ArticlePDF Available

Interchangeable RNA polymerase I and II enhancers

Authors:

Abstract and Figures

The RNA polymerase I (pol I) enhancer of Saccharomyces cerevisiae contains at least three elements commonly associated with RNA polymerase II (pol II) enhancers, binding sites for the transcriptional activators general regulatory factor 2 and autonomously replicating sequence-binding factor I, and a thymidine-rich element. When the particular form of the thymidine-rich element found in the pol I enhancer was placed in front of a pol II promoter, transcription was stimulated 43-fold, comparable to the effect of a powerful pol II activator such as Gal4. Conversely, when two copies of a thymidine-rich element from a pol II enhancer were placed upstream of a pol I promoter, transcription was stimulated 38-fold. This functional reciprocity of pol I and II enhancers may reflect similarities in the mechanisms of transcriptional activation. The pol I enhancer also contains an element that appears to be pol I-specific and prevent the activation of pol II.
Content may be subject to copyright.
Proc.
Nati.
Acad.
Sci.
USA
Vol.
87,
pp.
8202-8206,
November
1990
Biochemistry
Interchangeable
RNA
polymerase
I
and
II
enhancers
(Saccharomyces
cerevisiae/transcription/upstream
activation
sequence/thymidine-rich
element)
YAHLI
LORCH*,
NEAL
F.
LUE,
AND
ROGER
D.
KORNBERG
Department
of
Cell
Biology,
Fairchild
Center,
Stanford
University
School
of
Medicine,
Stanford,
CA
94305
Communicated
by
E.
Peter
Geiduschek,
August
6,
1990
(received
for
review
May
18,
1990)
ABSTRACT
The
RNA
polymerase
I
(pol
I)
enhancer
of
Saccharomyces
cerevisiae
contains
at
least
three
elements
com-
monly
associated
with
RNA
polymerase
H
(pol
II)
enhancers,
binding
sites
for
the
transcriptional
activators
general
regula-
tory
factor
2
and
autonomously
replicating
sequence-binding
factor
I,
and
a
thymidine-rich
element.
When
the
particular
form
of
the
thymidine-rich
element
found
in
the
pol
I
enhancer
was
placed
in
front
of
a
pol
U
promoter,
transcription
was
stimulated
43-fold,
comparable
to
the
effect
of
a
powerful
pol
II
activator
such
as
Gal4.
Conversely,
when
two
copies
of
a
thymidine-rich
element
from
a
pol
U
enhancer
were
placed
upstream
of
a
pol
I
promoter,
transcription
was
stimulated
38-fold.
This
functional
reciprocity
of
pol
I
and
U
enhancers
may
reflect
similarities
in
the
mechanisms
of
transcriptional
activation.
The
pol
I
enhancer
also
contains
an
element
that
appears
to
be
pol
I-specific
and
prevent
the
activation
of
pol
H.
Eukaryotic
RNA
polymerases
I
and
II
(pol
I
and
II)
have
much
in
common,
including
a
high
degree
of
sequence
homology
of
the
two
largest
subunits
(1, 2),
three
subunits
identical
in
the
two
enzymes
(3,
4),
and
similarities
in
promoter
structure.
Both
enzymes
are
capable
of
initiation
at
a
minimal
promoter,
consisting
of
sequences
in
the
immediate
vicinity
of
the
transcription
start
site;
in
both
cases
the
frequency
of
initiation
is
greatly
increased
by
the
presence
of
additional
sequence
elements,
termed
enhancers,
which
may
be
located
a
variable
distance
from
the
promoter
and
in
either
orientation
(5-11).
The
question
arises
whether
enhancers
are
functionally
interchangeable
between
pol
I
and
pot
II
promoters.
Others
have
found
(12)
that
a
number
of
pol
II
enhancers
fail
to
stimulate
initiation
by
mammalian
pol
I,
suggesting
that
the
two
enzymes
are
activated
by
different
mechanisms.
We
report
here
on
findings
that
lead
to
the
opposite
conclusion.
RNA
polymerase
I
is
responsible
for
the
synthesis
of
a
large
precursor
of
ribosomal
RNAs,
35S
in
yeast
and
40S
in
higher
organisms.
The
35S
precursor
is
encoded
by
-=120
tandem
repeats
of
a
gene
and
2.5-kilobase
(kb)
spacer.
Two
parts
of
the
spacer
are
involved
in
initiation
of
the
35S
precursor:
""210
base
pairs
(bp)
immediately
upstream
of
the
initiation
site
and
"'190
bp
located
2.2
kb
further
upstream
(10,
11,
13).
