 Figure 6. FTIR spectra of model proteins. A. Hemoglobin, B. Concanavalin A. The absorbance scale is in arbitrary units. Method. Lyophilized proteins obtained from Sigma were dissolved in 200 mM NaCl in deuterium oxide solution (50 mg/ml) and allowed to equilibrate for 24 hrs. FTIR spectra were obtained and analyzed using a Nicolet 7199 FTIR instrument (courtesy of Dr. R. Hollingsworth). Three thousand co-added interferograms were obtained at 2 cm'1 nominal resolution. The sample volume and pathlength were 0.05 ml and 0.075 mm, respectively. Deconvolution parameters were VFO=20 cm‘1 and VF1=2.0 and Lorentzian line-shape was assumed. A v A ‘ Absorbance V 76 L 1710.3 3705.2 {592.0 {375.9 {565.0 3652.7 Wavenu'm ber (cm" ) Figure6 £539.: A 1626.4 5613.3 i600.2 77 Figure 7. FTIR spectra of GAL4-VP16 proteins A. Full-length (90 amino acids) VP16 activation domain, B. truncated VP16 activation domain, C. GAL4-VP16, D. GAL4-FF 442, E. GAL4-del456, F. GAL4 1-147. The absorbance scale is in arbitrary units. Method. The activation domain peptides corresponding to the full- length and truncated activation domains were separated from the fusion proteins GAL4-VP16 and GAL4-del456 by tryptic cleavage and purified as described in the text. Proteins were prepared for FTIR by dialysis against 5 mM HEPES (pH 7.5) and 0.2 M NaCl. Samples were then lyophilized, resuspended in D20 and allowed to equilibrate for 24-48 hr. FTIR spectra were then obtained as described in the legend of Figure 6. Absorbance 6710.3 78 A A _‘ ‘ A L ‘ A ‘ ‘ 3705.2 {092.0 1070.0 1000.0 3002.7 1039.0 1020.4 10:3.3 1000.2 Wavenumber'lcm'1 ) Figure 7 79 major peaks in the spectral region corresponding to beta-sheet structure. These data suggest that these proteins are composed predominantly of a mixture of helix and coil substructures. This observation is consistent with published CD data which estimates that GAL4 1-147 is 40% alpha-helix, 40% coil and 20% beta-sheet (Pan and Coleman, 1989). The IR spectra of the purified activation domains are notably broader than the spectra of the proteins containing GAL4. While the maxima of helix and coil IR spectra are very close, the IR spectrum of coil structure is characteristically broader than that of alpha helix (Figure 4 of Byler and Suzi, 1986). The observation that the pure activation domain peptides have broad IR spectra relative to GAL4 1-147 suggests that they possess a higher percentage of coil structure. The fusion proteins GAL4-VP16, GAL4-del456, GAL4-FP 442, and the purified full-length and truncated activation domains appear to be predominantly a mixture of helix and coil structures. Unfortunately, the IR resonances corresponding to these substructures are highly overlapping making it very difficult to determine relative contributions of these two substructures to the overall structure. Thus, I have chosen to rely on a complementary method of structural analysis, circular dichroism (CD), which is much more sensitive to differences in helix and coil substructures for quantitative estimates (see the next section). I have, however, used FTR qualitatively to study the affects of various solvents upon the structures of GAL4 1-147 and the GAL4-VP16 fusion protein. One important finding of these comparisons was the great sensitivity of the GAL4 1-147 DNA-binding domain to conditions of low salt and low pH. At pH 7 .5, the GAL4 80 1-147 protein is insoluble below 0.2 M NaCl, whereas the GAL4-VP16 fusion is soluble down to 0.1 NaCl. Apparently the highly charged activation domain improves the solubility of the fusion protein, perhaps by charge repulsion between the negatively charged domains. Neither GAL4 1-147 nor GAL4-VP16 show significant changes in the their FTIR spectra in the range of NaCl concentration from 0.2 M to 1.0 M, however the GAL4-VP16 protein does show an overall shift toward lower wavenumbers at 0.1 M N aCl (where GAL4 1-147 precipitates), probably reflecting the structural changes leading to the precipitation of GAL4 1-147 in low salt. At 0.2 M NaCl both GAL4 1-147 and GAL4-VP16 precipitate at pH less than 7.0, although no structural changes are evident in the pH range from 7 to 9. Negatively charged residues are important in the cation-binding domains of proteins such as calmodulin and troponin C. Upon divalent-cation binding both calmodan and troponin C undergo conformation rearrangements which are evident using FTIR spectroscopy (Trewhella gg, 1989). Although the VP16 activation domain does not have striking sequence homology to the cation-binding domains of calmodulin or troponin C it was considered a possibility that the VP16 activation domain might interact with such divalent cations. Thus, the effects of the cations K+, Ca++, Mg++, Mn++, Fe++, and Zn++ upon the structure of GAL4-VP16 were evaluated using FTIR. GAL4-VP16 (0.2 mM) was dialyzed against 2 mM EDTA, 0.2 M NaCl, and 5 mM HEPES buffer (pH 7.5) to remove bound cations, then against the same buffer without EDTA. The sample was lyophilized and resuspended in D20. Specific salt solutions (KCl, CaClz, MgClz, MnClZ, FeC12 and ZnC12 in D20) were then added and FTIR spectra obtained after 24-48 hours. Surprisingly, 2 mM FeClz or 2 mM 81 ZnC12 precipitated 0.2 mM GAL4-B58 (equimolar concentrations did not). This effect was clearly specific to the GAL4 DNA-binding domain, since GAL4 1-147 was also precipitated by equimolar FeClz, ZnC12 or 113CdC12. This phenomenon is reflected in the protocols of investigators who have previously studied the GAL4 Zn-binding domain (Pan and Coleman, 1989 and 1990), but an adequate explanation for its mechanism is not available. The cations K+, Ca++, Mg”, and Mn++ (in ten-fold excess) did not have any strong effects on