Critical residues for histone acetylation
by Gcn5, functioning in Ada
and SAGA complexes, are also required
for transcriptional function in vivo
Lian Wang,
1
Lin Liu,
1
and Shelley L. Berger
2
Molecular Genetics Program, The Wistar Institute, Philadelphia, Pennsylvania 19104 USA
Several previously known transcription cofactors have been demonstrated in vitro recently to be histone
acetyltransferases and deacetyltransferases, suggesting that remodeling of chromatin through histone
acetylation plays a fundamental role in gene regulation. Clear evidence has not yet been obtained, however, to
demonstrate that histone acetylation is required for gene activation in vivo. In this study we performed an
alanine-scan mutagenesis through the HAT (h
istone acetyltransferase) domain identified previously by
deletion mapping in recombinant yeast Gcn5. We identified multiple substitution mutations that eliminated
completely Gcn5’s ability to potentiate transcriptional activation in vivo. Strikingly, each of these mutations
was also critical for free and nucleosomal histone acetylation by Gcn5 functioning within the native yeast
HAT complexes, Ada, and SAGA. Moreover, the growth phenotypes of these mutations as measured by colony
size and liquid growth assay closely tracked transcription and HAT activities. In contrast, mutations that did
not affect in vivo function of Gcn5 were able to acetylate histones. These data argue strongly that acetylation
is required for gene regulation by Gcn5 in vivo, and support previous arguments that nucleosomal histones are
among the physiological substrates of acetylation by Gcn5.
Gene expression is a highly regulated process at the level
of transcriptional activation. Recruitment of the tran-
scriptional basal machinery by activators and cofactors
has been shown to be an important component of gene
regulation in vitro (Zawel and Reinberg 1995) and in vivo
(Struhl 1995). Transcriptional activator proteins facili-
tate association of components of the basal machinery
with the TATA box and start site of transcription
(Triezenberg 1995). Under physiological conditions,
DNA is complexed with histones forming nucleosomes
that assemble into higher order chromatin (Wolffe 1992).
Chromatin assembly strongly inhibits transcription
(Owen-Hughes and Workman 1994; Paranjape et al.
1994), probably by blocking binding of transcription fac-
tors to their cognate DNA-binding sites. Therefore, in
vivo, transcription factors must overcome the DNA ac-
cess problem posed by nucleosomes. Moreover, genetic
manipulations of histones confirm their important role
in regulation of transcription (Grunstein 1990).
Both transcriptional activators and cofactors have
critical roles in altering the repressive chromatin struc-
ture to promote binding of basal factors. Activation do-
mains have been shown to be involved in remodeling
chromatin (Svaren et al. 1994), and nucleosome struc-
tural changes occur in the absence of transcription sug-
gesting that activators may induce transcription by caus-
ing chromatin changes before initiation (Fascher et al.
1993). Transcriptional cofactors also activate or repress
transcription by destablizing or stablizing the repressive
nucleosome structure (Kingston et al. 1996). For ex-
ample, the SWI/SNF family of protein complexes acti-
vates transcription by disrupting the nucleosome struc-
ture in an ATP-dependent manner to increase the bind-
ing of transcription factors (Cote et al. 1994; Peterson et
al. 1994).
Nucleosomes are composed of 146 bp of DNA
wrapped around the histone octamer, consisting of two
molecules each of the histones H2A, H2B, H3, and H4
(Wolffe 1992). Each core histone contains a globular do-
main and extended basic amino-terminal tail that stabi-
lizes histone octamer association with DNA through
ionic interactions. Covalent modifications of nucleo-
somal histones have been correlated with modulation of
chromatin structure and function (Bradbury 1992). These
modifications include phosphorylation, ubiquitination
and, in particular, acetylation of e amino groups of ly-
sines in the histone tails. Acetylation is associated with
both gene activation (hyperacetylation) and gene silenc-
ing (hypoacetylation) (for review, see Loidl 1994; Turner
and O’Neill 1995). Acetylation of histones increases the
affinity of transcription factors for nucleosomal DNA
(Lee et al. 1993; Vettese-Dadey et al. 1996).
1
These authors contributed equally to this work.
2
Corresponding author.
E-MAIL [email protected]; FAX (215) 898-0063.
640 GENES & DEVELOPMENT 12:640–653 © 1998 by Cold Spring Harbor Laboratory Press ISSN 0890-9369/98 $5.00; www.genesdev.org
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
The recent identification of histone acetyltransferase
(HAT) and histone deacetylase (HDAC) enzymes as pre-
viously recognized transcription cofactors (Pazin and Ka-
donaga 1997; Wade and Wolffe 1997) suggests a direct
mechanistic connection between the activity of tran-
scription factors and changes in chromatin structure that
precede initiation. The first nuclear HAT enzyme, p55,
was purified and cloned from Tetrahymena, and was
found to be closely related to yeast Gcn5 (Brownell et al.
1996), a known transcription coactivator/adaptor (Guar-
ante 1995). The ability of yeast Gcn5, and its binding
partner yeast Ada2, to associate with activation domains
(Barlev et al. 1995; Chiang et al. 1996; Silverman et al.
1994) and with the TATA-binding protein (TBP; Barlev
et al. 1995), suggests that acetylation of nucleosomal his-
tones through activation domain recruitment of Gcn5
may be a fundamental and novel feature of gene regula-
tion. Supporting this notion were subsequent revelations
that several other transcription coactivators/adaptors
possess intrinsic HAT activity, including additional
Gcn5 family members [hGcn5 (Wang et al. 1997; Yang et
al. 1996) and P/CAF (Yang et al. 1996)] as well as unre-
lated CBP/p300 (Bannister and Kouzarides 1996; Ogry-
zko et al. 1996) and TAF
II
250 (Mizzen et al. 1996).
Conversely, the first HDAC was identified (Taunton
et al. 1996) and strikingly, it also was a previously
known transcription factor, Rpd3 (Vidal and Gaber
1991), possessing transcriptional repression activity.
Several other transcriptional repressors have been shown
to be HDACs, including Sin3 (Hassig et al. 1997; Kadosh
and Struhl 1997; Laherty et al. 1997; Zhang et al. 1997)
and the corepressor N-CoR/SMRT (Alland et al. 1997;
Heinzel et al. 1997). Despite these provocative observa-
tions, however, direct evidence has not yet been ob-
tained that acetylation or deacetylation of histones has a
direct role in gene regulation.
Although recombinant proteins have been tested for
histone acetylation and deacetylation activity, recent
studies demonstrate that, in vivo, the HATs (Grant et al.
1997; Horiuchi et al. 1997; Saleh et al. 1997) and HDACs
(Rundlett and Grunstein 1996; Zhang et al. 1997) func-
tion as components of high-molecular-weight com-
plexes. For example, yeast Gcn5 is the catalytic HAT
subunit present in two complexes, termed Ada and
SAGA (S
pt, Ada, Gcn5 acetyltransferase; Grant et al.
1997). The Ada complex also contains Ada2 and Ada3,
which were similarly isolated as suppressors of growth
toxicity caused by overexpression of the potent chimeric
activator composed of the yeast Gal4 DNA-binding do-
main fused to the herpes virus VP16 activation domain
(Berger et al. 1992; Pin˜a et al. 1993; Marcus et al. 1994).
In addition to the Ada proteins, the SAGA complex con-
tains Spt7 (Gansheroff et al. 1995) and Spt20 (Roberts and
Winston 1996) proteins, which were isolated as suppres-
sors of Ty element insertions. Interestingly, Spt20/Ada5
was identified in both Ada and Spt genetic selections
(Marcus et al. 1996; Roberts and Winston 1996) and is
present in the SAGA complex (Grant et al. 1997). The
importance of determining acetylation activity within
these native complexes is illustrated by the observation
that recombinant yeast Gcn5 acetylates only free his-
tones, whereas, in association with Ada and SAGA com-
plexes, Gcn5 acetylates nucleosomal histones (Grant et
al. 1997). In addition, genetic studies demonstrate that
SAGA contains functions involved in transcriptional ac-
tivation distinct from those mediated by the Gcn5/
Ada2/Ada3 group (Roberts and Winston 1997), under-
scoring further the importance of testing HAT activity in
the context of native complexes.
The boundaries of the HAT domain in recombinant
yeast Gcn5 were mapped and the region was shown to be
required for Gcn5’s role in transcriptional activation in
vivo (Candau et al. 1997). This region of the Gcn5 family
members Tetrahymena p55 (Brownell et al. 1996), yeast
Gcn5 (Georgakopoulos and Thireos 1992), human Gcn5
(Candau et al. 1996), and human P/CAF (Yang et al.
1996) is remarkably conserved in primary sequence
(Brownell et al. 1996). Certain sequence motifs have
been identified in these HAT domains and are found in
other acetyltransferases (Reifsnyder et al. 1996; Neuwald
and Lansman 1997). Paradoxically, the other transcrip-
tional coactivators possessing intrinsic HAT activity,
CBP/p300 and TAF
II
250, display no obvious sequence
homology within the putative HAT domain compared
with the Gcn5 family (Bannister and Kouzarides 1996;
Mizzen et al. 1996; Ogryzko et al. 1996). In fact, recent
data demonstrates that, in addition to histones, p300
acetylates the transcriptional activator p53, and, surpris-
ingly, with nearly equal strength (Gu and Roeder 1997).