The
190-bp
segment
stimulates
initiation
17-fold
and
functions
at
a
variable
distance
from
the
initiation
site
in
either
orientation,
leading
to
its
designation
as
an
enhancer.
Progressive
deletions
of
the
190-bp
segment
gradually
dimin-
ish
enhancer
activity
(14),
suggesting that
the
enhancer
is
made
up
of
multiple
DNA
elements.
Similarities
between
some
of
these
elements
and
previously
identified
pol
II
enhancer
elements
prompted
us
to
investigate
the
question
of
functional
reciprocity.
One
pol
I
enhancer
element
proved
highly
effective
in
the
activation
of
pol
II
transcription,
whereas
other
pol
I
enhancer
elements
may
play
accessory
roles
at
both
pol
I
and
pol
II
promoters.
MATERIALS
AND
METHODS
Plasmid
DNAs.
Plasmids
for
transcriptional
activation
of
pol
II
were
members
of
the
pCZ
family
of
Escherichia
col-Saccharomyces
cerevisiae
shuttle
vectors,
containing
a
polylinker
upstream
of
a
yeast
CYCJ-E.
coli
lacZ
fusion
gene
(15).
The
synthetic
oligonucleotides
listed
in
Table
1
were
ligated
with
the
1-kbp
EcoRI-Cla
I
and
7.8-kbp
Cla
I-BamHI
fragments
of
pCZ.
The
control
with
no
oligonucleotide
(des-
ignated
A)
inserted
in
the
polylinker
was
as
described
(15).
For
construction
of
a
plasmid
with
a
Gal4-binding
site
and
residues
157-180
of
the
pol
I
enhancer
in
pCZ
(plasmid
designated
pCZGALpolI157-180),
the
synthetic
oligonucleo-
tide
had
the
sequence
CGGGTGACAGCCCTCCGAAGGC-
AAAGATGGGTTGAAAGAGAAGG,
with
termini
as
de-
scribed
in
Table
1.
Synthetic
oligonucleotides
containing
the
minimal
en-
hancer
with
sequences
downstream
or
upstream
of
the
thy-
midine-rich
(T-rich)
element
deleted
(Fig.
3
downstream
del.,
upstream
del.)
were
also
ligated
with
the
1-kbp
EcoRI-Cla
I
and
7.8-kbp
Cla
I-BamHI
fragments
of
pCZ.
The
entire
and
minimal
enhancers
[EcoRI-Xba
I
and
EcoRI-HindIII
frag-
ments
of
pSES5,
respectively
(ref.
18;
gift
of
S.
Roeder,
Yale
University)]
depicted
in
Fig.
3
were
ligated
with
the
2.2-kbp
EcoRI-Bgl
II
and
6.6-kbp
BgI
II-Xho
I
fragments
of
pCZ
and
the
following
Xho
I-Xba
I
or
Xho
1-HindIII
adapter
oligo-
nucleotides:
5'-TCGAGGAAGGGGTTCCCTTCCCCAAGGTCA-5'
and
5'-TCGAGGAAGGGGTTCCCTTCCCCAAGTCGA-5'
(Xmn
I
sites
in
the
adapters
are
underlined).
Orientation
of
the
enhancer
fragments
was
reversed
in
constructions
des-
ignated
entire.,,
and
minimalrey
by
ligation
of
the
same
fragments
of
pSES5
with
the
6.6-kbp
EcoRI-Bgl
II
and
2.2-kbp
BgI
II-Xba
I
or
2.2-kbp
BgI
II-HindIII
fragments
of
pCZ.
The
region
adjacent
to
the
minimal
enhancer
(HindIII-
Xba
I
fragment
of
pSES5)
was
ligated
with
the
6.6-kbp
Xba
I-Bgl
II
and
2.2-kbp
Bgl
II-HindIII
fragments
of
pCZ.
Plasmids
for
transcriptional
activation
of
pol
I
were
deriv-
atives
of
pCZ;
the
pol
I
promoter
was
supplied
by
a
640-bp
Xba
I
(end-filled)-BamHI
fragment
of
pSES5,
extending
from
212
bp
upstream
of
the
initiation
site
of
the
35S
rRNA
precursor
into
vector
sequences
428
bp
downstream.
Inser-
tion
of
the
640-bp
pol
I
promoter
fragment
between
the
EcoRI
(end-filled)
and
BamHI
sites
of
pCZ(DED48)2
(17)
gave
pCpolIZA.
Insertion
of
the
pot
I
promoter
fragment
between
the
Xho
I
(end-filled)
and
BamHI
sites
of
pCZ(DED48)2
gave
pCpolIZ-(DED48)2.
