the j biological c © 2005 by the american society for ... biol chem. 2005.pdf ·...

14
Unraveling the Mechanism of a Potent Transcriptional Activator* S Received for publication, May 4, 2005 Published, JBC Papers in Press, May 10, 2005, DOI 10.1074/jbc.M504895200 Zhen Lu‡§, Steven P. Rowe§, Brian B. Brennan§, Sarah E. Davis‡, Renee E. Metzler‡, Johnathan J. Nau‡, Chinmay Y. Majmudar, Anna K. Mapp**, and Aseem Z. Ansari‡ ‡‡§§ From the Department of Biochemistry and ‡‡The Genome Center, University of Wisconsin, Madison, Wisconsin 53706 and the Department of Chemistry and Department of Medicinal Chemistry, University of Michigan, Ann Arbor, Michigan 48109 Despite their enormous potential as novel research tools and therapeutic agents, artificial transcription factors (ATFs) that up-regulate transcription robustly in vivo remain elusive. In investigating an ATF that does function exceptionally well in vivo, we uncovered an unexpected relationship between transcription func- tion and a binding interaction between the activation domain and an adjacent region of the DNA binding do- main. Disruption of this interaction leads to complete loss of function in vivo, even though the activation do- main is still able to bind to its target in the transcrip- tional machinery. We propose that this interaction par- allels those between natural activation domains and their regulatory proteins, concealing the activation do- main from solvent and the cellular milieu until it binds to its transcriptional machinery target. Inclusion of this property in the future design of ATFs should enhance their efficacy in vivo. The expression of specific genes is mediated by transcrip- tional regulators that bind to DNA and interact with various protein partners to exert transcriptional control. There is in- creasing evidence that malfunctioning transcriptional regula- tors play a central role in the etiology of human diseases as diverse as cancer and diabetes (1–3). The development of arti- ficial transcription factors (ATFs), 1 molecules that target spe- cific genes and control their transcriptional levels, has thus become a highly active area of research. ATFs will likely have a clear role in the study and eventual treatment of transcrip- tion-based disorders (2, 4 –7). A significant challenge has been the creation of fully functional artificial transcriptional activators that seek out and robustly up-regulate specific genes (4, 5, 7). Endogenous activators are modular proteins minimally com- posed of two domains as follows: a DNA binding domain that dictates the genes to be targeted and an activation domain that governs the nature and the extent of the transcriptional re- sponse through interactions with the transcriptional machin- ery (Fig. 1A) (6). The two domains typically operate in an independent fashion such that the DNA binding domain of one protein can be attached to the activation domain of another to generate a fully functional activator. The nature of the linker connecting the two domains is variable but most often consists of a dimerization domain that facilitates specific and high affinity DNA binding (6). Inspired by the architecture of natu- ral activators, artificial activators have typically been gener- ated through a modular replacement strategy in which the endogenous DNA binding domain is replaced with an engi- neered counterpart with desired DNA targeting properties (4, 7, 8). By using this strategy, protein, oligonucleotide, and small molecule DNA binding domains have been used in the con- struction of ATFs that function well in vitro and/or in cell culture (4 –11). In contrast to the diversity of DNA binding domains used in ATF construction, it has proven much more challenging to identify non-natural activation domains that function robustly in vivo (4, 5, 7). In fact, most ATFs that elicit strong expression of a targeted gene in cells utilize activation domains that are derived from natural transcription factors, such as VP16 or Gal4 (4, 10, 12–16). Attempts to develop non-natural activation domains have predominantly yielded peptides or, more re- cently, RNA molecules that function only moderately relative to natural systems (11, 17–26). There are, however, two artifi- cial activators that are exceptions to this trend and function strongly in vivo; these thus serve as excellent models for the design or discovery of potent artificial activation domains. One potent activation domain is P201, a hydrophobic peptide (YL- LPTCIP, see Fig. 1B) that when coupled to the DNA binding domain of Gal4 (residues 1–100) activates transcription to lev- els that are comparable to those obtained by some of the most potent natural activators (Fig. 1) (19, 20). In contrast to natural activators that typically interact with a number of transcrip- tional machinery proteins, the function of P201 is most likely mediated by targeting the Gal11 component of the transcrip- tional machinery. In support of this, a single point mutation in Gal11 (T322K) abolishes the ability of P201 to activate transcrip- tion in vivo and the ability to bind to that protein in vitro (20). The second example of a potent artificial activation is ob- served when Gal4(1–100), a fragment of the transcriptional activator Gal4, is present in a yeast strain bearing a mutated version of Gal11. The first 50 residues of Gal4 compose the minimal DNA binding domain, and the next 50 residues form an extended dimerization domain (see Fig. 1B). In normal yeast strains, Gal4(1–100) is an inert DNA-binding protein. How- * The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. S The on-line version of this article (available at http://www.jbc.org) contains Figs. 1–3. § These authors contributed equally to this work. Supported by National Institutes of Health Grant GM07863 and NASA Graduate Research Fellowship F008372. ** Supported by National Institutes of Health Grant GM65330, the Burroughs Wellcome Fund, and the March of Dimes. To whom corre- spondence may be addressed: Dept. of Chemistry, University of Mich- igan, Ann Arbor, MI 48109. E-mail: [email protected]. §§ Supported by the March of Dimes Foundation, the Steenback Career Development Award, and Industrial and Economic Develop- ment Research funds from University of Wisconsin-Madison and Wis- consin Alumni Research Foundation. To whom correspondence may be addressed: Dept. of Biochemistry, 433 Babcock Dr., University of Wis- consin, Madison, WI 53706. Tel.: 608-265-4690; Fax: 608-262-3453; E-mail: [email protected]. 1 The abbreviations used are: ATFs, artificial transcription factors; GST, glutathione S-transferase. THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 280, No. 33, Issue of August 19, pp. 29689 –29698, 2005 © 2005 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A. This paper is available on line at http://www.jbc.org 29689 at University of Wisconsin-Madison on December 4, 2015 http://www.jbc.org/ Downloaded from at University of Wisconsin-Madison on December 4, 2015 http://www.jbc.org/ Downloaded from at University of Wisconsin-Madison on December 4, 2015 http://www.jbc.org/ Downloaded from at University of Wisconsin-Madison on December 4, 2015 http://www.jbc.org/ Downloaded from at University of Wisconsin-Madison on December 4, 2015 http://www.jbc.org/ Downloaded from

Upload: others

Post on 26-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

Unraveling the Mechanism of a Potent Transcriptional Activator*□S

Received for publication, May 4, 2005Published, JBC Papers in Press, May 10, 2005, DOI 10.1074/jbc.M504895200

Zhen Lu‡§, Steven P. Rowe§¶�, Brian B. Brennan§¶, Sarah E. Davis‡, Renee E. Metzler‡,Johnathan J. Nau‡, Chinmay Y. Majmudar¶, Anna K. Mapp¶**, and Aseem Z. Ansari‡ ‡‡§§

From the ‡Department of Biochemistry and ‡‡The Genome Center, University of Wisconsin, Madison, Wisconsin 53706and the ¶Department of Chemistry and Department of Medicinal Chemistry, University of Michigan,Ann Arbor, Michigan 48109

Despite their enormous potential as novel researchtools and therapeutic agents, artificial transcriptionfactors (ATFs) that up-regulate transcription robustlyin vivo remain elusive. In investigating an ATF that doesfunction exceptionally well in vivo, we uncovered anunexpected relationship between transcription func-tion and a binding interaction between the activationdomain and an adjacent region of the DNA binding do-main. Disruption of this interaction leads to completeloss of function in vivo, even though the activation do-main is still able to bind to its target in the transcrip-tional machinery. We propose that this interaction par-allels those between natural activation domains andtheir regulatory proteins, concealing the activation do-main from solvent and the cellular milieu until it bindsto its transcriptional machinery target. Inclusion of thisproperty in the future design of ATFs should enhancetheir efficacy in vivo.

The expression of specific genes is mediated by transcrip-tional regulators that bind to DNA and interact with variousprotein partners to exert transcriptional control. There is in-creasing evidence that malfunctioning transcriptional regula-tors play a central role in the etiology of human diseases asdiverse as cancer and diabetes (1–3). The development of arti-ficial transcription factors (ATFs),1 molecules that target spe-cific genes and control their transcriptional levels, has thusbecome a highly active area of research. ATFs will likely havea clear role in the study and eventual treatment of transcrip-tion-based disorders (2, 4–7). A significant challenge has beenthe creation of fully functional artificial transcriptional activatorsthat seek out and robustly up-regulate specific genes (4, 5, 7).