the FTHI spectra of GAL4-B58 (0.2 mM) nor did equimolar concentrations of FeC12 or ZnClz. These result suggest that the VP16 activation domain structure is not affected by interaction with cations and probably suggests that the VP16 activation domain is not a specific cation-binding protein. Circular Dichroism FTIR analysis is limited in its ability to resolve alpha-helix and coil structures, however the VP16 activation domain appears to be composed of a mixture of these two substructures. To estimate the relative contributions of helix and coil to the total structure of the VP16 activation domain, circular dichroism (CD) was used. Because a CD instrument is not available on campus, I collaborated with Dr. B. Krueger and M. Prairie at the Upjohn company in Kalamazoo, M1 to obtain a CD spectrum of the full-length VP16 activation domain. Figure 8 shows the CD spectrum of the full- length VP16 activation domain as obtained in 10 mM sodium phosphate buffer (pH 7.2) containing 0.2 M NaCl. This CD spectrum clearly reveals that the activation domain has little alpha-helical content as indicated by its relatively low negative 82 MOLAR ELLIPTICITY X ID I _‘a—l ’12 I I I I I I I I I 175 195 215 235 255 275 WAVELENGTH (NM) Figure 8. CD analysis of the full-length VP16 activation domain. Method. The full-length VP16 activation was prepared for CD analysis by dialysis against 10 mM sodium phosphate buffer (pH 7 .2) and 0.2 M NaCl. The CD spectrum above is the average of 16 scans obtained in a 0.1 mm cell at room temperature using a Jasco 600C CD spectropolarimeter. 83 ellipticity at 222 nm. The CD spectrum does however suggest considerable coil structure is present. Unfortunately, the CD spectrum of the full—length tail did not yield acceptable results using secondary structure calculations (Compton and Johnson, 1986; Manavalan and Johnson, 1987). This observation suggests that the unlmown structure is not well represented in the data base of proteins used (Dr. B. Krueger, personal communication). It is possible that the structure observed for the purified activation domain peptide is not the natural structure of the domain, but that the activation domain needs to be attached to the GAL4 DNA-binding domain (or to the specificity domain of VP16) to form an active structure. I do not expect that this is the case, however, since the isolated VP16 activation domain (with only an added nuclear localization signal) has been expressed in Hela cells and shown to inhibit transcription from reporter promoters (Tasset gg, 1990) suggesting that no other sequences are required for it to form an active structure. Native Gel Electrophoresis UV analysis suggests that Phe 442 is buried within the structure of GAL4—VP16, thus Phe 442 may be critical in maintaining a particular protein conformation. If Phe 442 is maintaining such a conformation then it is likely that inactivating substitutions at position 442 might disrupt that structure. Furthermore, it might be possible to detect this change of conformation using native-gel electrophoresis, since the mobility of a native protein within an electric field is 84 inversely proportional to its frictional coefficient which is in turn a function of its overall conformation. I tested the possibility that substitutions at Phe 442 might alter the mobility of GAL4-VP16 fusion proteins in the context of a DNA-binding/ mobility shift assay (Fried and Crothers, 1981). In this assay, an oligonucleotide composing the GAL4-binding site was incubated in the presence of various GAL4-VP16 fusion proteins. The complexes were then electrophoresed on a nondenaturing gel and the relative mobility of the complexes compared (see Figure 9). The first lane of Figure 9 shows the mobility of the GAL4 1-147:DNA complex which has the slowest mobility of all of the GAL4-VPl6zDNA complexes despite being the smallest protein. The second lane of Figure 9 shows the mobility of GAL4-AH which adds a fifteen amino acid peptide designed to be an AAH to the GAL4 DNA-binding domain (Gininger and Ptashne, 1988). The increased mobility of this proteinzDNA complex relative to the GAL4:DN A complex is apparently due to the three negative charges added by the AH peptide. The next seven lanes of Figure 9 compare the mobilities of GAL4-del456 with six mutant derivatives. These lanes reveal that mutations of Phe 442 do not affect the mobility of GAL4-VP16 fusion proteinzDNA complexes in this system, since GAL4-del456, -FP 442, -FA 442, and -FY 442:DNA complexes all have indistinguishable mobilities. Ianes labeled DN 2578 and DN 2579 reveal the importance of negative charge in this system since removal of four negative charges considerably slows the mobility of these derivatives relative to GAL4-del45 6. The final lane of Figure 9 shows the relative mobility of GAL4-VP16. Despite having considerably more negative charge than GAL4-del456 derivatives the mobility of 85 'l I I to <2- " r~ LO SI. “3, 3 <2- 3 ‘1' N N :3 1- I d) D. v n. < 0- Z Z ‘3: T? u. it. It < o o >, V V q. l “I. l I I . V _. _n —I 3 _r 3 3 3. 