For
construction
of
pCpolIZ-ACT,
the
pol
I
promoter
fragment
and
large
EcoRI-BamHI
fragment
of
Abbreviations:
pol
I
and
II,
RNA
polymerase
I
and
II,
respectively;
GRF2,
general
regulatory
factor
2;
ABFI,
autonomously
replicating
sequence-binding
factor
I;
T-rich,
thymidine-rich.
*To
whom
reprint
requests
should
be
addressed.
8202
The
publication
costs
of
this
article
were
defrayed
in
part
by
page
charge
payment.
This
article
must
therefore
be
hereby
marked
"advertisement"
in
accordance
with
18
U.S.C.
§1734
solely
to
indicate
this
fact.
Proc.
Natl.
Acad.
Sci.
USA
87
(1990)
8203
Table
1.
RNA
pol
I
enhancer
elements
activate
RNA
pol
II
transcription
Oligonucleotide
A
GRF2
binding
35SRNAl
35SRNA2
RAP1
X40
ABFI
binding
rABFI
SPT2
pol
I
enhancer
residues
8-29
40-56
Relative
transcription
1.0
3.7
2.8
1.3
2.5
1.8
10.7
T-rich
r(dAdT)
116-156
43
r(dAdT)rev
156-116
3.2
The
synthetic
oligonucleotides
listed
were
inserted
in
the
poly-
linker
of
pCZ.
Sequences
of
the
35SRNA1,
35SRNA2,
RAPI,
and
X40
oligonucleotides
are
given
in
ref.
16
and
that
of
SPT2
in
ref.
17.
The
rABFl
and
T-rich
oligonucleotides
contained
the
residues
listed
5'-GATCC
G-5'
(Fig.
1),
with
before
the
first
residue
and
3'-G
CAATfPG-3'
after
the
last
residue.
The
resulting
plasmids
were
introduced
into
yeast
strain
15C
(a
leu2-31,
12,
ura3-52,
trpi,
his4-580,
pep4-3),
and
,/-galactosidase
activities
in
cell
extracts
(expressed
as
units
per
mg
of
protein)
were
determined.
Relative
transcription
refers
to
units
of
,8-galactosidase
activity
per
mg
of
protein
divided
by
the
result
obtained
for
the
plasmid
with
no
oligonucleotide
(A)
inserted
in
the
polylinker
(0.6
unit
per
mg
of
protein).
Each
value
listed
is
an
average
of
at
least
two
determinations
on
independent
transformants.
pCZ(DED48)2
were
ligated
with
the
following
double-
stranded
oligonucleotide:
AATTCTCTGTCACCCGGCCTCTATTTTCCATTTTCTTCTTTACCCGCCAC
GAGACAGTGGG
CGGAGATAAAAGGTAAAAGAAGAAAIGGG
GGTG
[general
regulatory
factor
2
(GRF2)-binding
sites
are
under-
lined;
ref.
16].
The
pol
I
promoter
fragment
was
inserted
between
the
Xho
I
(end-filled)
and
BamHI
sites
of
pCZ-3GAL
(19)
to
give
pCpolIZ-3GAL
with
three
Gal4-binding
sites
in
front
of
the
promoter.
Assays
of
Transcriptional
Activation.
Plasmid
DNAs
were
introduced
into
yeast
by
the
lithium
acetate-transformation
procedure
(20).
Transformants
were
picked,
grown,
har-
vested,
and
disrupted,
and
B-galactosidase
assays
were
done
as
described
(16).
Values
of
relative
transcription
or
-fold
activation
represent
3-galactosidase
units
per
mg
of
protein
for
the
fragment
or
oligonucleotide
in
question
divided
by
that
for
no
insert
(A,
0.6
unit
per
mg).
GRF2
RESULTS
The
pol
I
Enhancer
Contains
Three
Sequence
Elements
Commonly
Associated
with
pol
II
Promoters.
Residues
14-23
of
the
pol
I
enhancer
(Fig.
1)
match
the
consensus
sequence
YNNYYACCCG
(Y
=
C
or
T;
N
=
A,
G,
C,
or
T)
for
recognition
by
the
pol
II
activator
protein
GRF2
(16)
and
bind
the
protein
in
vitro
(16).
A
protein
termed
REB1
that
binds
this
region
of
the
pol
I
enhancer
has
been
identified
(21)
and
appears
identical
with
GRF2
(16).
Binding
sites
for
GRF2
occur
upstream
of
many
pol
II
promoters
and
function
synergistically
with
other
upstream
activation
sequence
ele-
ments,
possibly
by
excluding
nucleosomes
from
these
ele-
ments
(16,
22,
23).