Endogenous activators are modular proteins minimally com-posed of two domains as follows: a DNA binding domain thatdictates the genes to be targeted and an activation domain thatgoverns the nature and the extent of the transcriptional re-sponse through interactions with the transcriptional machin-ery (Fig. 1A) (6). The two domains typically operate in anindependent fashion such that the DNA binding domain of oneprotein can be attached to the activation domain of another togenerate a fully functional activator. The nature of the linkerconnecting the two domains is variable but most often consistsof a dimerization domain that facilitates specific and highaffinity DNA binding (6). Inspired by the architecture of natu-ral activators, artificial activators have typically been gener-ated through a modular replacement strategy in which theendogenous DNA binding domain is replaced with an engi-neered counterpart with desired DNA targeting properties (4,7, 8). By using this strategy, protein, oligonucleotide, and smallmolecule DNA binding domains have been used in the con-struction of ATFs that function well in vitro and/or in cellculture (4–11).

In contrast to the diversity of DNA binding domains used inATF construction, it has proven much more challenging toidentify non-natural activation domains that function robustlyin vivo (4, 5, 7). In fact, most ATFs that elicit strong expressionof a targeted gene in cells utilize activation domains that arederived from natural transcription factors, such as VP16 orGal4 (4, 10, 12–16). Attempts to develop non-natural activationdomains have predominantly yielded peptides or, more re-cently, RNA molecules that function only moderately relativeto natural systems (11, 17–26). There are, however, two artifi-cial activators that are exceptions to this trend and functionstrongly in vivo; these thus serve as excellent models for thedesign or discovery of potent artificial activation domains. Onepotent activation domain is P201, a hydrophobic peptide (YL-LPTCIP, see Fig. 1B) that when coupled to the DNA bindingdomain of Gal4 (residues 1–100) activates transcription to lev-els that are comparable to those obtained by some of the mostpotent natural activators (Fig. 1) (19, 20). In contrast to naturalactivators that typically interact with a number of transcrip-tional machinery proteins, the function of P201 is most likelymediated by targeting the Gal11 component of the transcrip-tional machinery. In support of this, a single point mutation inGal11 (T322K) abolishes the ability of P201 to activate transcrip-tion in vivo and the ability to bind to that protein in vitro (20).

The second example of a potent artificial activation is ob-served when Gal4(1–100), a fragment of the transcriptionalactivator Gal4, is present in a yeast strain bearing a mutatedversion of Gal11. The first 50 residues of Gal4 compose theminimal DNA binding domain, and the next 50 residues forman extended dimerization domain (see Fig. 1B). In normal yeaststrains, Gal4(1–100) is an inert DNA-binding protein. How-

* The costs of publication of this article were defrayed in part by thepayment of page charges. This article must therefore be hereby marked“advertisement” in accordance with 18 U.S.C. Section 1734 solely toindicate this fact.

□S The on-line version of this article (available at http://www.jbc.org)contains Figs. 1–3.

§ These authors contributed equally to this work.� Supported by National Institutes of Health Grant GM07863 and

NASA Graduate Research Fellowship F008372.** Supported by National Institutes of Health Grant GM65330, the

Burroughs Wellcome Fund, and the March of Dimes. To whom corre-spondence may be addressed: Dept. of Chemistry, University of Mich-igan, Ann Arbor, MI 48109. E-mail: [email protected].

§§ Supported by the March of Dimes Foundation, the SteenbackCareer Development Award, and Industrial and Economic Develop-ment Research funds from University of Wisconsin-Madison and Wis-consin Alumni Research Foundation. To whom correspondence may beaddressed: Dept. of Biochemistry, 433 Babcock Dr., University of Wis-consin, Madison, WI 53706. Tel.: 608-265-4690; Fax: 608-262-3453;E-mail: [email protected].

1 The abbreviations used are: ATFs, artificial transcription factors;GST, glutathione S-transferase.

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 280, No. 33, Issue of August 19, pp. 29689–29698, 2005© 2005 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

This paper is available on line at http://www.jbc.org 29689

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 2: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

ever, when a single point mutation (N342V) is introduced intoGal11, Gal4(1–100) stimulates high levels of transcription (27–29). Several studies indicate that this mutant version of Gal11,called “potentiator” or Gal11P, retains the normal function ofGal11 but is able to interact with the dimerization domain of

Gal4 (Gal4dd, residues 50–100; see Fig. 1, B and D) (28). Struc-ture-activity studies indicate that the primary interactionswith Gal11P occur in helix 3 and loop 1 of Gal4dd, both of whichcontain a number of hydrophobic residues (28).

Gal11, a primary target of two of the most potent artificialactivators, is also one of the targets of natural yeast activators(20, 27, 29–32). Gal11 resides in the mediator complex; thismultiprotein complex interacts with RNA polymerase II and isthought to mediate the positive or negative signals of transcrip-tional regulators bound to their respective genes (33, 34). Insupport of this model, a number of mediator subunits havebeen found to interact with transcriptional activators (35–37),and several of the targeted components, including Gal11, re-side in the tail of this tri-lobed complex (Fig. 1C) (38, 39).Because both of the potent artificial activators described abovefunction through an interaction with Gal11, this suggests thatGal11 could be a privileged target for the development of potentactivation domains.

Recent evidence has emerged indicating that exclusive tar-geting of Gal11 does not provide potent transcriptional acti-vation domains. For example, peptides selected specificallyfor their ability to bind to Gal11 function only modestly astranscriptional activators despite affinities for Gal11 compa-rable with natural activation domains (23). This suggeststhat the binding interaction with Gal11 is unlikely to be thesole contributor to the unusually potent function of P201. Togain insight into additional properties of P201 that make it arobust artificial activator, we examined its ability to activatetranscription in yeast strains bearing various mutants ofGal11 and to bind these Gal11 mutants in vitro. These ex-periments indicate the existence of a second binding interac-tion that is essential for P201-mediated activation. Remark-ably, this interaction is not with a transcriptional machineryprotein but with the hydrophobic dimerization domain ofGal4 (Gal4dd). We propose that this dynamic interaction be-tween Gal4dd and P201 plays a role analogous to the inter-actions between natural activation domains and their regu-latory proteins, preserving the hydrophobic activationsequences from nonproductive binding events with other cel-lular proteins. This finding provides an important guidingprinciple for the future design of robust artificial activators.

MATERIALS AND METHODS

Yeast Strains and Plasmids—The yeast strain used wasJPY52::JP188 MAT� his3�200 leu2�1 trp1�63 ura3-52::JP188lys2�385 gal4�11 gal11::LYS2. This strain contains a GAL1-lacZ re-porter with two GAL4 DNA-binding sites 191 bp upstream of the TATAbox integrated at the URA3 locus. The Gal4dd or P201 constructs wereon a CEN4 plasmid under the control of a �-actin promoter with a HIS�

selection. The GAL11 constructs were based on YCplac111 (CENLEU�).

Expression and Purification of GST-Gal11(186–619)—The plasmidpGEXVsh1 was transformed into chemically competent BL21(DE3) pL-ysE Escherichia coli (Invitrogen), and cells were plated onto LB agarplates supplemented with ampicillin (0.1 mg/ml) and chloramphenicol(0.034 mg/ml). Cultures (50 ml) from single colonies were grown over-night at 37 °C (275 rpm) in LB supplemented with ampicillin (0.1mg/ml) and chloramphenicol (0.034 mg/ml) before addition to 1 liter ofLB supplemented with ampicillin (0.1 mg/ml). After 3 h, the cultureswere cooled to 16 °C, and expression was induced with isopropyl 1-thio-�-D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cellpellet was lysed using sonication, and the GST-tagged protein wasisolated from the cell lysate using glutathione-Sepharose beads (Amer-sham Biosciences). Elution from the beads was accomplished with 50mM Tris�HCl buffer, pH 8.0, containing 15 mM glutathione and 0.1%Nonidet P-40. The protein solution thus obtained was concentrated to�20–25 �M and the buffer exchanged to storage buffer (phosphate-buffered saline, pH 7.4, 1 mM dithiothreitol, 10% glycerol (v/v), and0.01% Nonidet P-40) using a Millipore Ultrafree centrifugal filter de-vice. The protein solution was stored in 50-�l aliquots at �80 °C untilneeded. The protein concentration was measured by using a Bradford