3 _. < < < < < < < < < < (5 (5 (5 (5 (5 (5 (5 <5 (5 (5 Figure 9. Native gel electrophoresis of GAL4-VP16 proteins. The proteins assayed are indicated by name below the lanes. Method. Binding assays (20 pl total volume) contained 20 mM HEPES (pH 7.5), 50 mM KCl, 5 mM MgC12, 10 pM ZnClz, 6% glycerol, 200 pg/ml BSA, 2 ng of 3ZZP—labeled double-stranded oligonucleotide (5’—CGGACGAGAGTCTTCCG-3’) and 20 ng of each protein. Binding was allowed to take place for 15 min at room temperature, then 4 pl of 5X loading dye (0.25 g/ml ficoll with 0.01% bromophenol blue) was added and the sample loaded onto a pre-electrophoresed 4.5% (w/v) polyacrylamide gel containing 90 mM TRIS base, 90 mM boric acid, 2 mM EDTA, and 1% glycerol. Electrophoresis was carried out in gel buffer (minus glycerol) at 15 Volts/cm until the bromophenol blue had entered the lower buffer tank. The gel was fixed for 30 minutes in 20% (v/v) methanol and 10% (v/v) acetic acid, dried onto Whatman paper and an autoradiograph obtained by exposure to X-ray film. 86 the GAL4-VP16 complex is only slightly greater suggesting that the GAL4-V1316 protein possesses a measurably greater frictional coefficient than its deleted counterpart. The observation that mutations at Phe 442 didn’t affect the mobility of GAL4-del456:DNA complexes suggests that the presumed structural differences are too small to detect in the context of this assay, but are otherwise inconclusive. Because of the net negative charge of GAL4-VP16 fusion proteins it was also possible to compare their mobilities in the absence of DNA using the same TRIS-Borate (pH 8.3) buffer system used to compare mobilities of proteinzDNA complexes. Unfortunately, the mobilities of GAL4-del456 derivatives having substitutions at Phe 442 were also indistinguishable in absence of DNA ligand. It is possible that structural changes are present, but that the relative mobilities of these proteins, at the relatively high pH of the buffer system used, are simply dominated by their negative charge. Unfortunately, GAL4 fusion proteins are insoluble below neutral pH, thus low pH electrophoresis buffers which might maximize the contribution of protein conformation toward electrophoretic mobility (while eliminating the dominant contribution of so many acidic residues) could not be used. Differential Scanning Calorimetry UV analysis suggests that the VP16 activation domain folds into a specific structure which insulates Phe 442 from solvent. I presume that mutations at Phe 442 in some way disrupt this structure, since they inactivate the domain, however, I was not able to detect any differences in structure using native gel electrophoresis. This 87 observation may indicate that the overall shapes of inactive VP16 derivatives are not very different than wildtype. However, the inactive derivatives might be less stable thermodynamically, thus I attempted to measure the thermodynamic stabilities of various activation domain derivatives using differential scanning calorimetry (DS C) (reviewed by Krishnan and Brandts, 1978; Donovan, 1984). Before DSC could be used to explore differences in activation domain stability among various derivatives it was necessary to first determine if the VP16 activation domain unfolded in a cooperative fashion such that it would yield a measurable denaturation enthalpy. Figure 10 shows DSC scans of GAL4 1-147 and GAL4-VP16. GAL4 1-147 did not tolerate heating. At temperatures above 47 °C GAL4 1-147 began to gradually precipitate, thus the DSC scan of GAL4 1-147 is not meaningful. Varying solvent conditions (NaCl up to l M; and pH up to 9) did slightly raise the temperature at which GAL 1-147 began to precipitate, but did not allow me to obtain DSC data of GAL4 1-147 in solution. The GAL4-VP16 fusion protein on the other hand remained soluble up to 100 °C; again, as in the case of low salt, the VP16 activation domain keeps the GAL4 DNA-binding domain in solution. Unfortunately GAL4—VP16 did not yield any obvious peaks in heat capacity corresponding to the unfolding of individual protein domains. Instead, a gradual increase in heat capacity was ob served over the entire temperature range monitored. This observation probably indicates that the GAL4-VP16 protein is not composed of one or two major cooperatively folded units, but rather is composed of several small units that unfold independently with overlapping melting temperatures. These hypothetical units could 88 Figure 10. DSC scans of GAL4 1-147 and GAL4-VP16. A. 2 mg/ml GAL4 1-147 B. 2 mg/ml GAL4-VP16 Method. Proteins were prepared for DSC by dialysis against 20 mM HEPES (pH 7.5), 0.1 M NaCl, 20 mM monothioglycerol, and 10 pM zinc acetate. Samples and reference buffers were deaerated with stirring under vacuum. DSC experiments were conducted at a scan rate of 90 °C/hr in a Microcal MC-2 scanning calorimeter (courtesy of Dr. J. Wilson). meal/degree K meal/degree K 89 ll 1 l l l l ' l 20 30 4O 50 60 70 80 90 100 Temperature (degrees C) l— l L l l l 1 LI 20 30 40 50 60 70 80 90 100 Temperature (degrees Cl Figure 10 90 correspond to the Zn—binding, DNA-specificity, and dimerization domains of GAL4 1-147 and the two activating regions of the VP16 activation domain. Conclusions The work in this chapter represents the first biophysical characterization of a transcriptional activation domain. The results of this chapter suggest two conclusions. First, Phe 442 of the VP16 activation domain is probably buried within the tertiary structure of the protein, whereas Tyr 465 and Tyr 488 are exposed to solvent. This observation suggests that the critical functional role of Phe 442 discovered in the previous chapter is to maintain the internal structure of the domain through hydrophobic interactions. The second conclusion that can be made from this chapter is that the VP16 activation domain contains very little alpha—helix. This observation is consistent with the results obtained using mutagenesis with regard to the functional importance of predicted alpha-helix. The data presented in this and the preceding chapter support a model in which internal hydrophobic interactions involving Phe 442 are important in maintaining the structure of the VP16 activation domain. More detailed information regarding the structure of the VP16 activation domain will require that the structure of the domain be determined using methods such as 