The data described above raise important questions—
what are the critical residues for acetylation of histone
substrates in Gcn5 and are these residues required for
Gcn5’s role in transcriptional activation? If so, this
would argue that first, acetylation is a critical and direct
determinant of gene regulation in vivo, and second, that
histones are likely to be a critical physiological sub-
strate. In the current study, we have undertaken alanine-
scan mutagenesis to address these issues. The role of
specific mutations was tested on Gcn5 functioning
within the context of native yeast Ada and SAGA com-
plexes. Our results demonstrate clearly that critical resi-
dues for histone acetylation in Gcn5 are also required for
its in vivo function in growth and as a transcription co-
factor. These data provide compelling evidence that
acetylation by Gcn5 is crucial for transcriptional activa-
tion in vivo.
Results
Construction of alanine-scan mutants
within the putative HAT domain of Gcn5
The four Gcn5 family members that have been identified
are yeast Gcn5, human Gcn5, human P/CAF, and Tet-
rahymena p55. Although the overall sequence identity
between the four proteins is <40%, the amino acid se-
quence in the putative HAT domain is 67% identical and
82% similar (Fig. 1A). The primary structural similarity
is reflected in conservation of function between the hu-
man and yeast HAT domains, whereas no other domain
Gcn5’s HAT activity required for function in vivo
GENES & DEVELOPMENT 641
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
of human Gcn5 is capable of function in yeast (Wang et
al. 1997). The HAT domain of Gcn5 can be divided into
four subdomains, I, II, III, and IV (Brownell et al. 1996),
based on exceptionally strong sequence identity between
all family members (Fig. 1A).
To identify critical amino acids for HAT enzymatic
activity and to determine whether HAT activity is cru-
cial for Gcn5’s role in transcription and growth, an ala-
nine-scan mutagenesis was performed. Three or four ad-
jacent amino acids were grouped (indicated as three or
four amino acid letter code) and mutagenized to alanine
(Fig. 1B) using site-directed mutagenesis. The mutagenic
oligonucleotides (Table 1) contained a PstI restriction
site allowing initial positive identification of each mu-
tant, which was confirmed by DNA sequencing. Thirty-
one substitution mutants were created and their pheno-
types studied.
Growth complementation by Gcn5 substitution
mutants in the gcn5
strain
The ability of the mutant GCN5 genes to complement
phenotypes caused by GCN5 disruption in yeast was
tested. Yeast lacking GCN5 grow poorly on minimal
synthetic media, and transformation of wild-type GCN5
completely complements the growth defect (Georgako-
poulos and Thireos 1992; Marcus et al. 1994). Comple-
mentation by the Gcn5 substitution mutants was tested
by transforming plasmids bearing each of the mutant
genes into the gcn5
strain and assessing growth using
colony size (Fig. 2A) and doubling time (Fig. 2B). The
mutant genes were expressed using the natural promoter
and terminator of GCN5 to obtain normal levels of pro-
tein expression. The majority of Gcn5 substitution mu-
tants (19 of the 30 mutants shown, as well as GGI lo-
cated between regions I and II) complemented growth as
well as wild-type Gcn5, as the yeast transformants
yielded colony sizes that were indistinguishable from
wild type (Fig. 2A). Several mutants (YIA, RGY, DNY,
KDY242, and TLM) displayed variable intermediate
colony sizes, ranging between that achieved by yeast
bearing wild-type GCN5 or a disruption of the gene (Fig.
2A). Six mutants (KQL, PKM, FAE, KDY196, AIGY, and
FKK) completely failed to complement the colony
growth defect in the gcn5
strain (Fig. 2A).
Relative growth phenotypes were quantitated by mea-
suring doubling times in liquid minimal growth media.
The Gcn5 mutants that exhibited the poorest growth in
the colony assay were analyzed, as well as several repre-
sentative mutants that exhibited wild-type growth. The
results paralleled the colony assays because the most
defective mutants (KQL, PKM, FAE, KDY196, AIGY, and
FKK) grew at essentially the rate of the gcn5
strain, and
the wild-type mutants grew comparably with GCN5
+
(Fig. 2B). Furthermore, the RGY mutant that showed in-
termediate-sized colonies was also intermediate in its
rate of doubling (Fig. 2B). Based on the two growth
complementation assays, we tentatively grouped the
mutants into three classes—those that were essentially
wild type, those that exhibited significant loss of growth
complementation, and those that were completely de-
fective.
The failure of Gcn5 mutants to complement growth
may be caused by defective function of the Gcn5 substi-
tution mutants, or alternatively, by structural defects in
the protein causing either instability or inability to form
requisite protein–protein interactions in vivo. To address
the question of protein stability, the expression levels of
the mutant Gcn5 proteins in yeast were determined rela-
tive to wild-type Gcn5. Gcn5 was expressed under con-
trol of the galactose-inducible GAL1-10 promoter and
intracellular protein expression levels were determined
by Western analysis (Fig. 3, top). Although there was
variation in protein levels, all Gcn5 mutants were pre-
sent at levels comparable with wild-type Gcn5, and, in
particular, all mutants displaying partial or complete de-
fects in the growth assay were detectable in vivo (Fig. 3,
top).
To determine the structural integrity of the mutant
proteins in vivo, we tested interaction with Ada2, as our
previous results demonstrated that Ada2 interaction is
critical for Gcn5 function in vivo (Candau et al. 1997).
Figure 1. Amino acid sequence within
the putative HAT domains of the Gcn5
family. (A) Sequence comparison between
the putative HAT domains of yeast Gcn5,
human P/CAF, human Gcn5 and Tetrahy-
mena p55. The protein sequence within
the putative HAT domains are shown by
single amino acid letter-code. Numbers at
left indicate amino acid residues. The
highly conserved regions I–IV (Brownell et
al. 1996) are indicated at the top. Solid
boxes indicate amino acid identity and
shaded boxes indicate similarity. (B) Yeast
Gcn5 substitution mutants created in the
study. The substitution mutants are
named according to the group of amino ac-
ids mutated to alanine, each of which is
indicated by brackets.
Wang et al.
642 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
Yeast ADA2 was fused to glutathione-S-transferase
(GST) followed by cotransformation into yeast of expres-
sion plasmids bearing GST–ADA2 and GCN5. Yeast ex-
tracts were incubated with GST beads and the binding of
wild-type Gcn5 or Gcn5 mutants to GST–Ada2 was de-
termined by immunoblot. The Gcn5 mutants were able
to bind to Ada2 and importantly, mutants defective for
growth complementation also interacted with Ada2 (Fig.
3, bottom). Overall, the yeast expression experiments
and Ada2-interaction results indicate that the defective
phenotypes of the substitution mutants were not due to
global structural defects.
Transcriptional activation in the presence of Gcn5
mutants in the gcn5
strain
The ability of Gcn5 substitution mutants to support
transcriptional activation in the gcn5
strain was tested.
Full transcriptional potency by certain transcriptional
activators requires Gcn5 (Georgakopoulos and Thireos
1992; Marcus et al. 1994). Two chimeric activators were
tested. The first was the yeast Gal4 DNA-binding do-
main fused to the potent activation domain of herpes
simplex virus activator VP16 (Gal4–VP16), and the sec-
ond was the bacterial LexA DNA binding domain fused
to the activation domain from the yeast activator Gcn4
(LexA–Gcn4). The expression reporters were bacterial
LacZ driven by yeast Gal4 DNA-binding sites or bacte-
rial LexA DNA-binding sites, and transcriptional po-
tency was determined by quantitative b-galactosidase as-
says.
Both Gal4–VP16 and LexA–Gcn4 transcriptional activ-
ity were reduced in the GCN5 deletion strain to 9% of
wild type (Fig. 4). The Gcn5 substitution mutants di-
vided into three classes. Most Gcn5 mutants were ca-
pable of potentiating Gal4–VP16 or LexA–Gcn4 activity
to approximately the level observed in the GCN5
+
strain
(Fig. 4). Several mutants that displayed partial comple-
mentation in the growth assay also showed an interme-
diate level of activity in potentiating transcriptional ac-
tivation by both activators (Fig. 4). The six mutants
(KQL, PKM, FAE, KDY196, AIGY, and FKK) that were
unable to complement the growth defect of gcn5
, also
failed to complement gcn5
function in the transcription
assay for either activator (Fig. 4).