FIG. 1. Scaled illustration of the artificial activator and theyeast RNA polymerase II holoenzyme. A, the construction of atranscriptional activator. The DNA binding domain (DBD) is shown asa green square and the activation domain (AD) is shown in red. B, thestructural model of the Gal4 dimerization domain (residues 53–97, seeRef. 28) with key structural elements denoted (�-helices, �1–3 and loopsL1 and L2) is shown along with a scaled representation ofFVQD�YLLPTCIP (Gal4(97–100)�P201). C, the composite activatorGal4(1–100)�P201 (in green) was modeled by superimposing the crys-tal structure of Gal4 residues 1–64 (59) on the NMR-derived structureof Gal4 residues 53–97 (28). P201 and residues 98–100 of Gal4 areshown as fully extended peptides (brown). RNA polymerase II holoen-zyme structure (�36 Å) is constructed from electron microscopy (39).The polymerase is colored in turquoise, and the head (H) and middle (M)lobes of the Mediator complex are colored in dark blue. The tail (T)domain of the Mediator is colored in purple; Gal11 the target of theartificial activator as well as several natural activators is a componentof this module. The Gal4-binding sites can be separated from the pro-moter (TATA box) by variable distances, and this is represented by agap in the DNA template. The interaction between the activator and theholoenzyme is sufficient to elicit the expression from the promoter.However, additional factors that are not shown in this illustration suchas the TATA box-binding protein and other general transcription fac-tors (TFIIB, TFIIF, TFIIE, TFIIH, and perhaps TFIIA) are required fortranscription (27). D, the replacement of asparagine 342 in Gal11 (pur-ple) with valine (shown as a green patch within Gal11) leads to abinding interaction with the dimerization domain of Gal4 (green). Thisinteraction converts Gal4dd into a potent artificial activator.

Examining the Potency of an Artificial Transcription Factor29690

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 3: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

assay (Bio-Rad) with bovine serum albumin as the standard. The iden-tity and purity of the fusion protein was verified by reducing SDS-PAGE with appropriate molecular weight standards.

�-Galactosidase Assays—For quantitative measurement of activity,freshly transformed colonies were used to inoculate 5-ml cultures of SCmedia containing 2% raffinose, 2% galactose but lacking uracil, histi-dine, and leucine. The cultures were incubated overnight at 30 °C at250 rpm. Following incubation, these cultures were used to inoculate5-ml cultures of SC media, which were subsequently incubated over-night at 30 °C with agitation to an A600 of 0.6–0.9. The yeast cells wereharvested, and the culture was then lysed with glass beads in Z buffer(for 500 ml: 4.26 g of sodium phosphate, 2.76 g of sodium phosphatemonobasic, 0.37 g of potassium chloride, 0.123 g of magnesium sulfate,pH 7.0) for 5 min in the cold. A portion of the cell extract was used tomeasure �-galactosidase activity via incubation with o-nitrophenyl-�-D-galactopyranoside at 4 mg/ml in Z buffer. The reaction was stopped byadding 1 M Na2CO3, and the A420 was measured. The activity reportedwas normalized to the total protein concentration of the extract, whichwas measured using a Bradford assay kit (Bio-Rad) with bovine serumalbumin as the standard.

Peptide Synthesis and Fluorescent Labeling—XLY (LTGLFVQDYL-LPTCIP) and XLR (LTGLFVQDRLLPTCIP) were synthesized on Rinkamide resin from Fmoc (N-(9-fluorenyl)methoxycarbonyl)-protectedamino acids, using standard protocols as in Ref. 61. At the conclusion ofthe synthesis, the peptides were fully deprotected, precipitated withcold ether, and purified to homogeneity using reversed-phase highpressure liquid chromatography (C18 column with a gradient solventsystem: buffer A, 0.1% trifluoroacetic acid; buffer B, CH3CN). Eachpeptide was characterized using electrospray mass spectrometry. Thepeptides were labeled at the N terminus with fluorescein using 5/6-carboxyfluorescein succinimidyl ester (Pierce) in accordance with themanufacturer’s instructions, purified using reversed-phase high pres-sure liquid chromatography, and characterized using electrospray massspectrometry. The labeled peptides were divided into 2-nmol aliquotsand stored as dry pellets at �80 °C.

Dissociation Constant Measurements—Dissociation constant meas-urements were carried out on a Spex Fluoromax-2 fluorimeter at roomtemperature. Prior to each experiment, a 2-nmol aliquot of fluorescein-labeled XLY or XLR was resuspended in 1 ml of Buffer S (phosphate-buffered saline, pH 7.4, 1 mM dithiothreitol, 10% glycerol (v/v), and0.01% Nonidet P-40), and the concentration of the solution was deter-mined by using the absorbance of fluorescein. The concentration of thesolution was adjusted to 1.25 �M with Buffer S, and 1 �l of this solutionwas added to 49 �l of GST�Gal11, GST�Gal11P, or GST�Gal11DM(20 �M final protein concentration for most measurements). After incu-bation for 10 min at room temperature, the first measurement wastaken (excitation wavelength, 487 nm; emission wavelength, 516 nm).For each successive measurement, the solution was diluted by additionof Buffer S containing 25 nM fluorescein-labeled XLY or XLR (to keep theconcentration of the labeled component constant throughout the exper-iment), manually mixed, and incubated for 10 min at room tempera-ture. At concentrations above 25 �M, GST�Gal11(186–619) aggregatesto a significant degree, and dissociation constants greater than 10 �M

thus were not measured. The same procedure was used for bindingexperiments with GST�Gal4dd or GST�Gal4dd�P201, except that a 10nM concentration of XLY or XLR was used. The data obtained wereplotted in Origin 7.0 and fit to the following equation using the Leven-berg-Marquardt least squares method: Y � ((a � X)/(Kd � X)) � b,where a is the amplitude of the curve and b is the anisotropy of the freelabeled peptide. All experiments were performed in triplicate, and er-rors were calculated using the mean � S.D. of the three experiments.

Gel Shift Experiments—Reaction buffer (phosphate-buffered saline,pH 7.4, 1 mM dithiothreitol, 10% glycerol (v/v), 0.01% Nonidet P-40),Gal4(1–100), DNA master mix (32P-5�-end-labeled DNA, salmonsperm DNA, reaction buffer), and Gal11, Gal11P, or Gal11DM wereused to make up 10-�l reactions. The reactions were incubated atroom temperature for 45 min. A 7% acrylamide, 3% glycerol gel waspre-run for 15 min. The 10-�l reactions were loaded while the gel wasrunning. Gels were run at room temperature at 95 V, dried, exposedto a PhosphorImager screen, and visualized with a PhosphorImager(Amersham Biosciences).

koff Measurements—All koff determinations were carried out at 25 °Con a Beacon 2000 polarization system (Invitrogen). For the Gal11-XLY

interaction, fluorescein-labeled XLY (25 nM in Buffer S) was incubatedwith 18 �M GST�Gal11(186–619) until equilibrium was reached, asdetermined by anisotropy measurements. Following equilibration, a1700-fold excess of unlabeled XLY was added by mixing, and anisotropymeasurements were taken every 11.5 s to monitor the release of the

labeled peptide. An identical procedure was used for the XLY�Gal4dd

experiment, except that a 10 nM concentration of fluorescein-labeledXLY was used and a 50 �M concentration of GST�Gal4dd. A controlexperiment was carried out to determine the loss in anisotropy duesolely to the dilution of the protein and accounted for in the dataanalysis. All data were plotted using Origin 7.0 and fit to a single phaseexponential decay using the Levenberg-Marquardt least squaresmethod.

RESULTS

In Vivo Activation—P201 was originally identified by ascreen in Saccharomyces cerevisiae of random 8-residue pep-tides attached to Gal4(1–100). We initially postulated that thepotent function of Gal4(1–100)�P201 could be due to someresidual affinity of Gal4dd (residues 50–100 of Gal4) for wild-type Gal11. Thus the overall level of activation observed withGal4(1–100)�P201 would be the result of two binding interac-tions, P201�Gal11 and Gal4dd�Gal11. To delineate the relativecontributions of P201 and Gal4dd, we first looked at the func-tion of Gal4(1–100)�P201 and Gal4(1–100) in yeast strainsbearing wild-type Gal11 or its mutants. The first of the keyGal11 mutations is the substitution of threonine 322 to lysine(T322K), a change that disrupts binding with P201. The secondmutation results from replacement of asparagine 342 with avaline (N342V, also referred to as the Gal11P mutation), andthis alteration promotes an interaction with Gal4dd (20, 28). Wealso tested the function of Gal4(1–100)�P201 and Gal4(1–100)in a yeast strain bearing the Gal11 double mutant (Gal11 DM)that has both the T322K and N342V mutations.