A
gene
for
GRF2
has
been
isolated
from
yeast
and
is
essential
for
viability;
the
deduced
amino
acid
sequence
of
the
protein
exhibits
some
homology
to
that
of
human
c-MYB
(39).
Residues
39-58
of
the
pol
I
enhancer
match
the
consensus
sequence
RDCNYNNNNNACGAD
(R
=
A
or
G,
D
=
A,
G,
or
T)
for
recognition
by
the
pol
II
activator
protein
autono-
mously
replicating
sequence-binding
factor
I
(ABFI)
(17),
except
for
the
substitution
of
thymidine
for
cytidine
at
the
third
position.
Substitution
of
adenosine
for
cytidine
at
this
position
was
previously
shown
to
reduce,
but
not
abolish,
ABFI
binding
(17),
and
an
oligonucleotide
with
the
sequence
of
residues
39-58
(Table
1,
rABFI)
competes
with
other
ABFI
sites
for
ABFI-binding
in
vitro
(16).
A
protein
termed
REB2
that
binds
this
region
of
the
pol
I
enhancer
has
been
identified
(21)
and
is
probably
identical
with
ABFI.
Like
GRF2,
ABFI
binds
to
many
pol
II
promoters
and
exerts
a
modest
effect
on
transcription
on
its
own
but
functions
synergistically
with
other
activators
(17).
The
third
region
of
the
poI
I
enhancer
that
resembles
previously
described
pol
II
activation
sequences
is
the
thy-
midine-rich
(T-rich)
stretch
of
residues
116-155.
Although
the
sequence
of
this
region
varies
somewhat
among
the
many
copies
of
the
pol
I
enhancer
in
the
yeast
genome,
the
first
part
(Fig.
1,
residues
116-131),
containing
three
stretches
of
2-5
thymidines,
and
the
last
part
(Fig.
1,
residues
143-153),
containing
nine
contiguous
thymidine
residues,
are
con-
served
among
all
sequences
so
far
determined
(10,
24,
25).
Similar
T-rich
sequences
occur
upstream
of
many
yeast
pol
II
promoters
and
contain
runs
of
3-11
thymidine
residues,
with
as
few
as
15
thymidines
sufficing
for
activation
of
transcription
(26-30).
From
the
previous
deletion
analysis
of
the
pol
I
enhancer
(14),
it
is
apparent
that
all
three
regions,
the
GRF2
and
ABFI
sites
and
the
T-rich
element,
contribute
to
activating
pol
I
transcription.
Deletion
of
residues
7-21,
invading
the
GRF2
site
(Fig.
1),
reduced
transcription
2.7-fold.
Extending
the
deletion
of
residue
32,
removing
the
rest
of
the
GRF2
site,
had
ABFI
GAATTCTATGATCCGGGTAAAAACATGTATTGTATATATCTATTATAATATACGAT
GGAGGATGATAGTGTGTAAGAGTGTACCAT
II I
20
30
40
50
60 70
80
T-rich
TTACTAATGTATGTAAGTTACTATTTACTATTTGGTCTTTTTATTTTTTATTTTTTTTTTTTTTTTCGTTGCAAAGATGGGTTGAAA
I
90
100
110
120
130
140
150
160
170
Hind
avj=
GAGAAGGGCTTTCACAAAGCTTCCCGAGCGTGAAAGGATTTGCCCGGACAGTTTGCTTCATGGAGCAGTTTTTTCCGC
-
7/
A
-
II
1a0
190
200
210
220
230
240
250
336
FIG.1.
Nucleotide
sequence
of
an
RNA
polymerase
I
enhancer
from
S. cerevisiae,
as
determined
by
Stewart
and
Roeder
(18).
GRF2-binding,
ABFI-binding,
and
T-rich
sequences
analyzed
as
synthetic
oligonucleotides
in
the
present
study
are
shaded
and
in
boldface
type.
Biochemistry:
Lorch
et
al.
Proc.
Natl.
Acad.
Sci.
USA
87
(1990)
no
greater
effect,
but
deletion
of
residues
7-46,
encroaching
on
the
ABFI
site,
reduced
transcription
a
further
2.4-fold.
An
even
larger
deletion,
removing
the
entire
T-rich
element,
caused
an
additional
2.3-fold
drop
in
transcription.
The
T-Rich
Region
of
the
pot
I
Enhancer
Activates
pot
II
Transcription.
In
view
of
the
similarity
of
the
GRF2,
ABFI,
and
T-rich
regions
of
the
pot
I
enhancer
to
pol
II
activation
sequences,
we
tested
the
capacity
of
the
regions
to
stimulate
pol
II
transcription
in
vivo.