Therefore, the majority of the Gcn5 substitution mu-
tants had similar phenotypes exhibited by wild-type
Gcn5 in growth and transcription assays (hereafter re-
Table 1. Sequences of mutagenic oligonucleotides
Name
a
Amino acid
b
Sequence of oligonucleotides
LKN 120 GATGGTCCTAACTGGAGCTGCAGCCATTTTTCAAAAGCAATTACC
IFQ 123 CTAACTGGATTAAAAAACGCTGCAGCAAAGCAATTACCAAAAATG
KQL 126 GGATTAAAAAACATTTTTCAAGCTGCAGCACCAAAAATGCCC
PKM 129 GGATTAAAAAACATTTTTCAAAAGCAATTAGCTGCAGCGCCCAAAGAATACATTGCC
PKE 132 CAATTACCAAAAATGGCTGCAGCATACATTGCCAGGTTAG
RLV 135 ATGCCCAAAGAAGCTGCAGCCAGGTTAGTCTATG
YIA 138 CAAAGAATACATTGCCGCTGCAGCCTATGATCGAAGTCATC
YDR 141 GCCAGGTTAGTCGCTGCAGCAAGTCATCTTTCCATG
SHL 144 GTTAGTCTATGATCGAGCTGCAGCTTCCATGGCTGTCATTAGG
GGI 159 CGATAAGAGAGAAGCCGCAGCAATTGTTTTCTGTGCC
FAE 171 CCATTGACTGTCGTAGCTGCAGCAACATATCGACCTTTCG
IVF 174 GAATTCGCAGAAGCTGCAGCCTGTGCCATCAGTTCG
CAI 177 CGCAGAAATTGTTTTCGCTGCAGCCAGTTCGACGGAACAG
EQV 183 GCCATCAGTTCGACGGCTGCAGCACGCGGTTATGGTGCG
RGY 186 GGAACAGGTAGCTGCAGCTGGTGCGCATC
GAHL 189 GGTACGCGGTTATGCTGCAGCTGCAATGAATCACTTAAAAGAC
MNHL 193 GGTTATGGTGCGCATCTAGCGGCTGCAGCAAAAGACTATGTTAG
KDY196 196 GAATCACTTAGCTGCAGCTGTTAGAAATACC
NIK 205 GTTAGAAATACCTCGGCTGCAGCATATTTTTTGACATATGC
YFL 208 CCTCGAACATAAAAGCTGCAGCGACATATGCAGATAATTAC
TYA 211 CGAACATAAAATATTTTTTGGCTGCAGCAGATAATTACGC
DNY 214 TATTTTTTGACATATGCAGCTGCAGCCGCTATTGGATACTTT
AIGY 217 GCAGATAATTACGCTGCTGCAGCCTTTAAAAAGCAAGGC
FKK 221 TACGCTATTGGATACGCTGCAGCGCAAGGCTTCACTAA
QGF 224 CGCTATTGGATACTTTAAAAAGGCTGCAGCCACTAAAGAAATCACGTTGG
TKEI 227 AAAAAGCAAGGCTTCGCTGCAGCAGCCACGTTGGATAAAAG
GYI 239 GTATATGGATGGCTGCAGCTAAAGATTATGAAGG
KDY242 242 GGATGGGATATATTGCTGCAGCTGAAGGTGGT
EGG 245 GGATATATTAAAGATTATGCTGCAGCTACGCTGATGCAATG
TLM 248 GATTATGAAGGTGGTGCTGCAGCGCAATGTTCTATGTTACC
PRI 256 GCAATGTTCTATGTTAGCTGCAGCACGATATTTGGACGC
a
The name of each mutant represents the amino acids mutagenized.
b
The position of the first amino acid mutagenized in each mutant is indicated.
Gcn5’s HAT activity required for function in vivo
GENES & DEVELOPMENT 643
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
ferred to as wild-type mutants; light gray bars in Fig. 4).
The assays identified six Gcn5 mutants (KQL, PKM,
FAE, KDY196, AIGY, and FKK) that profoundly reduced
Gcn5 function in vivo (‘‘defective’’ mutants; black bars
in Fig. 4). In addition, several mutants (YIA, RGY, DNY,
KDY242, and TLM) displayed various degrees of partial
function and are likely to represent important, but not
crucial, residues for Gcn5 activity (‘‘intermediate’’ mu-
tants; dark gray bars in Fig. 4).
Dominant-negative effect of Gcn5 mutants
Mutations in catalytic site residues may cause mutant
proteins to exert dominant-negative effects when over-
expressed in vivo. Dominant-negative phenotypes are
likely caused by the mutant enzyme competing with the
wild-type counterpart for association with other proteins
or assembly into protein complexes. To determine
whether the phenotypically defective Gcn5 mutants ex-
ert a dominant defective growth effect in vivo, each was
overexpressed in yeast containing normal levels of wild-
type Gcn5.
Overexpression of wild-type Gcn5 itself caused yeast
to grow slowly compared with transformed vector (Fig.
5), perhaps attributable to sequestration of associated
proteins (Gill and Ptashne 1988). Overexpression of the
wild-type Gcn5 mutants showed similar phenotypes as
wild-type Gcn5 (data not shown) and were therefore
judged not to be dominant-negative. Overexpression
of most defective Gcn5 mutants caused even stronger
growth inhibition relative to wild-type Gcn5 (Fig. 5).
Figure 3. Stability and Ada2 interaction of Gcn5 substitution
mutants in yeast. Wild-type GCN5, mutant GCN5, or vector
alone were cotransformed with GST–ADA2, and proteins were
induced using the GAL1-10 promoter. (Top) Gcn5 protein levels
in the yeast extracts were determined by Western blot analysis
using Gcn5 antisera. (Bottom left) The yeast extracts were in-
cubated with GST beads to allow GST–Ada2 to bind. Wild-type
or mutant Gcn5 bound to GST–Ada2 was determined by West-
ern analysis using Gcn5 antiserum. (Bottom right) Co-expres-
sion of wild-type Gcn5 with GST alone was used as a negative
control.
Figure 2. Growth phenotypes of GCN5 substitution
mutants in the gcn5
strain. Wild-type GCN5, each of
the GCN5 mutants, or vector alone were introduced
into gcn5
strain. (A) Colony growth assay. Transfor-
mants were streaked onto synthetic minimal medium.
Colony growth was assessed after 2 days of growth at
30°C. (B) Doubling time of the Gcn5 substitution mu-
tants. Aliquots of cultures in liquid synthetic minimal
medium were taken every 2 hr and the A
600
of the cul-
tures was measured. The generation time was calculated
as the OD doubling time during exponential growth.
Wang et al.
644 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
This was true both of mutants that were completely
(KQL, PKM, FAE, AIGY, and FKK) or partially (YIA,
RGY, and KDY242) defective (Fig. 5). Two defective mu-
tants (completely defective KDY196 and partially defec-
tive TLM) exhibited a similar phenotype as wild-type
Gcn5 (Fig. 5) and therefore were not dominant-negative.
The mutants are likely to exert their dominant-negative
effects by incorporation into normal protein complexes
(Grant et al. 1997) because first, they are vastly overex-
pressed relative to wild-type Gcn5, and second, they in-
teract normally with Ada2 (Fig. 3) and within native
complexes (Fig. 6B,C) in the gcn5
strain where wild-
type Gcn5 is not present.
Therefore, two lines of evidence indicate that the
Figure 5. Dominant-negative effects of Gcn5 HAT mutants. Wild-type GCN5, the GCN5 mutants, or vector alone were transformed
into yeast strain bearing wild-type GCN5. The transformants were restreaked onto galactose plates to induce expression of trans-
formed Gcn5 proteins and grown at 25°C for 6 days.
Figure 4. Transcriptional activation in
the presence of Gcn5 substitution mu-
tants. GAL4–VP16 (left panel) and LexA–
GCN4 (right panel) were cotransformed
with wild-type GCN5, each of the GCN5
mutants or vector alone, along with the ap-
propriate LacZ reporters. b-Gal activities
of the mutants are shown as a percentage
of wild-type Gcn5, which was set at 100%.
(Solid bars) Mutants that were defective in
the growth and transcription assays; (dark
gray bars) mutants having intermediate
phenotypes; (light gray bars) mutants hav-
ing activity similar to wild-type Gcn5. Er-
ror bars represent the standard error about
the mean from three independent experi-
ments.
Gcn5’s HAT activity required for function in vivo
GENES & DEVELOPMENT 645
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
Gcn5 mutants displaying defects in vivo are functional.
First, all of the defective mutants possess structural in-
tegrity as manifested by their ability to interact with
Ada2 (Fig. 4), and second, the majority of the defective
mutants exhibit dominant negativity in vivo (Fig. 5).
Histone acetylation by Gcn5 substitution mutants
within Ada and SAGA complexes
Previous studies have tested HAT activity of recombi-
nant proteins. The HATs and HDACs are active within
high-molecular-weight complexes, however, which in-
fluences their specificity (Rundlett and Grunstein 1996;
Grant et al. 1997). To directly assay whether the defec-
tive phenotypes of the substitution mutants were caused
by loss of ability to acetylate histones, HAT activity of
the Gcn5 substitution mutants was tested in the context
of high-molecular-weight native HAT complexes (Grant
et al. 1997). Two high-molecular-weight native yeast
complexes have been identified. Both contain Gcn5 as
the HAT catalytic subunit, and specifically acetylate his-
tone H3 and, to a lesser extent, histone H2B. The smaller
complex (800 kD) was named Ada and the larger one (2
MD) SAGA, based on the identity of their constituent
proteins (Grant et al. 1997). We partially purified Ada
and SAGA complexes from yeast expressing the Gcn5
substitution mutants, to determine whether the muta-
tions affected the HAT activity within these physiologi-
cal HAT complexes.