Gal4(1–100)�P201 functions well in the wild-type Gal11-bearing strain, and as expected, the T322K substitution inGal11 abolishes P201 activity (Fig. 2, A and B). In contrast,Gal4dd does not activate transcription in yeast strains contain-ing either wild-type Gal11 or Gal11 with the T322K mutation(Fig. 2, C and D). Introduction of a second mutation in Gal11(N342V, Gal11DM) restores activation function to Gal4dd tolevels nearly identical to those observed with Gal11P (Fig. 2, Eand F). The latter result demonstrates that the T322K muta-tion has little effect on the ability of Gal4dd to activate tran-scription. The Gal11P mutant contains binding sites for bothGal4dd and for P201, but as illustrated in Fig. 2G, the levels oftranscription elicited by Gal4(1–100)�P201 in this strain aresimilar to those obtained in the presence of wild-type Gal11(Fig. 2A).

In the final experiment, we examined the function of Gal4(1–100)�P201 in yeast bearing Gal11DM to decouple the effect ofP201 on activation mediated by Gal4dd (Fig. 2H). The T322Kmutation in Gal11 abrogates P201-mediated activation (Fig.2B) but has little effect on the function of Gal4dd (Fig. 2F). Itwould be expected, then, that Gal4(1–100)�P201 would acti-vate transcription in Gal11DM yeast primarily because of theinteraction of Gal4dd with the binding surface generated by theN342V mutation (Fig. 2H). However, activation was signifi-cantly diminished in this strain (Fig. 2H). In other words, P201somehow blocks the function of Gal4dd. Identical results wereobtained in yeast strains bearing a single binding site for Gal4(see supplemental Fig. 1 for details). Taken together, theseresults suggest that although Gal4dd and P201 do not appear tofunction cooperatively, there is a functional interplay betweenthe two peptides. Subsequent experiments were therefore de-signed to first more clearly define the P201�Gal11 interaction,and second to investigate a possible Gal4dd�P201 interaction.

Defining the Minimal Gal11-binding Peptide—The resultsfrom the in vivo activation studies above indicate that Gal4dd

does not interact with wild-type Gal11 in the context of Gal4(1–100)�P201. However, previous mutagenesis studies haveshown that residue 97 in helix 3 of Gal4dd is important for P201function, and there thus remained the possibility that some

Examining the Potency of an Artificial Transcription Factor 29691

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 4: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

portion of Gal4dd is necessary for P201 to interact effectivelywith Gal11 (19). To examine the role of helix 3 of Gal4dd inP201-mediated activation, we carried out alanine-scanningmutagenesis of this region (Fig. 3A). To avoid any effect onDNA binding, we replaced the DNA binding domain of Gal4with the DNA binding domain of the bacterial protein LexA. Asillustrated in Fig. 3A, the replacement of residues 96–98 inGal4dd with alanine eliminated P201-mediated transcription,whereas substitution of residues 92–94 had modest effects. Thein vivo results were correlated in vitro by examining the inter-action of the peptides P201, Gal4(96–100)�P201, and Gal4(93–100)�P201 with Gal11. These experiments revealed that onlyGal4(93–100)�P201 binds to Gal11 with any appreciable affin-ity (Fig. 3B and data not shown). Thus, both the in vivo func-tional data and the in vitro binding experiments are consistentwith a requirement for both the P201 sequence and residues93–100 of Gal4dd for interaction with Gal11 and for transcrip-tional activation. All subsequent binding experiments weretherefore carried out with this minimal peptide (Gal4(93–100)�P201, LTGLFVQDYLLPTCIP), referred to as XLY.

Interaction of XLY with Gal4dd and Gal11—The data of Fig.2H suggest that XLY binds directly to Gal4dd and interfereswith its ability to interact with Gal11DM. To probe a possibleinteraction between XLY and Gal4dd and to measure the affin-ity of XLY for Gal11, we designed a series of fluorescencepolarization assays. For this purpose, XLY was synthesized bystandard methods and labeled with carboxyfluorescein (61).Gal4dd and the central region of Gal11, residues 186–619, wereexpressed and purified as GST fusion proteins. The centralregion of Gal11 was used in these experiments because it istargeted by XLY as well as Gal4dd; it can be readily generatedusing a bacterial expression system, and it is the largest frag-ment of Gal11 that is well behaved at micromolar concentra-tions in vitro.

As shown in Fig. 3, B and C, XLY binds to both Gal11 andGal4dd with dissociation constants in the micromolar range andinteracts with Gal11 �2-fold more tightly (2.2 versus 5 �M,respectively). We also measured the affinity of XLY forGal4dd�P201. This interaction is roughly 3.5-fold weaker thanthe interaction with Gal4dd lacking the fused P201 sequence(compare Fig. 3, C with D). This result indicates that the fusedP201 peptide does not inhibit the XLY�Gal4dd interaction. As acontrol, we also examined the binding properties of a P201sequence containing an arginine rather than a tyrosine residue(XLR, LTGLFVQDRLLPTCIP); this sequence is inactive as atranscriptional regulator in vivo (19). In our experiments XLR

did not bind to Gal11, Gal4dd, or to Gal4dd�P201 under anyconditions examined (data not shown).

Next, we tested to see if XLY would interact with Gal11proteins bearing either the 11P (N342V) mutation singly or incombination with the T322K mutation (Gal11DM). Consistentwith the in vivo data, we found that introduction of the 11Pmutation causes a 3-fold reduction in affinity (Fig. 3E). Alsoconsistent with the lack of activation in vivo (Fig. 2, B and H),XLY does not bind to Gal11DM within the limits of our assay.Furthermore, the inactive version of Gal4(93–100)�P201, XLR,shows no measurable binding interactions with either Gal11Por Gal11DM (data not shown).

NMR studies of Gal4dd suggest that it has several solvent-exposed hydrophobic residues. XLY is also a very hydrophobicpeptide. To probe the specificity of the Gal4dd�XLY interaction,we examined the interaction of Gal4dd with another hydropho-bic Gal11-binding peptide, AHYYYPSE (23). This peptide has asimilar affinity for Gal11 (4.8 �M) relative to XLY (2.2 �M) butis far less robust than XLY as a transcriptional activationdomain (23). Moreover, as illustrated in Fig. 3F, this peptideexhibits no detectable binding to Gal4dd. Consistent with theinability to bind Gal4dd in vitro, we find that this peptide,unlike XLY, does not interfere with the ability of Gal4dd tostimulate transcription in strains bearing the Gal11DM mu-tant (see supplemental Fig. 2 for details). In further experi-ments we found that XLY does not bind nonspecifically tounrelated proteins such as bovine gamma globulin that displayaffinity for hydrophobic surfaces (62) (Fig. 3G). Thus, theGal4dd�XLY binding event cannot simply be ascribed to nonspe-cific hydrophobic interactions.

Interaction of Gal4dd with Gal11—We next determined thestrength of the interaction between Gal4dd and the Gal11 vari-ants by electrophoretic mobility shift assays. Fluorescence po-larization was not used for these experiments because the largesize of the two components (Gal4dd bound to DNA � Gal11 andits mutants) leads to a minimal change in anisotropy uponcomplex formation. The Gal11 proteins and the assay condi-tions (buffer, temperature, and incubation times) employedwere identical to those used in the fluorescence polarizationassays outlined above to allow for meaningful comparisons of

FIG. 2. A–H, levels of transcription elicited by the two artificialactivators in a strain bearing different Gal11 alleles. The reporterbears two cognate 17-bp Gal4-binding sites 191 bp upstream of the Gal1promoter (TATA element) fused to the lacZ reporter gene. This positionis analogous to the location of the natural upstream activating sequence(UASg) at this promoter. The transcriptional activation potential of theactivators is determined by measuring the activity of �-galactosidase(�-gal) in strains transformed with single copy plasmids expressingdifferent alleles of Gal11 at endogenous levels. Only the hydrophobicloop (68–72), the hydrophilic loop, and the second and third helices ofGal4 are shown. The P201 peptide (along with residues 97–100 of Gal4)is shown as an extended peptide. The hydrophobic loop and the thirdhelix (marked by brackets) interact with Gal11P or Gal11DM. Consist-ent with previous results, Gal4dd is inert in the absence of the 11Pmutation. Various alleles of Gal11 are shown. The N342V mutation inGal11P and Gal11DM is shown as a gray patch and the T342K muta-tion as a white patch. The �-galactosidase activity is reported based onquadruplicate independent measurements (see “Materials and Meth-ods”). In parallel control experiments, activation by full-length naturalGal4 (residues 1–881) was not affected in the four strains bearingdifferent Gal11 mutants, indicating that these alleles do not alter thenatural Gal11 functions (data not shown, see also Ref. 20).