1H NMR or X—ray crystallography. CHAPTER IV DISCUSSION Introduction An important question, related to the central topic of this dissertation, is how the VP16 activation domain works. When I began this project the mechanism of activation by VP16, or any other activator, was primarily a matter of speculation. Since that time however, the VP16 activation domain has been extensively studied and has became somewhat of a paradigm for transcriptional activators. During the course of my investigation I generated a number of VP16 derivatives which have become important reagents in addressing the identity of the VP16 target. In this section I will summarize several investigations which have utilized several of my VP16 activation domain derivatives toward identifying the activation domain target(s). Mechanism of Action of the VP16 Activation Domain. It is becoming clear that the VP16 activation domain may utilize a number of mechanisms to activate transcription. For instance, the VP16 activation domain can clearly activate transcription by the mechanism of antirepression (Croston _e_t_g, 1991). The VP16 activation domain may also interact with the basal transcription factor TFHD. Stringer, Ingles and Greenblatt (1990) performed the first experiments that indicated that the VP16 activation domain might interact directly with TFHD. In 91 92 their experiments, they utilized the VP16 activation domain as a column ligand in affinity chromatography. They found that the VP16 affinity column selectively retained human TFHD and purified recombinant yeast TFHD. This result was consistent with the hypothesis that TFHD might be a target of the VP16 activation domain, however, it was open to the criticism that the interaction between the two proteins could be relatively nonspecific due to the net negative charge of the VP16 activation domain and the net positive charge of the TFHD protein [(pKa 9.5 (Hahn Qt g, 1989)]. Substitutions of Phe 442 of the minimal VP16 activation domain with alanine, serine and proline abolish activation without affecting negative charge. If the interaction of TFHD with VP16 was simply the result of charge attraction then these inactivating substitutions should have no effect on the mutual affinity between TFHD and VP16. To test the specificity of the interaction between TFHD and the VP16 activation domain, I sent the Ingles group a collection of VP16 activation domain derivatives. In this collaboration (Ingles girl, in press) we measured the relative affinities of the VP16 derivatives del456, FP 442, FS 442, FA 442, FY 442, DN 2789 and DN 2579 for yeast recombinant TFHD. The deleted version of the VP16 activation domain del45 6 had a ten-fold lower affinity for TFHD than did full-length VP16, despite the fact that it 10Ses only 50% activity of the full-length activation domain. However, when VP16 derivatives having substitutions at Phe 442 were compared in the context of the deleted version, their affinities for TFHD correlated perfectly with their relative activities. The inactive VP16 derivatives FP, FS and FA 442 had little measurable affinity for TFHD, 93 whereas the partially active derivative FY 442 had an affinity for TFIID partially reduced from that of del456. This observation clearly rules out the possibility that the affinity of TFHD to VP16 is purely a nonspecific ionic attraction. Surprisingly, the derivatives DN 27 89 and 25 79 which possess 20 and 60 % relative activity, respectively, both have lower affinities for TFHD than FY 442 (despite the fact that DN 2579 is more active in transcription). The observations that affinity for TFHD does not correlate perfectly with measured relative activities of VP16 derivatives might suggest that TFHD binding may not be the sole mechanism of activation. Alternatively, the imperfect correlation between affinity for TFHD and activation may simply reflect that net negative charge is important in attracting and binding TFIID, but that hydrophobic interactions are critical in the activation phenomenon (which might involve a change in conformation). Thus, the mutant DN 2579 may activate transcription better than FY 442 because DN 2579 better preserves hydrophobic interactions critical in activation, whereas FY 442 binds TFHD better because it preserves negative charges critical in attracting TFHD. Lin and Green (1991) have recently discovered evidence that the VP16 activation domain specifically interacts with yet another basal transcription factor, TFHB. In this work, Lin and Green used the VP16 activation domain as a chromatography affinity matrix in a manner similar to Stringer _elgl (1990). They passed Hela cell nuclear extracts over the VP16 affinity column and observed binding of TFHD to the column. However, Lin and Green went on to demonstrate that TFHD is not the only basal transcription factor bound to the VP16 column. TFHB was also bound and appeared to be bound with greater affinity than is TFIID (with 94 respect to the concentration of KCl used as eluant). Stringer e1; gl, (1990) did not eliminate the possibility that TFHB was bound to their columns. Lin and Green recognized the possibility that the affinity of TFHB to the VP16 column could be due to non-specific charge interactions. Thus, they requested the VP16 activation domain derivatives del45 6 and FP 442 which they used as control affinity matrices. They found that TFHB bound to the del45 6 protein and not to the FP 442 derivative under identical conditions, thus the affinity of TFHB (like that of TFHD, described above) for VP16 was specific to transcriptionally active versions of VP16 suggesting that the ob served interaction is biologically relevant. Activation of transcription by VP16 may not be limited