Wild-type Gcn5 or the substitution mutants were
transformed into the gcn5
strain using vectors to ex-
press normal levels of protein. Whole-cell extracts were
prepared from 26 out of the 31 mutants created and were
fractionated through Ni
2+
agarose and MonoQ ion ex-
change resins (Grant et al. 1997). HAT activities were
determined on free histone substrates for 22 out of these
26 extracts, which were determined by Western analysis
to have nearly normal Gcn5/Ada2 protein levels in both
the Ada and SAGA complexes (see below). These in-
cluded 11 mutants that were partially or completely de-
fective in growth (Fig. 2) and transcription (Fig. 4), as well
as 11 wild-type mutants.
The HAT activity of the peak SAGA fraction from
wild-type or mutant Gcn5 was then determined by liq-
uid HAT assay quantitation of [
3
H]acetyl group incorpo-
ration into histones using scintillation counting and vi-
sualization of histones by PAGE and fluorography.
SAGA fractions derived from strains bearing wild-type
mutant Gcn5 were able to acetylate histones, as shown
by the examples of LKN, PKE, YDR, GAHL, and GYI
(Fig. 6A). Four of these had levels as high as, or nearly as
high as, wild-type Gcn5 (LKN, PKE, YDR, and GYI) as
quantitated by scintillation counting (Fig. 6A, top) and
confirmed by visualization of [
3
H]acetyl incorporation
into histone H3 (Fig. 6A, middle). These mutant com-
plexes displayed Gcn5 and Ada2 protein levels that were
comparable with wild-type Gcn5 (Fig. 6B, top). The re-
maining wild-type mutant (GAHL) possessed >50% of
wild-type activity (Fig. 6A) and was found to be some-
what reduced in protein level compared with wild-type
Gcn5 and the other wild-type mutants (Fig. 6B, top; data
not shown). In addition, visualization of [
3
H]acetyl in-
corporation into histone H3 catalyzed by SAGA bearing
GAHL indicated clearly higher activity than the defec-
tive Gcn5 mutants (Fig. 6A, middle).
The six severely defective mutants (KQL, PKM, FAE,
KDY196, AIGY, and FKK), as well as partially defective
RGY, displayed greatly reduced ability to acetylate his-
tones, using both the quantitative assay as well as his-
tone visualization (Fig. 6A, top and middle). Five of the
extracts showed background levels of HAT activity
(25% of wild-type level). The level of acetylation by
SAGA prepared from the defective FKK mutant was 45%
of wild type in the quantitative assay, but was nearly as
low as background levels by [
3
H]-histone fluorography.
(As is obvious in Fig. 6, we consistently observe that
fluorography is a more sensitive measure of acetylation
than scintillation quantitation.) The protein levels of
both Ada2 and the defective mutant Gcn5 proteins were
variable, but the levels were similar to those in the strain
expressing wild-type Gcn5 (Fig. 6B, bottom). Therefore,
consistent with the stability assay (Fig. 3), the defects of
these mutants were not attributable to lack of protein
expression, or importantly, inability to be incorporated
into the SAGA complex.
We then determined the HAT activity of the Ada com-
plexes using the identical set of mutant Gcn5 proteins as
used for assay of the SAGA complexes. In the case of the
Ada complex only [
3
H]histone-H3 fluorography was used
to determine HAT activity, as a strong overlapping
Gcn5-independent histone H4 acetylation activity
(Grant et al. 1997; see Fig. 7) obscured quantitation by
scintillation counting of the Gcn5-dependent histone H3
activity. As observed for the SAGA complex, each of the
wild-type mutants displayed activities comparable with
wild-type Gcn5, whereas HAT activity by the defective
mutants was not detectable (Fig. 6A, bottom). Western
analysis of Ada2 and Gcn5 in the Ada complex (Fig. 6C)
of the wild-type (Fig. 6C, top) and defective (Fig. 6C, bot-
tom) Gcn5 mutants indicated comparable levels of pro-
tein with wild-type Gcn5. Note that the Ada complex
containing the GAHL mutant eluted primarily in frac-
tion 20 from the MonoQ column (Fig. 6C, top left),
whereas the remainder of the wild-type mutants eluted
in a broader peak covering fractions from 20 (Fig. 6C, top
left) to 24 (Fig. 6C, top right).
Overall, the results demonstrate a clear correlation be-
tween complementation of function by the Gcn5 substi-
tution mutants in the gcn5
strain (Figs. 2 and 4) and
acetylation of histones (Fig. 6). Importantly, HAT activ-
ity by the mutant Gcn5 proteins was similar in both the
Ada and SAGA complexes, such that mutants that pos-
sessed wild-type levels of activity in the Ada complex
were also competent in the SAGA complex and mutants
that were defective in Ada were also defective in SAGA.
Acetylation of nucleosomal histones by Gcn5
substitution mutants
Both the SAGA and Ada complexes acetylate nucleo-
Wang et al.
646 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
somal histones. Because the level of nucleosomal acety-
lation compared with free histone acetylation is mark-
edly inhibited in the SAGA complex (D. Sterner, P.
Grant, J. Workman, and S.L. Berger, unpubl.), we tested
the ability of the Gcn5 substitution mutants to acetylate
nucleosomal histones in the context of the Ada complex.
In addition, the Gcn5-dependent H3/H2B HAT activity
in the Ada complex overlaps a Gcn5-independent H4/
H2A HAT activity (Grant et al. 1997; Fig. 7), which pro-
vides an internal control for variation in extract prepa-
ration.
Fractions between 18 and 30 were tested for HAT
activity on nucleosomal substrates using a liquid
HAT assay and incorporation of the [
3
H]acetyl group
into histones was visualized by PAGE and fluorogra-
phy. The peak of Gcn5-dependent H3/H2B activity in
the wild-type extract eluted between fractions 20–
24, whereas the H4/H2A activity peaked at fraction
24–26 (Fig. 7, top left). As observed previously (Grant
et al. 1997), in extracts prepared from the gcn5
strain, the H4/H2A HAT activity remained intact,
whereas the H3/H2B activity disappeared (Fig. 7,
left).
Strikingly, five mutants that failed to complement
loss of Gcn5 in vivo (KQL, PKM, FAE, AIGY, and FKK)
and were defective in acetylation of free histones in the
SAGA and Ada complexes also were completely defec-
tive in acetylation of nucleosomal histones in the Ada
complex (Fig 7, right). The remaining defective mutant
(KDY196) had severely reduced activity, which was
barely above background levels present in the gcn5
ex-
tract (Fig. 7, right). In addition, one of the partially de-
fective mutants in the in vivo assays (RGY) showed se-
verely impaired activity in the nucleosome acetylation
assay (Fig. 7, right), as also observed in the free histone
assay (Fig. 6A). In contrast, the Gcn5-independent H4/
H2A acetylation activity was not affected by these Gcn5
substitution mutations (Fig. 7, right). As described in the
previous section, the presence of Gcn5 and Ada2 in the
Ada complex was confirmed for all mutants by Western
analysis (Fig 6C).
In contrast, mutants that displayed wild-type pheno-
Figure 6. Free histone HAT activity of Gcn5 substitution mutants in SAGA and Ada complexes. (A) SAGA and Ada complexes
prepared from Gcn5 mutants displaying wild-type activity in functional assays (light gray bars in Fig. 4) and Gcn5-defective mutants
(solid bars in Fig. 4) were tested for HAT activity. SAGA and Ada complexes of each mutants containing peak Gcn5 and Ada2 proteins
detected by Western blots (B,C) were used. Complexes were incubated with free core histones and [
3
H]acetyl CoA. Reaction mixtures
were then subjected to liquid scintillation accounting (top) as well as SDS-PAGE and fluorography (middle and bottom). The activity
from SAGA complexes (middle) and Ada complexes (bottom) are indicated by arrows. Quantitation of corresponding scintillation
assays for SAGA is shown as a percentage of wild-type Gcn5. Error bars represent the standard error about the mean from three
independent experiments. Western blots of Gcn5 and Ada2 proteins from SAGA (B) and Ada (C) complexes. Peak fractions from SAGA
and Ada complexes were detected for the presence of Gcn5 and Ada2 using immunoblot analysis using Gcn5 and Ada2 antisera.
Gcn5’s HAT activity required for function in vivo
GENES & DEVELOPMENT 647
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
types in complementation of growth and transcription in
the gcn5
strain (LKN, PKE, YDR, GAHL, and GYI), were
competent for acetylation of nucleosomal H3/H2B with
varying efficiencies relative to wild-type Gcn5 (Fig. 7,
left). Significantly, all wild-type Gcn5 mutants were
relatively strong in nucleosomal H3/H2B acetylation
compared with all defective mutants. Therefore, the data
demonstrate that mutations that are most deleterious for
in vivo function of Gcn5 similarly completely impair its
histone acetylation function as a component of native
SAGA and Ada complexes.