Examining the Potency of an Artificial Transcription Factor29692

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 5: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

FIG. 3. Binding behavior of the min-imal activator peptide, XLY. A, resultsfrom alanine-scanning mutagenesis stud-ies on helix 3 of Gal4dd. The LexA DNAbinding domain (1–87) was fused in-frame to Gal4(40–100)�P201, and tran-scriptional activation was monitored by areporter bearing two LexA operators 191bp upstream of the Gal1 promoter drivingthe expression of the lacZ gene. The use ofthe LexA DNA binding domain ensuresthat any loss of activity upon mutagenesisis not due to loss of DNA binding ofGal4�P201 (for example see Ref. 28). The�-galactosidase (�-gal) activity averagedfrom quadruplicate measurements is re-ported, and the S.D. did not exceed 15%.As shown, residues 93–100 of this regionare essential for P201-mediated tran-scriptional activation. This, in combina-tion with the binding data of 3B, indicatesthat the minimal functional activator isGal4(93–100)�P201, named XLY. B, de-termination of the dissociation constantfor the XLY�Gal11 interaction by fluores-cence polarization. Fluorescein-labeledXLY (Gal4(93–100)� P201) at a constantconcentration was incubated at 25 °Cwith increasing concentrations of Gal11.The fluorescence polarization at eachconcentration was measured, and the re-sulting data fit using the Levenberg-Marquardt least squares method to ob-tain the dissociation constants. Eachexperiment was performed in triplicate(R2 0.98) with the error indicated. C,determination of the dissociation con-stant for the XLY�Gal4dd interaction byfluorescence polarization. Conditionswere identical to those used for B. D,determination of the dissociation con-stant for the XLY�Galdd�P201 interac-tion by fluorescence polarization. Condi-tions were identical to those used for B.E, determination of the dissociationconstant for the XLY�Gal11P interactionby fluorescence polarization. Conditionswere identical to those used for B. F,the hydrophobic activator peptideAHYYYPSE does not interact withGal4dd as monitored by fluorescence po-larization. Conditions were identical tothose used for B. G, XLY does not inter-act with bovine gamma globulin as mon-itored by fluorescence polarization.Conditions were identical to those usedfor B.

Examining the Potency of an Artificial Transcription Factor 29693

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 6: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

the binding affinities measured by the two techniques. A con-centration of 30 nM Gal4(1–100) was sufficient to bind satura-bly to its cognate DNA site, and this Gal4-DNA complex wasincubated with increasing concentrations of wild-type Gal11,Gal11P, or Gal11DM. As demonstrated in Fig. 4, Gal4dd doesnot interact with wild-type Gal11, consistent with the lack ofactivation observed in vivo. Introduction of the 11P mutation,however, leads to half of the Gal4-DNA complex being furthershifted to a ternary DNA-Gal4-Gal11P complex at a Gal11Pconcentration of �18 �M. This binding is virtually unchangedby the introduction of the second (T322K) mutation (compareunnumbered lanes 10 and 16 in Fig. 4). Because Gal11 and itsvariants begin to aggregate significantly and precipitate atconcentrations above 24 �M, the concentration of Gal11 couldnot be further increased to determine the complete bindingisotherm; however, the half-maximal binding at 18 �M providesa reasonable indication of the KD. Although it is difficult tomake a direct quantitative comparison because different tech-niques were used to assess binding, the results suggest that theaffinity of Gal4dd for Gal11P is somewhat weaker than for XLY.It is thus not surprising that the Gal4dd�XLY interaction inhib-its Gal4dd-mediated transcription but does not negatively im-pact XLY-mediated activation.

Site of the Gal4dd-XLy Interaction—Both the in vivo activa-tion data of Fig. 2 and the in vitro binding data of Fig. 3 areconsistent with an interaction between Gal4dd and XLY, buttwo key questions regarding this interaction remained. 1) Doesthis interaction contribute to XLY-mediated activation? 2)Where within Gal4dd is the site of this interaction? We ad-dressed these questions by a combination of deletion analysisand mutagenesis experiments (Fig. 5).

In the first set of experiments, the P201 sequence was at-

tached to the LexA DNA binding domain along with fragmentsof Gal4dd of increasing length. As depicted in Fig. 5A, XLY-mediated transcriptional activation was not observed until theentire Gal4dd region (residues 50–100) was incorporated intothe protein construct. The last result of this panel points to theimportance of helix 1 and loop 1 for this process, because theconstruct containing only helices 2 and 3 (Gal4(73–100))showed no activity.

To refine this picture further, additional alanine-substitu-tion mutagenesis experiments of Gal4dd were carried out. Forthis purpose, we again employed the chimeric LexA-Gal4 pro-tein described in Fig. 3A to avoid any impact on DNA binding.Using the structural model of Gal4dd to guide these efforts, wemutated residues in loop1 (L1), helix 2 (�2), loop2 (L2), andhelix 3 (�3) that are important for Gal4dd�Gal11P interactions(see Fig. 1B and Ref. 28). As shown in Fig. 5B, changingresidues 68, 69, 72, and 75 to alanine has a profound effect onXLY-mediated activation, with little or no activity observed,suggesting that loop 1 is the likely site of interaction with XLY.To probe this further, we examined the binding of XLY toGal4dd containing one of those mutations, 69A, again usingfluorescence anisotropy, and no detectable binding was ob-served (see supplemental Fig. 3 for details). As summarized inFig. 5C, several of these residues are also essential for theGal4dd-Gal11P-binding interaction (28). Most interestingly, notall substitutions affected XLY function; L86A and I89A mu-tants were functional (30% active). Both of these residues areessential for the Gal11P�Gal4dd interaction in vitro and forGal4dd transcriptional activity in vivo (28). In contrast, resi-dues Phe-68 and Leu-77 flanking loop 1 and Asp-84 at theC-terminal end of �2 are essential for XLY function but not forthe Gal4dd�Gal11P interaction.

The residues of Gal4dd that are important for XLY functionand for Gal11P interaction are also summarized in Fig. 5C.There are several residues in loop 1 that are essential both forXLY function and for the ability of Gal4dd to interact withGal11P. However, residues that are essential for XLY functionare not entirely coincident with those required for theGal4dd�Gal11P interaction. Taken together, the results suggestthat Gal4dd uses distinct but overlapping surfaces along itslength to interact with Gal11P and with XLY.

Dynamics of Binding—To further characterize theXLY�Gal4dd and the XLY�Gal11 interactions, we measured theoff-rates again by using fluorescence polarization methods (40).Fluorescein-labeled XLY was incubated with either Gal11 orGal4(52–100) under conditions of saturated binding (10–25 nM

XLY, 18 �M Gal11, or 50 �M Gal4(52–100)). Following equili-bration at room temperature, a 1700-fold excess of unlabeledXLY peptide was added as a single aliquot. This leads to ameasurable decrease in polarization as the bound fluorescentpeptide is exchanged with the unlabeled peptide. As shown inFig. 6, we find the dissociation of XLY from Gal4(52–100) isfaster than the dissociation from Gal11. In fact, virtually all ofthe exchange occurred prior to the first measurement (11.5 s).Thus, the koff value of 0.24 s�1 for XLY dissociating fromGal4(52–100) calculated from these data represent the lowerlimit of this rate. In contrast, the dissociation rate for theGal11�XLY interaction is significantly slower, with a koff of 0.06s�1. In contrast, the koff for Gal4dd dissociating from Gal11Phas been measured previously as 5 s�1 by NMR titration (28).Thus, although the measurements between methods are notdirectly comparable, this off-rate is significantly faster than theXLY�Gal4dd off-rate reported in Fig. 6.

DISCUSSION

The hydrophobic octapeptide P201 functions as an unusuallyrobust transcriptional activator when fused to the Gal4 DNA

FIG. 4. Electrophoretic mobility shift assays with Gal4(1–100).The mobility of a 30-bp oligonucleotide bearing a single Gal4-bindingsite (17 bp) decreases in the presence of Gal4(1–100) but not Gal11,Gal11P, or Gal11DM (lanes 2, 3, 5, and 11). The mobility of the Gal4-DNA complex is further decreased by increasing concentrations ofGal11P and Gal11DM but not by Gal11 (20 �M in lane 4). The finalconcentrations of Gal11P and Gal11DM in lanes 6–10 and 12–16 are 1,3, 9, 18, 20 �M, respectively. Where incubated with DNA, in the absenceof Gal4, the concentration of the Gal11 proteins is 20 �M (lanes 3, 5and 11).