to antirepression and recruitment of basal factors. In a collaboration with S. Berger and L. Guarente, we (A. Cress, S. Triezenberg and I) sought to identify the VP16 target using yeast i_n 21m transcription reactions and GAL4-VP16 fusion proteins (Berger gal, 1990). In this collaboration we compared the effects of GAL4-VP16 on three promoter templates: one having only a TATA box and two having upstream activating sequences, GAL4 UAS or dAsz UAS. The basal level of transcription in these experiments was defined as the level of transcription obtained from the promoter containing only a TATA box. The GAL4 UAS promoter gave a basal level of transcription in the yeast extracts in the absence of exogenous GAL4-VP16, whereas the dAsz UAS template was activated to a level ten-fold basal (probably due to the presence of a dAsz activating protein in the yeast extract). When GAL4-VP16 was added to the GAL4 UAS template transcription was activated roughly 100-fold over the basal level. However, GAL4-VP16 abolished transcription from templates 95 containing basal and dA:dT promoters. Mutations within the VP16 activation domain, such as FP 442, which abolished activation by VP16 also abolished its ability to inhibit, thus the mechanism of this inhibition presumably involves the ability of VP16 to bind to some transcription factors (or factors) and titrate them away from heterologous promoters. Addition of an oligonucleotide containing a GAL4-binding site which would prevent GAL4-VP16 from binding nonspecifically to DNA, however, was found to relieve inhibition of basal transcription of both the basal template and the dA:dT template, but could not restore transcription to the activated level. This observation suggested that GAL4-oligozGAL4-VP16 must be able to titrate some factor involved in activated transcription, but not required for basal transcription. An apparently identical activity which partially copurifies with TFIID (presumably due to their association) has been identified by other workers (Kelleher e_t g, 1990; Hoey et al., 1990; Pugh and Tjian, 1990; Peterson et al., 1990). The list of factors through which VP16 has the potential to mediate its effect is long. Is it possible that VP16 interacts with many factors to activate transcription? It has been noted that transcriptional activators work synergistically when multiple regulatory elements are present (Lin et al., 1988; Carey fl, 1990b). In experiments which have studied synergistic action by acidic activation domains it has been estimated that up to ten molecules of activator can synergize to activate transcription (Carey et_al., 1990b). If the mechanism of activation required interaction with a single factor then this would suggest that up to ten activation domains could contact the target simultaneously. As pointed out by Ptashne and Gann (1990), it is very difficult to imagine that ten domains could simultaneously contact a 96 1 single target protein such as TFHD, for instance, since its molecular weight is only 38 kD and TFHD also appears to interact in some manner with other basal factors, TFHB and TFHA. Perhaps a better explanation would be that the acidic activation domains indeed can recruit several components of the basal transcriptional machinery, in addition to its ability to relieve repression by histones. This would explain how multiple activators can function synergistically at a single promoter and would be consistent with the observation that strong activator proteins, such as the full-length VP16 activation domain, often have multiple activating regions (J. Regier and S. Triezenberg, unpublished results). The final answer to the question of how the VP16 activation domain functions will clearly require further study. As clones corresponding to the various transcription factors become available for use in in vitro transcription reactions they will undoubtably provide new insights into the mechanisms of transcriptional activation by VP16. Furthermore, the VP16 mutants described in this dissertation will most likely be valuable as reagents in the characterization of the activation mechanism. Critical Structural Elements of the VP16 Activation Domain The hypothesis that acidic transcriptional activation domains function as AAHs (Gininger and Ptashne, 1987) has stood untested for a number of years. To my knowledge, my work is the first in depth test of that hypothesis and the first structural characterization of a eukaryotic transcriptional activation domain. In this work, I believe that I have convincingly demonstrated that the AAH model is incorrect. 97 Unfortunately, I can not suggest a simple model to replace the AAH. In other words, I have not been able to identify the "leucine zipper" or " zinc finger" or any other telling anatomical feature of acidic activation domains, although I do observe some conservation of bulky hydrophobic residues among transcriptional activation domains. I propose two alternative models which are somewhat more general than the AAH, but which take into account the measured structural elements of the activation domain and seek to explain the critical role observed for hydrophobic residues. One model proposes that Phe 442 may make direct contact with the VP16 activation domain target. Phe 442 for instance, might fit into an hydrophobic cleft within the structure of the target molecule. This model is made particularly attractive when the similarities between the VP16 activation domain and the acidic tail of the anticoagulant hirudin are considered. The hirudin acidic tail binds to the exosite of thrombin and possesses a critical phenylalanine residue which inserts itself into a hydrophobic cleft of the thrombin molecule upon