Discussion
Several transcriptional cofactors have been revealed re-
cently to possess histone acetyltransferase activity in
vitro. These include members of the Gcn5 family as well
as the unrelated coactivators p300/CBP and TAF
II
250.
The emergence of this new class of transcription coacti-
vators functioning as HATs has raised important ques-
tions. First, do these proteins function as acetyltransfer-
ases in vivo, second, are nucleosomal histones their
natural substrates, and, finally, is acetyltransferase ac-
tivity required for transcription function? To address
these issues, we systematically created substitution mu-
tants in the putative HAT domain of yeast Gcn5, and
compared their effects on growth and transcriptional ac-
tivation in vivo with histone acetylation in vitro. If
acetylation of histones by Gcn5 is critical, then there
should be strong correlation between HAT activity and
in vivo function.
Correlation between in vivo function of Gcn5
and in vitro histone acetylation
We identified six substitution mutations (KQL, PKM,
FAE, KDY196, AIGY, and FKK) in Gcn5 that were highly
deleterious to Gcn5 function in vivo. Strikingly, five
were inert for acetylation of histones, either as free or
nucleosomal substrates, and the sixth mutant (KDY196)
had very low activity. Importantly, other mutants that
were intermediate in function in vivo (YIA, RGY, DNY,
KDY242, and TLM) were also impaired significantly in
both free histone and nucleosomal histone acetylation
(RGY acetylation is shown in Figs. 6 and 7; others not
shown). Finally, substitution mutants that displayed
wild-type function in vivo, displayed relatively high lev-
els of acetylation in vitro (five examples are shown in
Figs. 6 and 7; complexes from six others were prepared
and were found to possess significant levels of HAT ac-
tivity). Therefore, these data (summarized in Fig. 8) dem-
onstrate a clear correlation between residues within
Gcn5 that are critical for acetylation of histone sub-
strates, and residues that are absolutely required for
growth and transcriptional activity of Gcn5.
It is important to stress that we analyzed histone
acetylation by Gcn5 within the context of native yeast
SAGA and Ada complexes. Gcn5, as well as other HATs
(Chen et al. 1997) and HDACs (Rundlett and Grunstein
1996), exists in protein complexes containing multiple
components that, in the case of Gcn5, alter substrate
specificity in two ways. Recombinant yeast Gcn5 acety-
lates only free histones and primarily histone H3 (his-
tone H4 is acetylated to a far lesser degree) (Brownell et
Figure 7. Nucleosomal HAT activities of
Gcn5 substitution mutants in Ada com-
plexes. The same wild-type mutants and
defective mutants used in the free histone
assays (Fig. 6) were subjected to nucleo-
somal HAT assays using Ada complexes.
Even-numbered Mono Q column fractions
from Ada complex of each mutant were in-
cubated with oligonucleosome cores and
[
3
H]acetyl CoA. Reactions were loaded
onto SDS-PAGE and the gels were analyzed
by fluorography. (Left) Wild-type Gcn5,
vector alone, and wild-type Gcn5 substitu-
tion mutants; (right) defective Gcn5 mu-
tants. Positions of core histones on the gel
are indicated by arrows. Numbers at the
top are the Mono Q column fractions. The
H3/H2B nucleosomal activity is Gcn5-de-
pendent, and the H4/H2A nucleosomal
HAT activity is Gcn5-independent (Grant
et al. 1997).
Wang et al.
648 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
al. 1996). Gcn5 in native yeast SAGA and Ada complexes
acetylates both free and nucleosomal histones and tar-
gets both H3 and H2B (Grant et al. 1997). The altered
specificity of Gcn5 in complexes, as opposed to recom-
binant form, suggests that physiologically relevant
acetylation activity is best evaluated in the context of
these native HAT complexes, as we have done in the
current study.
Mutations that impair transcription
similarly impair growth
Gcn5 mutants were tested for their effects on transcrip-
tion using two heterologous model activators, Gal4–
VP16 and LexA–Gcn4. We showed previously that Ada2,
which is present in the Ada and SAGA complexes and is
required for Gcn5 function in vivo, binds to both of these
activation domains (Barlev et al. 1995). The Gcn5 sub-
stitution mutants showed very similar phenotypes in the
transcription assays, such that the wild-type, intermedi-
ate, and deleterious mutants displayed parallel effects for
function by each activator. In addition, the natural HIS3
and PHO5 promoters also display dependency on the de-
fective mutants, and have normal function in the pres-
ence of wild-type mutants (R. Bolotserkovskaya, N. Bar-
lev, and S.L. Berger, unpubl.). Therefore, the effects of the
substitution mutations that reduce histone acetylation
have similar effects on both synthetic model promoters
and natural yeast promoters, arguing that the effects are
generally consistent for activators that require Gcn5 for
function.
Moreover, we found a striking correlation between the
effect of the substitution mutations on Gcn5 function in
growth and in transcriptional activation assays, such
that colony size and doubling time of the wild-type, in-
termediate, and deleterious mutants mirrored transcrip-
tion function. The consistency between growth and tran-
scription suggests that results obtained using the model
activators Gal4–VP16 and LexA–Gcn4 reflect natural de-
pendency of yeast genes transcription on Gcn5.
Taken together with previous data regarding function
and biochemical interactions of the Ada, Gcn5, and Spt
proteins (Grant et al. 1997; Roberts and Winston 1997),
the data support the model that Gcn5 functions within
Ada and/or SAGA complexes to acetylate nucleosomal
histones in promoter regions of dependent genes. Ada
and SAGA may be recruited to promoters via interac-
tions of Ada2 or Gcn5 with activation domains (Silver-
man et al. 1994; Barlev et al. 1995; Chiang et al. 1996),
which may position the HAT domain to acetylate
nucleosomes proximal to the TATA box. Spt proteins
may have distinct roles in the SAGA complex, either for
complex integrity (Spt7 and Spt20/Ada5; Grant et al.
1997) or recruitment of TBP and associated proteins
(Spt7 and Spt8; Eisenmann et al. 1992, 1994). Based on
the strong correlation we observe between residues
within Gcn5 required for HAT activity and function, we
conclude that acetylation by Ada and/or SAGA appears
to be a critical in vivo function.
Residues important for function within the Gen5
HAT domain
We identified critical residues for function in all four
Figure 8. Comparison of mutant Gcn5
phenotypes in acetylation vs. transcrip-
tion. Quantitative data are taken from Fig.
6A (top) and Fig. 4 (the Gal4–VP16 and
LexA–Gcn4 data were averaged). Acetyla-
tion of free histones by the SAGA complex
is shown at left and transcriptional activa-
tion by the chimeric activators is shown at
right.
Gcn5’s HAT activity required for function in vivo
GENES & DEVELOPMENT 649
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
conserved subdomains of Gcn5. Mutations in regions I
(KQL and PKM), II (FAE and RGY), and III (AIGY and
FKK) were most debilitating for histone acetylation
within the Ada and SAGA complexes. Previous evolu-
tionary comparisons have identified conserved se-
quences in the putative HAT domain between the Gcn5
subfamily of acetyltransferases and many additional pro-
teins, leading to speculation that these proteins form an
extensive superfamily (Reifsnyder et al. 1996; Neuwald
and Lansman 1997). In particular, there is very strong
conservation of conserved Gcn5 regions II and III and
members of the superfamily (Neuwald and Lansman
1997). Because the apparent commonality within the su-
perfamily is acetyl-CoA binding, this region may bind
acetyl-CoA. In particular, RGY, AIGY, and FKK show
very strong conservation across the superfamily. There-
fore, our results that these mutations are extremely del-
eterious to histone acetylation are consistent with a role
in acetyl-CoA binding. Whereas AIGY and FKK are com-
pletely debilitating to function of Gcn5 in vivo, it was
surprising that RGY mutant was not affected as dramati-
cally. Because the two glycines in RGYG show the most
strong conservation in this short sequence (Neuwald and
Lansman 1997), it may be that the alanine substitution
preserves some function. The dominant negativity of all
three of these mutants RGY, AIGY, and FKK supports
the interpretation of a defect in acetyl-CoA binding, but
retention of other critical protein interactions.
Interestingly, the FAE mutant, which occurs at the
amino-terminal border of region II, is one of the most
dysfunctional mutants, but is not strongly conserved in
the superfamily. The biochemical function of the FAE
residues will be important to determine. In addition, an-
other region of conservation in the superfamily, located
between regions I and II, contains the conserved GGI
(Neuwald and Lansman 1997). In our study, the GGI mu-
tant was not affected in vivo, but, as is true for the
RGYG mutant, the alanine substitutions may not be
profound sequence changes for these particular residues.
Overall, our data support previous speculation that re-
gions II and III are involved in acetyl-CoA binding.