Examining the Potency of an Artificial Transcription Factor29694

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 7: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

binding and dimerization domains (residues 1–100) (19, 20).Our study demonstrates that in addition to the 8 amino acidsthat make up P201, residues 93–100 of Gal4 are essential fortranscriptional activation and also for interaction with Gal11(Fig. 3A). The minimal functional activator is thus the 16-residue sequence Gal4(93–100)�P201, named XLY. The po-tency of XLY as an activation domain is particularly remarka-ble when one considers that other artificial activator peptidesof greater size and surface area up-regulate only moderately(17, 22, 23). Moreover, the potency of XLY is not simply afunction of targeting Gal11, because artificial activator pep-tides of similar size and even similar affinity for Gal11 do notactivate nearly as well as XLY (23).

The first indication that an interaction with the dimerizationdomain of Gal4 might contribute to functional potency of XLY

came from studies in yeast strains bearing different mutants ofGal11 (Fig. 2). The most revealing result was obtained withGal4(1–100)�P201 in a yeast strain with Gal11 containing twopoint mutations (Gal11DM) (Fig. 2H). One of the mutations(N342V) generates a surface that is targeted by Gal4dd,whereas the second mutation (T322K) abrogates binding toXLY. If in the context of Gal4(1–100)�P201, Gal4dd and XLY

each target Gal11 independent of one another, then in theGal11DM strain one would expect to observe activation levelscomparable with those obtained with Gal4dd alone in either the

FIG. 5. Location of the XLY-binding site within Gal4dd. A, LexA-P201 chimera does not activate transcription, and only upon inclusionof the complete dimerization domain of Gal4 (LexA�Gal (50–100)�P201) is robust transcriptional activation observed. The reporterand conditions used for the experiment are identical to those used in theFig. 3A experiments. The activity reported is an average of quadrupli-cate independent measurements, and in all cases the S.D. was less than15%. B, alanine substitution mutagenesis was performed on key struc-tural elements of the Gal4dd. The LexA-Gal4(50–100)�P201 chimerashown in Fig. 3A was used for this study, and transcriptional activationwas measured using the same reporter setup as the Fig. 3A experi-ments. The activity reported is an average of quadruplicate independ-ent measurements, and in all cases the S.D. was less than 15%. Severalmutations dramatically affected the activation properties of XLY(Gal4(93–100)�P201). C, alanine mutagenesis data from Figs. 3A and5A are summarized and compared with data from Ref. 28. The symbolsL1 and L2 are the loops, and �1–3 represents the three helices as shownin Fig. 1A and derived by Hidalgo et al. (28). The residues that werereplaced with alanine are in boldface type, and their effect on activationby XLY or Gal4dd in the respective strains (wild type Gal11 and Gal11P)are marked by a plus or minus symbol. Any mutation that negativelyimpacted functional activity more than 20%, implicating that residue inthe XLY�Gal4dd-binding interaction, is scored with a minus sign. Themutants that had an effect on one interaction but not the other arenumbered. �-gal, �-galactosidase.

FIG. 6. Rates of dissociation of XLY from Gal11 and Gal4dd.Fluorescein-labeled XLY was incubated with Gal11 (A) or Gal4dd (B)such that saturated binding was observed at 25 °C. The complexes thusobtained were rapidly mixed with a 1700-fold excess of unlabeled XLYpeptide, and the change in polarization due to displacement of thefluorescently labeled XLY over time was monitored. The data thusobtained were fit to a single exponential decay to determine the respec-tive off rates. The data displayed are the average of three individualexperiments.

Examining the Potency of an Artificial Transcription Factor 29695

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 8: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

Gal11P or Gal11DM strains. Instead, we observed a significantdecrease in activation, suggesting that XLY blocks the interac-tion between the N342V surface of Gal11DM and Gal4dd. Con-sistent with this interpretation, the XLY�Gal4dd-binding inter-action (5 �M) is qualitatively stronger than that between Gal4dd

and Gal11P (�18 �M by electrophoretic mobility shift assays).Additional mutagenesis experiments shown in Fig. 5 reveal

that the solvent-exposed hydrophobic loop 1 of Gal4dd is themost likely site of the XLY�Gal4dd interaction and, furthermore,that this interaction is essential for the transcriptional activa-tion function of XLY. The mutation of several loop 1 residuesabrogates XLY-mediated transcription. Moreover, Gal4dd, inwhich residue 69 of loop 1 is replaced by an alanine, shows nodetectable binding interaction with XLY (see supplemental Fig.3). Thus, the XLY�Gal4dd interaction is just as essential as theXLY�Gal11-binding interaction for XLY to function as an acti-vator. Consistent with this picture, residues 97–100 of Gal4dd

appear conformationally mobile in the NMR structure ofGal4(50–103), and if fully extended, these residues along withthe eight residues of P201 would span �36 Å, more thansufficient to reach the solvent-exposed hydrophobic loop 1(Fig. 1B).

These data also provide a clear explanation for how XLY

inhibits Gal4dd-mediated activation in yeast strains bearingGal11DM. Structure-activity studies of Gal4dd employing ala-nine-scanning mutagenesis as well as binding studies indicatethat the hydrophobic residues in loop 1 and the third helix arerequired for interaction with Gal11P (28). The mutagenesisresults of Fig. 5 show that the function of XLY is dependent onseveral residues of Gal4dd that are also necessary for interac-tion with Gal11P. The data also demonstrate that althoughthere is significant overlap, not all residues important for XLY

function are important for the Gal4dd�Gal11P interaction andvice versa. Thus, Gal4dd interacts with XLY and with Gal11Pusing two partially overlapping surfaces. More importantly,the results show that interaction with Gal4dd is essential forthe ability of XLY to function as a robust activator in vivo. Thefunctional importance of this XLY�Gal4dd interaction is furthersupported by the observation that a similar peptide(AHYYYPSE) that interacts with Gal11 nearly as well as XLY

(23) but does not bind to Gal4dd is not able to activate tran-scription to the same extent as XLY (Fig. 3F and supplementalFig. 2).

The �2-fold weaker interaction between XLY and Gal4dd

relative to Gal11�XLY (Fig. 3) provides some explanation as towhy the XLY�Gal4dd interaction does not inhibit XLY-mediatedtranscription. In addition, the kinetic data of Fig. 6 are sugges-tive of a mechanism where rapid interconversion betweenclosed and open states of XLY correlates with inactive andactive states of the activator (discussed below). In future ex-periments, the bound and unbound states of XLY will be char-acterized in more detail through stopped-flow experiments ofthe constructs bound to DNA to probe the effects of cooperativeDNA binding and interactions between adjacent DNA-boundactivators on this process.

By having identified this unexpected XLY�Gal4dd interaction,we are faced with the question as to why this interaction isnecessary for XLY to function as a potent activator. One possi-ble explanation is that the XLY�Gal4dd complex presents XLY inan appropriate conformation for the interaction with Gal11.However, several observations lead us to disfavor this expla-nation (although further experimentation will be required torigorously exclude it). Perhaps most important is the transfer-ability of this phenomenon. For example, XLY also functions asa potent transcriptional activator when fused to the DNA bind-ing domain of Pho4 (a helix-loop-helix transcription factor),

suggesting that the XLY�Gal4dd complex is not the active com-ponent of XLY-mediated transcriptional activation (19). Simi-larly, we have recently identified several other peptidic activa-tion domains unrelated to XLY that function more strongly(10–100-fold enhancement) when attached to Gal4(1–100) rel-ative to other DNA binding domains; each of these activationpeptides interacts with Gal4dd, and some do not require Gal11to function as transcriptional activators.2 Furthermore, thehydrophobic residues of XLY that are required for interactionwith Gal11 (tyrosine, for example) are also required for XLY tobind to Gal4dd. The simultaneous participation of these resi-dues in interactions with Gal4dd and Gal11, which would berequired if the closed complex is the transcriptionally activeform, is unlikely.

Taken together, the in vivo activation data and the kineticand thermodynamic measurements are more consistent with amodel in which XLY when bound to Gal4dd is inactive andunable to interact with Gal11 (Fig. 7). We propose that it is onlyin the open or exposed state that the peptide is functionallyactive, interacting with Gal11, and facilitating the cooperativeassembly of the transcriptional machinery at the promoter(Fig. 7).