binding (Rydel e_t_al, 1990). The possibility that hirudin and the VP16 activation domain are structurally similar is appealing because the crystal structure of hirudin in complex with thrombin has been solved (Rydel §t_a_l,, 1990). The relevant sequences of hirudin and the VP16 activation domain compare as follows: VP16 -Asp-Ala-Leu-Asp-Asp-m-Asp-Leu-Asp-Met-LeuM7 I-Iirudin -His-Asn-Asn-Gly-Asp-le-Glu-Glu-Ile-Pro-Glu61 Although the direct comparison of sequences is not tremendously convincing, Dr. A. Tulinsky pointed out that the positions of negative charges and hydrophobic residues 98 relative to the central phenylalanine are identical if one of the sequences is considered in reverse (i. e. the sequence read C-terminal to N—terminal): VP16 N—Asp-Ala-Leu—Asp-ASP-P_hC-Asp-Leu-Asp-Met-Leu447 Hirudin C-Glu-Pro-11e—Glu-Glu-Pl'le—Asp-Gly-Asn-Asn-Hi551 . Comparing these sequences in reverse orientation is reasonable in the case of hirudin, since all hirudin/thrombin contacts in this region are side-chain contacts and thus do not necessarily involve the polarity of the polypeptide mainchain. The results of mutagenesis analyses of hirudin—like peptides further suggest that a comparison between hirudin and the VP16 activation domain is relevant. For instance, Stone fl (1989) determined that the affinity of hirudin-like peptides to thrombin was a function of net negative charge. Furthermore, it was found that the phenylalanine of hirudin could be replaced by tyrosine, but not by alanine, glutamate, or leucine or by a variety of nonprotein phenylalanine analogs including D-phenylalanine (Krstenansky et_al., 1987; Owen et_al., 1988). These results are very similar to results that J. Regier (unpublished) and I have observed for Phe 442 of the VP16 activation domain. Finally, prior to the determination of its crystal structure with thrombin (Rydel et_al., 1990) the acidic tail of hirudin was modeled as an AAH (Krstenansky et_al., 1987). A simple test which might strengthen the hirudin/VP16 connection would be to demonstrate that the VP16 activation could bind to the thrombin exosite or that the hirudin tail could activate transcription. The first of these test has been done. Unfortunately, (from a model building perspective) L. Graham and S. Triezenberg Could find no experimental evidence that the VP16 activation domain bound to 99 thrombin (unpublished data) using native gel-electrophoresis experiments. It has not yet been determined if the hirudin polypeptide can activate transcription. The alternative model proposes that Phe 442 is important in hydrophobic contacts involved in the self-folding of the VP16 activation domain. UV spectroscopy data supports this model, in that Tyr 442 of GAL4-FY 442 is buried. This model is also consistent with the observation that various activation domains have a conserved pattern of bulky hydrophobic residues. Future Studies The final determination of the three-dimensional structure of the VP16 activation domain will clearly distinguish between these two models. Fortunately, I have been able to initiate two projects with the potential to elucidate the structure of the VP16 activation domain. These investigations will clearly be aided by the availability of GAL4-VP16 fusion proteins purified from E. coli and the availability of various mutants to correlate structural elements with function. The first project, a collaboration with Dr. G. Wagner at Harvard University, will seek to determine the solution structure of the VP16 activation domain using NMR. Preliminary NMR analysis of the purified activation domain agrees with CD analysis in that the purified activation domain appears to contain virtually no helical structure in solution. The Second project, in collaboration with Dr. A. Tulinsky here at MSU, will seek to crystallize the GAL4-VP16 fusion protein and then to determine its crystal structure using X-ray diffraction. LIST OF REFERENCES Ace, C. I., Dalrymple, M. A., Ramsay, F. H., Preston, V. G. and Preston C. M. (1988). J. Gen. Virol. 69:2595-2605. Allison, L. A. and Ingles, C. J. (1989). Proc. Natl. Acad. Sci. USA 86:2794-2798. Amit, A. G., Mariuzza, R. A., Phillips, S. E. V. and Poljak, R. J. (1986). Science 233:747-753. Azizkhan, J. C., Vaughn, J ., Christy, R. J. and Hamlin, J. H. (1986). Biochemistry 25:6228-6236. Baker, A. and Schatz, G. (1987). Proc. Natl. Acad. Sci. USA 84:3117-3121. Batterson, W. and Roizman, B. (1983). J. Virol. 46:371-377. Berger, 8., Cress, W. D., Cress, A. P., Triezenberg, S. J. and Guarente, L. (1990). Cell 61:1199-1208. Brandl, C. I. and Struhl, K. (1989). Proc. Natl. Acad. Sci. USA 86:2652-2656. Breathnach, R. and Chambon, P. (1981). Annu. Rev. Biochem. 50:349-393. Brent, R. and Ptashne, M. (1985). Cell 43:729-736. Briggs, M. and Gierasch, L. (1986). Adv. Protein Chem. 38:109-180. Bunick, D., Zandomeni, R., Ackerman, S. and Weinmann, R. (1982). Cell 29:877-886. Burton, Z. R, Killeen, M., Sopta, M., Ortolan, L. G. and Greenblatt, J. (1988). Mol. Cell. Biol. 8:1602-1613. Buratowski, S., Hahn, S., Guarente, L. and Sharp, P. A. (1989). Cell 56:549-561. Byler, D. M. and Suzi, H. (1986). Biopolymers 25:469-487. 100 101 Bzik, D. J. and Preston, C. M. (1986). Nucleic Acids Res. 14:929-943. Campbell, M. E. M., Palfreyman, J. W. and Preston, C. M. (1984). J. Mol. Biol. 180:1-19. Carey, M., Leatherwood, J. and Ptashne, M. (1990a). Science 247:710-711. Carey, M., Lin. Y.-S., Green, M. R. and Ptashne, M. (1990b). Nature 345:361-364. Carcamo, J., Lobos, S., Merino, A., Buckbinder, L., Weinmann, R., Venkatachala, N. and Reinberg, D. (1989). J. Biol. Chem. 264:7704-7714. Carcamo, J., Maldonado, E., Cortes, P., Ahn, M.