Regions I and IV of the Gcn5 subfamily display no
significant homology to the other members of the super-
family (Neuwald and Lansman 1997). The KQL and PKM
mutations in region I, however, result in similar pheno-
types as the mutations in the putative acetyl-CoA-bind-
ing site in regions II and III. These phenotypes include
loss of histone acetyltransferase activity, and loss of
growth and transcription function in vivo, as well as
strong dominant negativity. One possibility is that re-
gion I is involved in binding of the particular substrates
acetylated by Gcn5. If this is true, then the lack of ho-
mology in region I within the superfamily would imply
that the other family members acetylate substrates
other than nucleosomal histones H3 and H2B. These al-
ternate substrates could include free histones, histones
H4 or H2A, or completely dissimilar acetylation targets.
The analysis of mutants KDY196 and KDY242 sug-
gests an important role for these residues in processes
other than direct nucleosome acetylation. Both mutants
had extremely debilitated function in vivo, but retained
low acetylation function (see Fig. 7 for KDY196; data not
shown for KDY242). These phenotypes suggest a pos-
sible regulatory function for these residues or interaction
with other proteins that are not involved in acetylation.
In particular, because region IV is not well conserved
with other acetyltransferases (Neuwald and Lansman
1997), and KDY242 was the only profoundly defective
mutant in this region, this is consistent with a function
of region IV specific to regulation of Gcn5.
Overall, the data demonstrate that all four highly con-
served subdomains contain important elements for func-
tion of Gcn5. Regions I, II, and III are likely to be in-
volved in enzymatic activity and substrate specificity,
whereas region IV may have a regulatory role.
Are histones bona fide physiological
substrates of Gcn5?
As noted above, the HAT domains of Gcn5 subfamily
members, although remarkably conserved with respect
to each other, display no obvious sequence conservation
with the putative HAT domains of either CBP/p300 or
TAF
II
250 (nor are the latter two HAT domains mutually
conserved). These observations raise the possibility that
the known HATs share yet unidentified structural mo-
tifs. A second, and not exclusive explanation is that sig-
nificant differences exist in the substrates of acetylation
by the Gcn5 subfamily compared with the other putative
acetyltransferases. Notably, p300 has been shown re-
cently to acetylate certain lysines in the carboxyl termi-
nus of the tumor suppressor and activator p53 (Gu and
Roeder 1997).
Our data are consistent with the hypothesis that
nucleosomal histones are critical physiological sub-
strates of the Gcn5 family. We scanned the conserved
residues of the entire HAT domain, and observed a
strong consistency between the effect of the most del-
eterious Gcn5 substitution mutants on function in vivo
and their effect on histone acetylation in vitro. More-
over, recent data indicate that normal remodeling of
nucleosomes at the PHO5 promoter during transcrip-
tional activation requires Gcn5, and in particular, the
remodeling mirrors the pattern of dependency on the
HAT domain mutations depicted here (P. Gregory, A.
Schmid, M. Zavari, L. Liu, S.L. Berger, and W. Ho¨rz, un-
publ.). Taken together the data strongly suggest that his-
tones are among the key substrates of Gcn5 in vivo.
Materials and methods
Yeast strains
The yeast strain PSY316 (MATa ade-101
D
his3-200 leu2-3,112
lys2 ura3-52) and its derivatives PSY316
D
gcn5 (Candau and
Berger 1996) were used for transformation of HAT domain sub-
stitution mutations in conserved regions II, III, and IV. The
TRP1 disruption of IPY3
D
gcn5 (Roberts and Winston 1997) was
constructed using the hisgURA3 cassette as described (Alani et
al. 1987), and was used for integration of the HAT domain sub-
stitution mutations in conserved region I.
Wang et al.
650 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
Plasmids
pRS414–GCN5 was made by amplifying GCN5 including its
natural promoter and terminator from genomic DNA using
PCR with primers containing 58-XhoI 38-BglII sites and then
cloning GCN5 into the low-copy yeast expression vector
pRS414 at XhoI and BamHI sites. GCN5 with BglII ends was
ligated into high-copy yeast galactose-inducible expression vec-
tor pRS424(P
Gal
) to produce pRS424(P
Gal
)-GCN5. GCN5 was
mutagenized in pRSET, pRS414 or pRS424(P
Gal
)backbone and
then subcloned into other vectors by exchanging BamHI–EcoRI
(for subregion I) or NcoI (subregion II–IV) fragments. GCN5 sub-
region I mutants were subcloned into pRS306 for integration.
The yeast-inducible expression vector containing GST was
made by inserting GST fragment (amplified by PCR from pGEX
plasmid with BamHI and KpnI sites) into the yeast galactose
inducible expression vector pYES. GST–ADA2 was made by
inserting ADA2 at BglII in frame with GST.
Site-directed mutagenesis
The Chameleon double-stranded site-directed mutagenesis kit
(Stratagene) was used to construct the alanine-scan mutants of
Gcn5 using mutagenic oligonucleotides shown in Table 1.
Briefly, the GCN5 plasmid was heat-denatured and hybridized
with two oligonucleotide primers. The selection primer
changed one nonessential unique restriction site to a new re-
striction site on the plasmid backbone. The second primer en-
coded the specific mutations, in which three or four amino acids
in Gcn5 were changed into three or four alanine residues, and a
new PstI restriction site was created simultaneously. Unmuta-
genized parental plasmids were selected against by digestion
using the mutagenized restriction site in the vector backbone,
and mutagenized progeny were identified using the introduced
PstI site and were confirmed by sequencing.
Growth and dominant-negative assays
Growth phenotypes of Gcn5 substitution mutations were ana-
lyzed after transformation of IPY37
D
gcn5
D
trp1 (conserved re-
gion I mutations) PSY316
D
gcn5
D
trp1 (conserved regions II–IV
mutations). Single colonies were streaked on minimal synthetic
medium, and were incubated at 30°C for 2 days. For the dou-
bling time assay, single colonies were inoculated into liquid
fully supplemented synthetic media and rotated overnight at
30°C. The cultures were then diluted to an A
600
of 0.02, and the
OD was checked every 2 hr up to a total of 28 hr.
Dominant-negative assays were carried out by transforming
high-copy expression plasmids pRS425(P
Gal
) or pRS424(P
Gal
)
containing GCN5 mutants into IPY36 or PSY316 strains bearing
wild-type GCN5. Transformants were plated on SD (2% glu-
cose) plates, incubated at 30°C for 3 days. Single colonies were
restreaked on SD (2% galactose) medium, and incubated at
room temperature for 3 days.
Transcription assays
Transcription assays were performed in IPY37
D
gcn5
D
trp1 (con-
served region I mutations) and PSY316
D
gcn5
D
trp1 (conserved
regions II–IV mutations). Plasmids expressing Gal4–VP16
FA
(Berger et al. 1990) and LexA–Gcn4 (Marcus et al. 1994) and the
corresponding reporters were cotransformed. b-Gal activity was
determined as units per milligram of protein. The LacZ report-
ers used were pLGSD5 (Guarente et al. 1982) for Gal4–VP16
activation, and LexA–8x (eight LexA-binding sites) (Candau et
al. 1996) for LexA–Gcn4 activation.
Yeast protein overexpression and in vivo GST-binding assay
Yeast PSY316
D
gcn5 strain transformed with GST–ADA2 and
pRS424(P
Gal
)–GCN5 or GCN5 substitution mutants were in-
oculated into 5 ml of synthetic glucose media and grown over-
night. The cultures were diluted in 100 ml of synthetic media
containing 2% lactate, and grown for 5–6 hr. Galactose was
added to 2% final concentration to induce protein expression.
Cells were harvested when cultures reached A
600
= 1. Cell pel-
lets were washed and resuspended in 400 µl of storage buffer (20
m
M HEPES at pH 8.0, 5 nM EDTA, 20% glycerol, 1 mM b-mer-
captoethanol, 1 µ
M PMSF). Cells were disrupted by vortexing
with glass beads and supernatants were collected. This extract
(200 µl) was incubated with 100 µl of GST bead slurry (50%) for
1 hr in PBS buffer–0.1% Triton-100, and beads were washed
extensively with PBS buffer. The proteins bound to GST beads
were analyzed by SDS-PAGE and immunoblotted using Gcn5
antiserum (Candau and Berger 1996).
Yeast protein complex preparation
Strains transformed with GCN5 substitution mutants were
grown selectively in 200 ml of synthetic media, and then trans-
ferred to 4 liters of YPD media and grown to A
600
= 2. Cells were
harvested and resuspended in 30 ml of extraction buffer (40 m
M
HEPES at pH 7.3, 350 mM NaCl, 0.1% Tween, 10% glycerol,
protease inhibitors). Cells were lysed by glass beads using a bead
beater, and extracts were clarified by centrifugation for 1 hr at
40,000g. Next, extracts were incubated with 10 ml Ni–NTA–
agarose slurry (50%) at 4°C for 2.5 hr. Resins were poured into
a column and washed subsequently with 10 ml of extraction
buffer, 10 ml of buffer A (20 m
M Imidazole at pH 7.0, 100 mM
NaCl, 0.1% Tween 20, 10% glycerol, protease inhibitors).