The concealment and exposure of XLY are similar to a fre-quently observed feature of natural activation domains inwhich the hydrophobic residues that are vital for activatorfunction are masked by either intramolecular or intermolecularinteractions with other regulatory proteins (41–53). For exam-ple, in the structure of the p53 activation domain complexedwith the inhibitor protein MDM-2, the hydrophobic residuesthat are known to play a critical role in activation are foundburied in a hydrophobic cleft on the surface of MDM-2 (41, 42).In response to physiological signals, the residues are released,and the activation domain is capable of stimulating the expres-sion of target genes. There are also several examples of in-tramolecular concealment of activation domains characterizedin yeast such as Leu3 and Put3. In both of these cases, thefull-length activator protein is nonfunctional when tested invivo or in vitro (49–53). In the presence of a specific metaboliteor small molecule, however, the activating region is revealed,and the protein stimulates the expression of genes (49–53).

2 J. K. Lum, C. Y. Majmudar, A. Z. Ansari, and A. K. Mapp, unpub-lished results.

FIG. 7. A model depicting the role of the XLY�Gal4dd interactionin XLY-mediated transcriptional activation. We propose that inthe ”closed“ conformation, the XLY peptide is inactive, whereas in theopen conformation it can interact with Gal11 and likely other proteinsthat may lead to nonproductive complexation or even degradation. Anassumption inherent to this model is that the affinity of XLY for Gal4ddis higher than its affinity for chaperones, the proteolytic machinery, orother undefined factors that target exposed hydrophobic peptides (de-picted by the size of the arrows). The assumed difference in affinitywould potentially shield the activator to a sufficient degree from non-productive interactions in the cytoplasm. In the nucleus, the interactionbetween DNA-bound XLY and mediator-associated Gal11 would recruitthe transcriptional machinery to the promoter. The cooperative assem-bly of the transcriptional machinery at the promoter would furtherstabilize the XLY�Gal11 interaction.

Examining the Potency of an Artificial Transcription Factor29696

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 9: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

Thus, in these examples another surface within the activatorprotein has evolutionarily co-evolved to bind and conceal theactivating region in the absence of the relevant signal and toexpose it when instructed to do so.

One purpose of the concealment or masking of natural acti-vation domains is regulatory, preventing uncontrolled stimu-lation of transcription. In addition, the masking prevents un-productive binding as well as premature degradation of theactivator, thus contributing to the potency of activator func-tion. We propose that it is the latter role that the XLY�Gal4dd

interaction plays. The closed conformation renders XLY inac-cessible not just to Gal11 but also to other cellular proteins thatlead to nonproductive complexation or even degradation ofactivators. In other words, the XLY�Gal4dd interaction enablesXLY to escape cellular surveillance by protein-folding chaper-ones that bind to unstructured hydrophobic peptides and/or bythe machinery that degrades them (54–56). This model as-sumes that the affinity of XLY for Gal4dd is higher than itsaffinity for chaperones, the proteolytic machinery, or otherundefined cellular factors that interact with exposed hydropho-bic peptides in vivo. Thus, in the absence of the interaction withGal4dd, interactions with a variety of cellular proteins leadeither to premature destruction of XLY or to nonproductiveinteractions, both of which would negatively impact the overalleffectiveness of XLY as a transcriptional activation domain.Gal11 has a significant advantage over the other possible XLY-binding partners because the XLY�Gal11 interaction initiatescooperative assembly of the transcriptional machinery complexat the promoter, likely a significant driving force. A similarcooperative assembly with chaperones or the proteolytic ma-chinery is, in contrast, highly unlikely.

The transient exposure of hydrophobic residues may alsoplay a role in the second example of a potent artificial activator,Gal4dd. The NMR structure of Gal4dd indicates that the keyhydrophobic residues in the third helix, critical for interactionwith Gal11P, are protected from solvent by dimerization of twoGal4dd monomers (28). These studies also indicate that theinter-monomer interaction displayed by the third helix of thedimerization domain is weak, and if it is stabilized by theintroduction of additional residues of Gal4, the ability of Gal4dd

to function as an activator in vivo and to bind Gal11P in vitrois severely diminished (28, 57, 58). Thus, in this example ofpotent artificial activation as well, the hydrophobic residuesthat interact with a target are likely transiently occluded/exposed from bulk solvent and most cellular proteins.

In closing, this property of concealment and exposure isunlikely to be restricted to XLY and Gal4dd. We propose thatany hydrophobic surface that permits a balance of exposureand occlusion of hydrophobic activation domains, comprised ofbiopolymers, unnatural polymers, or small molecules, wouldenhance the potency of the artificial activator. In support ofthis, preliminary results suggest that the ability to transientlyinteract with such a surface increases the potency of unrelatedshort peptides that are otherwise weak activators up to 2orders of magnitude.2 Inclusion of this property may greatlyaid in the design of synthetic transcriptional activators thatfunction robustly in a cellular context.

Acknowledgments—We thank Prof. E. Craig for helpful discussionsand C. Luty, and R. Casey for technical assistance.

REFERENCES

1. Darnell, J. E., Jr. (2002) Nat. Rev. Cancer 2, 740–7492. Pandolfi, P. P. (2001) Oncogene 20, 3116–31273. Duncan, S. A., Navas, M. A., Dufort, D., Rossant, J. & Stoffel, M. (1998) Science

281, 692–695

4. Ansari, A. Z. & Mapp, A. K. (2002) Curr. Opin. Chem. Biol. 6, 765–7725. Mapp, A. K. (2003) Org. Biomol. Chem. 1, 2217–22206. Ptashne, M. & Gann, A. (2001) Genes & Signals, Cold Spring Harbor Labora-

tory Press, New York7. Ansari, A. Z. (2001) Curr. Org. Chem. 5, 903–9218. Mapp, A. K., Ansari, A. Z., Ptashne, M. & Dervan, P. B. (2000) Proc. Natl.

Acad. Sci. U. S. A. 97, 3930–39359. Beerli, R. R., Dreier, B. & Barbas, C. F. (2000) Proc. Natl. Acad. Sci. U. S. A.

97, 1495–150010. Kuznetsova, S., Ait-Si-Ali, S., Nagibneva, I., Troalen, F., Le Villain, J. P.,

Harel-Bellan, A. & Svinarchuk, F. (1999) Nucleic Acids Res. 27, 3995–400011. Liu, B., Han, Y., Ferdous, A., Corey, D. R. & Kodadek, T. (2003) Chem. Biol. 10,

909–91612. Beerli, R. R. & Barbas, C. F. (2002) Nat. Biotechnol. 20, 135–14113. Gerber, H. P., Seipel, K., Georgiev, O., Hofferer, M., Hug, M., Rusconi, S. &

Schaffner, W. (1994) Science 263, 808–81114. Blair, W. S., Bogerd, H. P., Madore, S. J. & Cullen, B. R. (1994) Mol. Cell. Biol.

14, 7226–723415. Ansari, A. Z., Mapp, A. K., Nguyen, D. H. & Dervan, P. B. (2001) Chem. Biol.

8, 583–59216. Nyanguile, O., Uesugi, M., Austin, D. J. & Verdine, G. L. (1997) Proc. Natl.

Acad. Sci. U. S. A. 94, 13402–1340617. Ma, J. & Ptashne, M. (1987) Cell 51, 113–11918. Giniger, E. & Ptashne, M. (1987) Nature 330, 670–67219. Lu, X. Y., Ansari, A. Z. & Ptashne, M. (2000) Proc. Natl. Acad. Sci. U. S. A. 97,

1988–199220. Lu, Z., Ansari, A. Z., Lu, X. Y., Ogirala, A. & Ptashne, M. (2002) Proc. Natl.

Acad. Sci. U. S. A. 99, 8591–859621. Frangioni, J. V., LaRiccia, L. M., Cantley, L. C. & Montminy, M. R. (2000) Nat.

Biotechnol. 18, 1080–108522. Han, Y. & Kodadek, T. (2000) J. Biol. Chem. 275, 14979–1498423. Wu, Z., Belanger, G., Brennan, B. B., Lum, J. K., Minter, A. R., Rowe, S. P.,

Plachetka, A., Majmudar, C. Y. & Mapp, A. K. (2003) J. Am. Chem. Soc.125, 12390–12391

24. Sengupta, D. J., Wickens, M. & Fields, S. (1999) RNA 5, 596–60125. Saha, S., Ansari, A. Z., Jarrell, K. A. & Ptashne, M. (2003) Nucleic Acids Res.