-H., Hho-Ha, Kasai, Y., Flint, J. and Reinberg, D. (1990). Genes Dev. 4:1611—1622. Chasman, D. I., Leatherman, J., Carey, M., Ptashne, M. and Kornberg, R. D. (1989). Mol. Cell. Biol. 9:4746-4749. Chou, P. Y. and G. D. Fasman. (1978). Ann. Rev. Biochem. 47:251-276. Compton, L. A. and Johnson, Jr., W. C. (1986). Anal. Biochem. 155:155-167. Conaway, J. W., Travis, E. and Conaway, R. C. (1990). J. Biol. Chem. 265:7564-7596. Courey, A. J. and Tjian, R. (1988). Cell 55:887-898. Cousens, D. J., Greaves, R. Goding, C. R. and O’Hare, P. (1989). EMBO J. 8:2337-2342. Corden, J., Wasylyk, B., Buchwalder, A., Sassone-Corsi, P., Kedinger, C. and Chambon, P. (1980). Science 209:1405-1414. Cordingley, M. G., Campbell, M. E. M. and Preston, C. M. (1983). Nucleic Acids Res. 11:2347-2365. Corton, J. C. and Johnston, S. A. (1989). Nature 340:724-727. Croston, G. E., Kerrigan, L. A., Lira, L. M., Marshak, D. R. and Kadonaga, J. T. (1991). Science 251:643-649. Dalrymple, M. A., McGeoch, D. J., Davison, A. J. and Preston, C. M. (1985). Nucleic Acids Res. 13:7865-7979. 102 Davidson, B. L., Egly, J .-M., Mulvihill, E. R. and Chambon, P. (1983). Nature 301:680-686. Davis, R. L., Weintraub, H. and Lassar, A. B. (1987). Cell 51:987-1000. Devereux, J ., Haeberli, P. and Smithies, O. (1984). Nuc. Acids Res. 12:387-395. Dignam, D. L., Levobitz, R. M., and Roeder, R. G. (1983) Nuc. Acids Res. 11:1475-1485. Douglas, H. D. and Hawthorne, D. C. (1964). Genetics 49:837-844. Donovan, J. W. (1984). TIBS:340-344. Dynan, W. S. and Tjian, R. (1985). Cell 35:79-87. Eisenmann, D. M., Dollard, C. and Winston, F. (1989). Cell 58:1183-1191. Elgin, S. C. R. (1988). J. Biol. Chem. 263:19259-19262. Evans, R. M. (1988). Science 240:889-895. Fire, A., Samuels, M. and Sharp, P. A. (1984). J. Biol. Chem. 259:2509-2516. Flores, O., Maldonado, E. and Reinberg, D. (1989). J. Biol. Chem. 264:8913-8921. Forsburg, S. L. and Guarente, L. (1989). Genes Dev. 3:1166-1178. Freedman, L. P., Luisi, B. F., Korszun, Z. R., Basavappa, R., Sigler, P. B. and Yamamoto, K. R. (1988). Nature 334:543-546. Fried, M. and Crothers, D. M. (1981). Nucleic Acids Res. 9:6505-6525. Gadhavi, P. L., Raine, A. R. C., Alefounder, P. R. and Laue, E. D. (1990). FEBS 276249-53. Gill, G. and Ptashne, M. (1987). Cell 51:121-126. Giniger, E. and Ptashne, M. (1987). Nature 330:670-672. Gerster, T. and Roeder, R. G. (1988). Proc. Natl. Acad. Sci. U.S.A. 85: 6347-635 1 . Gibson, W. and Roizman, B. (1974). J. Virol. 13:155-165. 103 Gilmore, D. S. and Lis, J. T. (1986). Mol. Cell. Biol. 6:3984-3989. Godowski, P. J ., Picard, D. and Yamamoto, K. R. (1988). Science 241:812-816. Graves, B. J ., Eisenman, R. N. and McKnight, S. L. (1985). Mol. Cell. Biol. 5: 1948-1958. Greaves, R. and O’Hare, P. (1989). J. Virol. 63:1641-1650. Greaves, R. and O’Hare, P. (1990). J. Virol. 64:2716-2724. Hahn, S., Buratowski, S., Sharp, P. A. and Guarente, L. (1989). Cell 58:1173-1181. Hai, T., Horikoshi, M., Roeder, R. G. and Green, M. R. (1988). Cell 54:1043-1051. Heine, J. W., Honess, R. W., Cassai, E. and Roizman, B. (1974). J. Virol. 14:640-651. Hoey, T. Dynlacht, B. D., Peterson, M. G. Pugh, B. F. and Tjian, R. (1990). Cell 61:1179-1186. Hollenberg, S. M. and Evans, R. M. (1988). Cell 55:899-906. Hope, I. A. and Struhl, K. (1986). Cell 46:885-894. Hope, I. A., Mahadevan, S. and Struhl, K. (1988). Nature 333:635-640. Horikoshi, M., Hai, T., Lin, Y.-S., Green, M. R. and Roeder, R. G. (1988). Cell 54: 1033-1042. Hu, J. C., O’Shea, E. K., Kim, P. S. and Sauer, R. T. (1990). Science 250: 1400-1403. Ingles, C. J ., Shales, M., Cress, W. D., Triezenberg, S. J. and Greenblatt, J. (in press). Nature. Jones, K. A. and Tjian, R. (1985). Nature 317:179-182. Johnson. P. F., Landshultz, W. H., Graves, B. J. and McNight, S. L. (1987). Genes Dev. 1:133-146. Johnson, P. F. and McNight, S. L. (1989). Annu. Rev. Biochem 58:799-839. 104 Kaiser, C., Pruess, D., Grisafi, P. and Botstein, D. (1987). Science 235:312-317. Katan, M., Haigh, A., Verrijzer, C. P., Vandervliet, P. C. and O’Hare, P. (1990). Nucleic Acids Res. 18:6871-6880. Kauppinen, J. K., Moffat, D. J., Mantsch, H. H. and Cameron, D. G. (1981). Applied Spectroscopy 35:271-276. Kelleher, R. J., Flanagan, P. M. and Kornberg, R. G. (1990). Cell 61:1209-1215. Khoury, G. and Gruss, P. (1983). Cell 33:313-314. Kissinger, C. R., Liu, B., Martin-Blanco, E., Kornberg, T. B. and Pabo, C. O. (1990). Cell 63:579-590. Klapper, M. H. (1977). Biochem. Biophys. Res. Comm. 78:1018-1024. Krishnan, K. S. and Brandts, J. F. (1978) Methods Enzymol. 49:3-14. Kristie, T. M. and Roizman, B. (1984). Proc. Natl. Acad. Sci. USA 81:4065—4069. Kristie, T. M. and Roizman, B. (1987). Proc. Natl. Acad. Sci. USA 84:71-75. Kristie, T. M., LeBowitz, J. H. and Sharp, P. A. (1989). EMBO J. 8:4229-4238. Kristie, T. M. and Sharp, P. A. (1990). Genes Dev. 4:2383-2396. Krstenansky, J. L., Owen, T. J., Yates, M. T. and Mao, S. J. T. (1987). J. Med. Chem. 30:1688-1691. Kunkel, T. A. (1985). Proc. Natl. Acad. Sci. USA 82:488-492. Laemmli, U. D. (1970). Nature 227:680-685. LaMarco, K. L. and McKnight, S. L. (1989). Genes Dev. 3:1372-1383. Landschulz, W. H., Johnson, P. F. and McKnight, S. L. (1988). Science 240: 1759-1764. Lathrop, R. H., Webster, T. A., Smith, T. F. and Winston, P. H. (in press). In Artificial Intelligence at MIT:Expanding Frontiers (Winston, P. H. and Shellard, S. A. eds.), pp. 70-103, IVHT Press, Cambridge, MA. Levine, M. and Manley, J. L. (1989). Cell 59:405-408. 105 Lesk, A. M. and Chothia, C. J. (1980). J. Mol. Biol. 136:225-270. Lin, Y.-S., Carey, M. F., Ptashne, M. and Green, M. R. (1988). Cell 54:659—664. Lin, Y.—S. and Green, M. R. (1991). Cell 64:971-981. Ma, J. and Ptashne, M. (1987a). Cell 48:847-853. Ma, J. and Ptashne, M. (1987b). Cell 51:113-119. Mackem, S. and Roizman, B. (1982). J. Virol. 