Bound proteins were eluted with 10 ml of buffer A with 300 m
M
imidazole. The 300 mM imidazole eluate was directly loaded
onto a Mono Q HR5/5 column equilibrated in buffer B (50 m
M
Tris at pH 8.0, 100 mM NaCl, 0.1% Tween 20, 10% glycerol).
After a 10-ml wash with buffer B, bound proteins were eluted
with a 25-ml linear gradient of 100 m
M to 500 mM NaCl in
buffer B. Fractions were collected and subjected to HAT assay.
Western blot analysis
The substitution mutations in the HAT domain of Gcn5 unpre-
dictably altered the elution position of the SAGA complex from
the Mono Q column, and therefore the peak fraction of SAGA
elution for each mutant was determined by Western blot analy-
sis using both Ada2 and Gcn5 antisera (data not shown). Then
final Western blot analyses were performed with this peak frac-
tion from the MonoQ column (Fig. 6B). For the Ada complex,
peak fractions for protein analysis were fractions 20 or 24; frac-
tion 22 could not be used because of a strong nonspecific cross-
reacting protein (data not shown). Western blot analyses were
performed with Gcn5 and Ada2 antisera (Barlev et al. 1995;
Candau and Berger 1996).
HAT assays for free and nucleosomal histones
Different amount of fractions (2 and 6 µl) were used for the
nucleosomal HAT assay and similar histone acetylation pat-
terns (H3/H2B relative to H4/H2A) were observed (data not
shown), indicating that the amount of extracts used in the HAT
assay was in the linear range of HAT enzymatic activity. For
nucleosomal HAT assays, MonoQ fractions from Ada complex
of each Gcn5 mutant were used. Three microliters of each even
numbered fraction (from 18 to 30) were incubated with 1 µg of
Gcn5’s HAT activity required for function in vivo
GENES & DEVELOPMENT 651
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
oligonucleosome cores and [
3
H]acetyl CoA in HAT buffer
(Grant et al. 1997) at 30°C for 30 min. The reactions were sub-
jected to SDS-PAGE, and gels were then treated with Intensify
Buffer (DuPont NEN) and fluorographed. For free histone HAT
assays, the same fractions of SAGA and Ada complexes used in
Western analysis were incubated with 1 µg core histones and
[
3
H]acetyl CoA in HAT buffer. Reactions and fluorograph were
carried out as indicated above.
Acknowledgments
We are grateful to R. Marmorstein for preparation of Figure 1A.
We thank C.D. Allis and M.H. Kuo for communication of un-
published results. We thank J. Workman for oligonucleosomes
and D. Sterner and F. Winston for yeast strains. We thank P.
Lieberman, G. Moore, F. Rauscher, G. Rovera, and J. Workmen
for support and valuable discussions; N. Barlev and G. Moore for
critical reading of the manuscript; and A. Kulak for expert edi-
torial help. The work was supported by grants from the Na-
tional Science Foundation and The Council for Tobacco Re-
search to S.L.B; S.L.B is the recipient of an American Cancer
Society Junior Faculty Research Award.
The publication costs of this article were defrayed in part by
payment of page charges. This article must therefore be hereby
marked ‘‘advertisement’’ in accordance with 18 USC section
1734 solely to indicate this fact.
References
Alani, E., L. Cao, and N. Kleckner. 1987. A method for gene
disruption that allows repeated use of URA3 selection in the
construction of multiply disrupted yeast strains. Genetics
116: 541–545.
Alland, L., R. Muhle, H. Hou Jr., J. Potes, L. Chin, N. Schreiber-
Agus, and R.A. DePinho. 1997. Role for N-CoR and histone
deacetylase in Sin3-mediated transcriptional repression. Na-
ture 387: 49–55.
Bannister, A. and T. Kouzarides. 1996. The CBP co-activator is
a histone acetyltransferase. Nature 384: 641–643.
Barlev, N., R. Candau, L. Wang, P. Darpino, N. Silverman, and
S. Berger. 1995. Characterization of physical interactions of
the putative transcriptional adaptor, ADA2, with acidic ac-
tivation domains and TATA-binding protein. J. Biol. Chem.
270: 19337–19344.
Berger, S.L., W.D. Cress, A. Cress, S.J. Triezenberg, and L.
Guarente. 1990. Selective inhibition of activated but not
basal transcription by the acidic activation domain of VP16:
Evidence for transcriptional adaptors. Cell 61: 1199–1208.
Berger, S.L., B. Pina, N. Silverman, G.A. Marcus, J. Agapite, J.L.
Regier, S.J. Triezenberg, and L. Guarente. 1992. Genetic iso-
lation of ADA2: A potential transcriptional adaptor required
for function of certain acidic activation domains. Cell
70: 251–265.
Bradbury, E.M. 1992. Reversible histone modifications and the
chromosome cell cycle. BioEssays 14: 9–16.
Brownell, J., J. Zhou, T. Ranalli, R. Kobayashi, D. Edmondson, S.
Roth, and C. D. Allis. 1996. Tetrahymena histone acetyl-
transferase A: A transcriptional co-activator linking gene ex-
pression to histone acetylation. Cell 84: 843–851.
Candau, R. and S.L. Berger. 1996. Structural and functional
analysis of yeast putative adaptors: Evidence for an adaptor
complex in vivo. J. Biol. Chem. 271: 5237–5345.
Candau, R., P. Moore, L. Wang, N. Barlev, C. Ying, C. Rosen,
and S. Berger. 1996. Identification of functionally conserved
human homologues of the yeast adaptors ADA2 and GCN5.
Mol. Cell. Biol. 16: 593–602.
Candau, R., J. Zhou, C.D. Allis, and S.L. Berger. 1997. Histone
acetyltransferase activity and interaction with ADA2 are
critical for GCN5 function in vivo. EMBO J. 16: 555–565.
Chen, H., R. Lin, R. Schiltz, D. Chakravarti, A. Nash, L. Nagy,
M. Privalsky, Y. Nakatani, and R. Evans. 1997. Nuclear re-
ceptor coactivator ACTR is a novel histone acetyltransferase
and forms a multimeric activation complex with PCAF and
CBP/p300. Cell 90: 569–580.
Chiang, Y., P. Komarnitsky, D. Chase, and C. Denis. 1996.
ADR1 activation domains contact the histone acetyltrans-
ferase GCN5 and the core transcriptional factor TFIIB. J.
Biol. Chem. 271: 32359–32365.
Cote, J., J. Quinn, J.L. Workman, and C.L. Peterson. 1994.
Stimulation of GAL4 derivative binding to nucleosomal
DNA by the yeast SWI/SNF complex. Science 265: 53–60.
Eisenmann, D., K. Arndt, S. Ricupero, J. Rooney, and F. Wins-
ton. 1992. SPT3 interacts with TFIID to allow normal tran-
scription in Saccharomyces cerevisiae. Genes & Dev.
6: 1319–1331.
Eisenmann, D., C. Chapon, S. Roberts, C. Dollard, and F. Win-
ston. 1994. The Saccharomyces cerevisiae SPT8 gene en-
codes a very acidic protein that is functionally related to
SPT3 and TATA-binding protein. Genetics 137: 647–657.
Fascher, K., J. Schmitz, and W. Horz. 1993. Structural and func-
tional requirements for the chromatin transition at the
PHO5 promoter in Saccharomyces cerevisiae upon PHO5
activation. J. Mol. Biol. 231: 658–667.
Gansheroff, L., C. Dollard, P. Tan, and F. Winston. 1995. The S.
cerevisiase SPT7 gene encodes a very acidic protein impor-
tant for transcription in vivo. Genetics 139: 523–536.
Georgakopoulos, T. and G. Thireos. 1992. Two distinct yeast
transcriptional activators require the function of the GCN5
protein to promote normal levels of transcription. EMBO J.
11: 4145–4152.
Gill, G. and M. Ptashne. 1988. Negative effect of the transcrip-
tional activator GAL4. Nature 334: 4721–4724.
Grant, P.A., L. Duggan, J. Cote, S.M. Roberts, J. Brownell, R.
Candau, R. Ohba, T. Owen-Hughes, C.D. Allis, F. Winston,
S.L. Berger, and J.L. Workman. 1997. yGCN5 function
within multisubunit ADA and SPT/ADA adaptor com-
plexes to acetylate nucleosomal histones. Genes & Dev.
11: 1640–1650.
Grunstein, M. 1990. Nucleosomes: regulators of transcription.
Trends Genet. 6: 395–400.
Gu, W. and R. Roeder. 1997. Activation of p53 sequence-specific
DNA binding by acetylation of the p53 C-terminal domain.
Cell 90: 595–606.
Guarante, L. 1995. Transcriptional coactivators in yeast and
beyond. Trends Biol. Sci. 20: 517–521.