31, 1565–157026. Buskirk, A. R., Kehayova, P. D., Landrigan, A. & Liu, D. R. (2003) Chem. Biol.

10, 533–54027. Barberis, A. Pearlberg, J., Simkovich, N., Farrell, S., Reinagel, P., Bamdad, C.,

Sigal, G. & Ptashne, M. (1995) Cell 81, 359–36828. Hidalgo, P., Ansari, A. Z., Schmidt, P., Hare, B., Simkovich, N., Farrell, S.,

Shin, E. J., Ptashne, M. & Wagner, G. (2001) Genes Dev. 15, 1007–102029. Himmelfarb, H. J., Pearlberg, J., Last, D. H. & Ptashne, M. (1990) Cell 63,

1299–130930. Jeong, C. J., Yang, S.-H., Xie, Y., Zhang, L., Johnston, S. A. & Kodadek, T.

(2001) Biochemistry 40, 9421–942731. Lee, Y. C., Park, J. M., Min, S., Han, S. J. & Kim, Y. J. (1999) Mol. Cell. Biol.

19, 2967–297632. Park, J. M., Kim, H.-S., Han, S. J., Hwang, M.-S., Lee, Y. C. & Kim, Y.-J. (2000)

Mol. Cell. Biol. 20, 8709–871933. Kim, Y. J., Bjorklund, S., Li, Y., Sayre, M. H. & Kornberg, R. D. (1994) Cell 77,

599–60834. Koleske, A. J. & Young, R. A. (1994) Nature 368, 466–46935. Lee, T. I. & Young, R. A. (2000) Annu. Rev. Genet. 34, 77–13736. Myers, L. C. & Kornberg, R. D. (2000) Annu. Rev. Biochem. 69, 729–74937. Ansari, A. Z., Koh, S. S., Zaman, Z., Bongards, C., Lehming, N., Young, R. A.

& Ptashne, M. (2002) Proc. Natl. Acad. Sci. U. S. A. 99, 14706–1470938. Dotson, M. R., Yuan, C. X., Roeder, R. G., Myers, L. C., Gustafsson, C. M., Li,

Y., Kornberg, R. D. & Asturias, F. J. (2000) Proc. Natl. Acad. Sci. U. S. A.97, 14307–14310

39. Davis, J. A., Takagi, Y., Kornberg, R. D. & Asturias, F. A. (2002) Mol. Cell 10,409–415

40. Liu, Q., Krzewska, J., Liberek, K. & Craig, E. A. (2001) J. Biol. Chem. 276,6112–6118

41. Kussie, P. H., Gorina, S., Marechal, V., Elenbaas, B., Moreau, J., Levine, A. J.& Pavletich, N. P. (1996) Science 274, 948–953

42. Lee, C., Chang, J. H., Lee, H. S. & Cho, Y. (2002) Genes Dev. 16, 3199–321243. Ansari, A. Z., Reece, R. J. & Ptashne, M. (1998) Proc. Natl. Acad. Sci. U. S. A.

95, 13543–1354844. Sil, A. K., Alam, S., Xin, P., Ma, L., Morgan, M., Lebo, C., Woods, M. & Hopper,

J. (1999) Mol. Cell. Biol. 19, 7828–784045. Yano, K. & Fukasawa, T. (1997) Proc. Natl. Acad. Sci. U. S. A. 94, 1721–172646. Zenke, F. T., Engles, R., Vollenbroich, V., Meyer, J., Hollenberg, C. P. &

Breunig, K. D., (1996) Science 272, 1662–166547. Platt, A. & Reece, R. J. (1998) EMBO J. 17, 4086–409148. Lohr, D., Venkov, P. & Zlatnova, J. (1995) FASEB J. 9, 777–78749. Sze, J. Y., Woontner, M., Jaehning, J. A. & Kohlhaw, G. B. (1992) Science 258,

1143–114550. des Etages, S. G., Saxena, D., Huang, H. L., Falvey, D. A., Barber, D. &

Brandriss, M. C. (2001) Mol. Microbiol. 40, 890–89951. Sellick, C. A. & Reece, R. J. (2003) EMBO J. 22, 5147–515352. Wang, D., Zheng, F., Holmberg, S. & Kohlhaw, G. B. (1999) J. Biol. Chem. 274,

19017–1902453. des Etages, S. G., Falvey, D. A., Reece, R. J. & Brandriss, M. C. (1996) Genetics

142, 1069–108254. Horwich, A. L., Weber-Ban, E. U. & Finley, D. (1999) Proc. Natl. Acad. Sci.

U. S. A. 96, 11033–1104055. Lipford, J. R. & Deshaies, R. J. (2003) Nat. Cell Biol. 5, 845–85056. Muratani, M. & Tansey, W. P. (2003) Nat. Rev. Mol. Cell. Biol. 4, 192–20157. Farrell, S., Simkovich, N., Wu, Y., Barberis, A. & Ptashne, M. (1996) Genes

Examining the Potency of an Artificial Transcription Factor 29697

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 10: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

Dev. 10, 2359–236758. Shin, E. J. (1997) Honors Thesis in Biology, Ph.D. thesis, Harvard University,

Cambridge, MA59. Marmorstein, R., Carey, M., Ptashne, M. & Harrison, S. C. (1992) Nature 356,

408–41460. Adams, A., Gottschling, D. E., Kaiser, C. A. & Stearns, T. (1997) Methods in

Yeast Genetics, Cold Spring Harbor Laboratory Press, Cold Spring Harbor,NY

61. Chan, W. C. & White, P. D. (2000) Fmoc Solid-phase Peptide Synthesis: APractical Approach, Oxford University Press, New York

62. Kim, M., Saito, K., Furusaki, S., Sato, T., Sugo, T. & Ishigaki, I. (1991)J. Chromatogr. 585, 45–51

Examining the Potency of an Artificial Transcription Factor29698

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from

Page 11: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

0

20

40

60

80

100

Gal11WT

Gal4(1-100)LS201

DMGal11PMGal11Gal11P

Sup Figure 1Lu et al.

Ansari 101

Page 12: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

A

B

XLY + Gal4dd69A

KD 10 ± 5 µM

XLY + Gal4dd90A

0.051

0.053

0.049

0.047

0.045

0.043

0.0410 5.0 10.0 15.0 20.0

0.0 5.0[Gal4dd90A] (µM)

[Gal4dd69A] (µM)

25.015.0 35.0

Ani

sotr

opy

Ani

sotr

opy

Supp. Figure 3Lu et al.

0.046

0.048

0.044

0.042

0.040

0.038

0.036

0.034

Page 13: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

Supp. Figure 2 Lu et al.

Construct Gal11 Gal11P (N342V)

Gal11 (DM) (N342V, T322K)

Gal4(1-100) 2.7±0.2 242±16 260±20 Gal4(1-100)+AHYYYPSE 4.3±0.4 234±6 260±25

Transcriptional activity of Ga4(1-100)+AHYYYPSE in yeast strains bearing different alleles of Gal11. The constructs were expressed on a CEN4 plasmid under the control of a -actin promoter. The activation potential of each construct was measured in yeast strains expressing different alleles of Gal11. The -galactosidase activity is reported based on quadruplicate independent measurements and the error is calculated as the SDOM.

Page 14: THE J BIOLOGICAL C © 2005 by The American Society for ... Biol Chem. 2005.pdf · -D-galactopyranoside (final concentration 0.5 mM) for 5 h. The cell pellet was lysed using sonication,

Mapp and Aseem Z. AnsariJ. Nau, Chinmay Y. Majmudar, Anna K.Sarah E. Davis, Renee E. Metzler, Johnathan Zhen Lu, Steven P. Rowe, Brian B. Brennan,  Transcriptional ActivatorUnraveling the Mechanism of a PotentGenes: Structure and Regulation:

doi: 10.1074/jbc.M504895200 originally published online May 10, 20052005, 280:29689-29698.J. Biol. Chem. 

  10.1074/jbc.M504895200Access the most updated version of this article at doi:

  .JBC Affinity SitesFind articles, minireviews, Reflections and Classics on similar topics on the

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

Supplemental material:

  http://www.jbc.org/content/suppl/2005/05/12/M504895200.DC1.html

  http://www.jbc.org/content/280/33/29689.full.html#ref-list-1

This article cites 58 references, 32 of which can be accessed free at

at University of W

isconsin-Madison on D

ecember 4, 2015

http://ww

w.jbc.org/

Dow

nloaded from