44:939-949. Manavalan, P. and Johnson, Jr., W. C. (1987). Anal. Biochem. 167276-85. Maniatis, T., Goodboum, S. and Fischer, J. A. (1987). Science 236:1237—1244. Mardsen, H. S., Campbell, M. E. M., Haarr, L., Frame, M. C., Parris, D. S., Murphy, M., Hope, R. G., Muller, M. T. and Preston, C. M. (1987). J. Virol. 61:2428-2437. Matsui, T., Segall, J., Weil, A. P. and Roeder, R. G. J. Biol. Chem. 257:14419- 14427. McKnight, J. L. C., Kristie, T. M. and Roizman, B. (1987). Proc. Natl. Acad. Sci. USA 84:7061-7065. McKnight, S. L. and Kingbury, R. (1982). Science 217:616-324. McKnight, S. L. and Tjian, R. (1986). Cell 46:795-805. Mermelstein, F. H., Flores, O. and Reinberg, D. (1989). Biochemica and Biophysica Acta 1009: 1—10. Mermod, M., O’Neill, E. A., Kelly, T. J. and Tjian, R. (1989). Cell 58:741—753. Miller, J., McLachlan, A. D. and Klug, A. (1985). EMBO J. 4:1609—1644. Mitchell, P. J. and Tjian, R. (1989). Science 245:371-378. Murre, C, McCaw, P. S., Vaessin, H., Caudy, M., Jan, L. Y. ,,Jan Y. N., Cabrera, C., Buskin, D. C., Lassar, A. B. ,Weintraub, H. and Baltimore, D. (1989). Cell 58:537-544. Myers, R. M., Tilly, K. and Maniatis, T. (1986). Science 232:613—618. 106 Nakatani, Y., Horikoshi, M., Brenner, M., Yamamoto, T., Besnard, F., Roeder, R. G. and Freese, E. (1990). Nature 348:86—88. O’Shea, E. K., Rutkowski, R. and Kim, P. S. (1989). Science 243:538—542. O’Hare, P. and Goding, R. C. (1988). Cell 52:435-445. O’Hare, P., Goding, C. R. and Haigh, A. (1988). EMBO J. 7:4231-4238 Ohkuma, Y., Sumimoto, H., Horikoshi, M. and Roeder, R. G. (1990). Proc. Natl. Acad. Sci. USA 87:9163-9167. Owen, T. J., Krstenansky, J. L., Yates, M. T. and Mao, S. J. T. (1988). J. Med. Chem. 31:1009—1011. Pan, T. and Coleman, J. (1989). Proc. Natl. Acad. Sci. USA 86:3145-3149. Pan, T. and Coleman, J. (1990). Proc. Natl. Acad. Sci. USA 87:2077—2081. Pellet, P., McKnight, J. L. C., Jenkins, F. and Roizman, B. (1985). Proc. Natl. Acad. Sci. USA 82:5870-5874. Peterson, M. G., Tanese, N., Pugh, B. F. and Tjian, R. (1990). Science 248:1625-1630. Post, L. E., Mackem, S. and Roizman B. (1981). Cell 24:555—565. Preston, C. M., Cordingley, M. G. and Stow, N. D. (1984). J. Virol. 50:708-716. Preston, C. M., Frame, M. C. and Campbell, M. E. M. (1988). Cell 52:425—434. Ptashne, M. (1986). Nature 322:697-701. Ptashne, M. (1988). Nature 355:683—689. Ptashne, M. and Gann, A. A. F. (1990). Nature 346:329-331. Pugh, B. F. and Tjian, R. (1990). Cell 61:1187—1197. Reinberg, D. and Roeder, R. G. (1987). J. Biol. Chem. 262:3310—3321. Reinberg, D., Horikoshi, M. and Roeder, R. G. (1987). J. Biol. Chem. 262:3322-3330. 107 Roeder, R. G. (1976). In RNA Polymerases (Losick, R. and Chamberlin, M., eds.), pp. 285-300, Cold Spring Harbor, NY. Roizman, B. and Spector, D. (1991). In Herpesvirus Transcription and its Regulation (Wagner, E. ed.), pp. 17-28. CRC Press, Boca Raton, FL. Roizman, B. and Sears, A. E. (1990). In Virology (Fields, B. and Knipe, D. eds.), pp. 1795-1842. Raven Press, New York. Rougvie, A. E. and Lis, J. T. (1988). Cell 54:795-804. Rydel, T. J ., Ravichandran, K. G., Tulinski, A., Bode, W., Huber, R., Roistsch, C. and Fenton H, J. W. (1990). Science 249:277-280. Sadowski, 1., Ma, J., Triezenberg, S. J. and Ptashne, M. (1988). Nature. 335:563-564. Samuels, M., Fire, A. and Sharp, P. A. (1982). J. Biol. Chem 257:14419-14427. Sanger, F., Nicklen, S. and Coulson, S. A. (1977). Proc. Natl. Acad. Sci. USA 74:5463-5467. Sawadogo, M. and Sentenac, A. (1990). Annu. Rev. Biochem. 56:711-754. Schatz, G. (1987). Eur. J. Biochem. 165:1—6. Scott, M. P., Tamkun, J. W. and Hartzell HI, G. W. (1989). Biochim. Biophys. Acta 989:25-48. Sigler, P. (1988). Nature 333:210-212. Smale, S. T. and Baltimore, D. (1989). Cell 57:103-113. Smale, S. T., Schmidt, M. C., Berk, A. J. and Baltimore, D. (1990). Proc. Natl. Acad. Sci. USA 87:4509-4513. Spear, P. G. and Roizman, B. (1972). J. Virol. 9:431-439. Spector, D., Purves, F. and Roizamn, B. (1990). Proc. Natl. Acad. Sci. USA 87:5268-5272. Stanfield, R. L., Fieser, T. M., Lerner, R. A. and Wilson, I. A. (1990). Science 248:712-719. Stern, S., Tanaka, M. and Herr, W. (1989). Nature 341:624-630. 108 Stone, S. R., Dennis, S. and Hofsteenge, J. (1989). Biochemistry 28:6857—6863. Stringer, K. F., Ingles, C. J. and Greenblatt, J. (1990). Nature 344:783-786. Sumimoto, H., Ohkuma, Y., Yamamoto, T., Horikoshi, M. and Roeder, R. G. (1990). Proc. Natl. Acad. Sci. USA 87:9158-9162. Tasset, D., Tora, L., Fromental, C., Scheer, E. and Chambon, P. (1990). Cell 62:1177-1187. Triezenberg, S. J ., Kingsbury, R. C. and McKnight, S. L. (1988a). Genes Dev. 2:718-729. Triezenberg, S. J ., LaMarco, K. L. and McKnight, S. L. (1988b). Genes Dev. 21730—742. Trewhella, J ., Liddle, W. K., Heidom, D. B. and Strynadka, N. (1989). Biochemistry 28:1294-1301. Vinson, C. R., Sigler, P. B. and McKnight, S. L. (1989). Science 246:911-916. von Heijne, G. (1985). J. Mol. Biol. 184:99-105. Voronova, A. and Baltimore, D. (1990). Proc. Natl. Acad. Sci. USA 87:4722-4726. Wagner, E. K. (1990). In Herpesvirus Transcription and its Regulation (Wagner, E. K., ed.), pp 1-16. CRC Press, Boca Raton, FL. Werstuck, G. and Capone, J. P. (1989). Gene 75:213-224. Werstuck, G. and Capone, J. P. (1990). J. of Virology 63:5509-5513. Whitley, R. J. (1990). In Virology (Fields, B. and Knipe, D., eds.), pp. 1843-1888. Raven Press, New York. Workman, J. L., Taylor, I. C. A. and Kingston, R. E. (1991). Cell 64:533-544. Yanisch-Perron, C., Vieira, J. and Messing, J. (1985). Gene 33:103-119. Zhu, Q., Smith, T. F., Lathrop, R. H. and Figge, J. (1990). Proteins 82156-163. Zoller M. J. and Smith, M. (1982). Nucleic Acids Res. 10:6487-6500. CI V. ”Influlfilfllljlfllfllumll 91 111i“