Guarente, L., R.R. Yocum, and P. Gifford. 1982. A GAL10-
CYC1 hybrid yeast promoter identifies the GAL4 regulatory
region as an upstream site. Proc. Natl. Acad. Sci. 79: 7410–
7414.
Hassig, C.A., T.C. Fleischer, A.N. Billin, S.L. Schreiber, and D.E.
Ayer. 1997. Histone deacetylase activity is required for full
transcriptional repression by mSin3A. Cell 89: 341–347.
Heinzel, T., R.M. Lavinsky, T.M. Mullen, M. Soderstrom, C.D.
Laherty, J. Torchia, W.M. Yang, G. Brard, S.D. Ngo, J.R.
Davie, E. Seto, R.N. Eisenman, D.W. Rose, C.K. Glass, and
M.G. Rosenfeld. 1997. A complex containing N-CoR, mSin3
and histone deacetylase mediates transcriptional repression
[see comments]. Nature 387: 43–48.
Kadosh, D. and K. Struhl. 1997. Repression by Ume6 involves
recruitment of a complex containing Sin3 corepressor and
Rpd3 histone deacetylase to target promoters. Cell 89: 365–
371.
Wang et al.
652 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
Kingston, R.E., C.A. Bunker, and A.N. Imbalzano. 1996. Repres-
sion and activation by multiprotein complexes that alter
chromatin structure. Genes & Dev. 10: 905–920.
Laherty, C.D., W.M. Yang, J.M. Sun, J.R. Davie, E. Seto, and
R.N. Eisenman. 1997. Histone deacetylases associated with
the mSin3 corepressor mediate mad transcriptional repres-
sion. Cell 89: 349–356.
Lee, D., J. Hayes, D. Pruss, and A. Wolffe. 1993. A positive role
for histone acetylation in transcription factor access to
nucleosomal DNA. Cell 72: 73–84.
Loidl, P. 1994. Histone acetylation: Facts and questions. Chro-
mosoma 103: 441–449.
Marcus, G., N. Silverman, S. Berger, J. Horiuchi, and L.
Guarente. 1994. Functional similarity and physical associa-
tion between GCN5 and ADA2 - putative transcriptional
adaptors. EMBO J. 13: 4807–4815.
Marcus, G., J. Horiuchi, N. Silverman, and L. Guarente. 1996.
ADA5/SPT20 links the ADA and SPT genes, which are in-
volved in yeast transcription. Mol. Cell. Biol. 16: 3197–3205.
Mizzen, C., X.-J. Yang, T. Kokubo, J. Brownell, A. Bannister, T.
Owen-Hughes, J. Workman, L. Wang, S.L. Berger, T. Kouzari-
des, Y. Nakatani, and C.D. Allis. 1996. The TAF250 Subunit
of TFIID Has Histone Acetyltransferase Activity. Cell
87: 1261–1270.
Neuwald, A.F. and D. Lansman. 1997. GCN5-related histone
N-acetyltransferases belong to a diverse superfamily that in-
cludes the yeast SPT10 protein. Trends Biochem. Sci.
22: 154–155.
Ogryzko, V., R. Schlitz, V. Russanova, B. Howard, and Y. Na-
katani. 1996. The Transcriptional Coactivators p300 and
CBP are Histone Acetyltransferases. Cell 87: 953–959.
Owen-Hughes, T. and J.L. Workman. 1994. Experimental analy-
sis of chromatin function in transcription control. Crit. Rev.
Eukary. Gene Exp. 4: 403–441.
Paranjape, S.M., R.T. Kamakaka, and J.T. Kadonaga. 1994. Role
of chromatin structure in the regulation of transcription by
RNA polymerase II. Annu. Rev. Biochem. 63: 265–297.
Pazin, M. and J. Kadonaga. 1997. What’s up and down with
histone deacetylation and transcription? Cell 89: 325–328.
Peterson, C.L., A. Dingwall, and M.P. Scott. 1994. Five SWI/
SNF gene products are components of a large multisubunit
complex required for transcriptional enhancement [see com-
ments]. Proc. Natl. Acad. Sci. 91: 2905–2908.
Pin˜a, B., S. Berger, G.A. Marcus, N. Silverman, J. Agapite, and L.
Guarente. 1993. ADA3: A gene, identified by resistance to
GAL4-VP16, with properties similar to and different from
those of ADA2. Mol. Cell. Biol. 13: 5981–5989.
Reifsnyder, C., J. Lowell, A. Clarke, and L. Pillus. 1996. Yeast
SAS silencing genes and human genes associated with AML
and HIV-1 Tat interactions are homologous with acetyl-
transferases. Nature Genet. 14: 44–49.
Roberts, S. and F. Winston. 1996. SPT20/ADA5 encodes a novel
protein functionally related to the TATA-binding protein
and important for transcription in Saccharomyces cerevi-
siae. Mol. Cell. Biol. 16: 3206–3213.
———. 1997. Essential functional interactions of SAGA, a Sac-
caromyces cerevisiae complex of Spt, Ada and GCN5 pro-
teins, with the Snf/Swi and Srt/mediator complexes. Genet-
ics 147: 451–465.
Rundlett, S.E. and M. Grunstein. 1996. HDA1 and RPD3 are
members of distinct yeast histone deacetylase complexes
that regulate silencing and transcription. Proc. Natl. Acad.
Sci. 93: 14503–14508.
Saleh, A., V. Lang, R. Cook, and C. Brand. 1997. Identification of
native complexes containing the yeast coactivator-repressor
proteins NGG1-ADA3 and ADA2. J. Biol. Chem. 272:
5571–5578.
Silverman, N., J. Agapite, and L. Guarente. 1994. Yeast ADA2
protein binds to the VP16 protein activation domain and
activates transcription. Proc. Natl. Acad. Sci. 91:
11665–11668.
Struhl, K. 1995. Yeast transcriptional regulatory mechanisms.
Annu. Rev. Genet. 29: 651–674.
Svaren, J., J. Schmitz, and W. Horz. 1994. The transactivation
domain of Pho4 is required for nucleosome disruption at the
PHO5 promoter. EMBO J. 13: 4856–4862.
Taunton, J., C.A. Hassig, and S.L. Schreiber. 1996. A mamma-
lian histone deacetylase related to the yeast transcription
regulator Rpd3p. Science 272: 408–411.
Triezenberg, S.J. 1995. Structure and function of transcriptional
activation domains. Curr. Opin. Genet. Dev. 5: 190–196.
Turner, B.M. and L.P. O’Neill. 1995. Histone acetylation in
chromatin and chromosomes. Sem. Cell Biol. 6: 229–236.
Vettese-Dadey, M., P.A. Grant, T.R. Hebbes, C. Crane-Robin-
son, C.D. Allis, and J.L. Workman. 1996. Acetylation of his-
tone H4 plays a primary role in enhancing transcription fac-
tor binding to nucleosomal DNA in vitro. EMBO J. 15: 2508–
2518.
Vidal, M. and R. Gaber. 1991. RPD3 encodes a second factor
required to achieve maximum positive and negative tran-
scriptional states in S. cerevisiae. Mol. Cell. Biol. 11: 6317–
6327.
Wade, P. and A. Wolffe. 1997. Histone acetyltransferases in con-
trol. Curr. Biol. 7: 82–84.
Wang, L., C. Mizzen, C. Ying, R. Candau, N. Barlev, J. Brownell,
C.D. Allis, and S. Berger. 1997. Histone Acetyltransferase
activity is conserved between yeast and human GCN5 and
required for complementation of growth and transcriptional
activation. Mol. Cell. Biol. 17: 519–527.
Wolffe, A. 1992. Chromatin: Structure and function. Academic
Press, London, UK.
Yang, X., V. Ogryzko, J. Nishikawa, B. Howard, and Y. Naka-
tani. 1996. A p300/CBP-associated factor that competes
with the adenoviral E1A oncoprotein. Nature 382: 319–324.
Zawel, L. and D. Reinberg. 1995. Common themes in assembly
and function of eukaryotic transcription complexes. Annu.
Rev. Biochem. 64: 533–561.
Zhang, Y., R. Iratni, H. Erdjument-Bromage, P. Tempst, and D.
Reinberg. 1997. Histone deacetylases and SAP18, a novel
polypeptide, are components of a human Sin3 complex. Cell
89: 357–364.
Gcn5’s HAT activity required for function in vivo
GENES & DEVELOPMENT 653
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from
12:1998, Genes Dev.
Lian Wang, Lin Liu and Shelley L. Berger
vivo
and SAGA complexes, are also required for transcriptional function in
Critical residues for histone acetylation by Gcn5, functioning in Ada
References
http://genesdev.cshlp.org/content/12/5/640.full.html#ref-list-1
This article cites 57 articles, 24 of which can be accessed free at:
License
Service
Email Alerting
click here.right corner of the article or
Receive free email alerts when new articles cite this article - sign up in the box at the top
Cold Spring Harbor Laboratory Press
Cold Spring Harbor Laboratory Press on September 2, 2024 - Published by genesdev.cshlp.orgDownloaded from