Synthetic
oligonucleotides
with
sequences
of
all
three
regions
were
inserted
in
a
polylinker
16
bp
upstream
of
the
major
TATA
element
of
the
yeast
CYCI
promoter
fused
to
the
E.
coli
lacZ
gene
in
a
yeast
centromeric
plasmid.
Levels
of
8-galactosidase
activity
were
determined
in
extracts
of
yeast
harboring
the
various
constructions
(Table
1).
The
T-rich
sequence
from
the
pol
I
enhancer
activated
pol
II
transcription
to
a
remarkable
extent
[r(dAdT)
oligonucleotide,
43-fold
stimulation
above
background
with
no
oligonucleotide],
more
than
a
T-rich
sequence
from
a
pol
II
promoter
(the
DED)
T-rich
element,
in
the
form
of
the
DED48
oligonucleotide,
7.4-fold
stimulation;
ref.
15).
Acti-
vation
by
the
pol
I
T-rich
sequence
was
strongly
orientation
dependent
[reduced
from
43-
to
3.2-fold
stimulation
on
re-
versing
orientation
in
the
r(dAdT)re,
oligonucleotide],
as
observed
previously
for
the
DEDI
T-rich
element
(15).
The
effect
of
the
pot
I
T-rich
sequence
was
comparable
to
those
of
the
most
potent
pol
II
activation
sequences
known,
such
as
a
single
Gal4-binding
site
(-100-fold;
ref.
31).
Two
controls
were
done
to
confirm
that
transcription
from
the
CYCI
promoter
constructs
was
from
pot
II
and
not
from
pol
I
recruited
to
the
promoter
by
the
pol
I
T-rich
sequence.
(i)
A
form
of
the
CYCJ
promoter
was
used
from
which
the
major
TATA
element
had
been
deleted
(pCThal,
ref.
32).
This
deletion
abolishes
pol
II
transcription
from
the
promoter
in
vitro
and
in
vivo
(32)
and
reduced
transcription
stimulated
by
the
pol
I
T-rich
element
5.8-fold.
The
low
level
of
transcrip-
tion
remaining
may
be
attributable
to
two
TATA-like
se-
quences
(TATTTT)
within
the
pol
I
T-rich
element.
(it)
The
second
control
was
to
compare
the
transcription
start
sites
obtained
with
a
known
pot
II
activation
sequence
to
those
found
in
the
presence
of
the
pol
I
T-rich
element.
For
this
purpose,
RNA
was
isolated
from
cells
bearing
plasmids
with
either
a
Gal4-binding
site
or
the
pot
I
T-rich
element
in
front
of
the
CYCI
promoter.
The
5'
ends
of
CYCI
transcripts,
revealed
by
RNase
protection
mapping
with
32P-labeled
RNA
probe,
were
the
same
in
the
two
cases
(Fig.
2),
with
a
larger
amount
of
RNA
probe
protected
for
the
Gal4binding
plas-
mid,
congruent
with
the
somewhat
greater
enhancer
activity
of
the
Gal
element
than
the
T-rich
sequence.
The
pattern
of
CYCI
transcription
start
sites
was
the
same
as
reported
(32).
Despite
the
capacity
of
the
pot
I
T-rich
sequence
to
activate
pot
II
transcription,
the
entire
pot
I
enhancer
failed
to
do
so.
The
enhancer
was
initially
tested
in
the
form
of
a
larger
fragment,
including
the
190-bp
minimal
enhancer
region
described
above
and
sequences
immediately
adjacent
to
this
region
because
the
adjacent
sequences
also
contribute
to
enhancer
function
(18).
The
larger
fragment
failed
to
activate
pol
II
transcription
in
either
orientation
(Fig.
3,
entire
and
entirerey),
so
the
minimal
enhancer
region
and
the
adjacent
sequences
were
tested
separately,
but
again
there
was
no
effect
(Fig.
3,
minimal,
minimalre,,
and
adjacent).
Appar-
ently,
sequences
flanking
the
T-rich
element
within
the
minimal
enhancer
not
only
fail
to
contribute
to
activation
of
pol
II
transcription
but
are
actually
inhibitory.
Location
of
the
inhibitory
component
was
determined
by
deleting
the
flanking
sequences.
Removal
of
the
region
down-
stream
from
the
T-rich
element
restored
function
to
nearly
the
level
obtained
with
the
element
alone
(Fig.
3,
downstream
del.).
On
the
other
hand,
addition
of
only
the
downstream
residues
157-180
to
the
T-rich
element
abolished
function
(Fig.
3,
upstream
del.).
We
conclude
that
residues
157-180
1
2
3
603-
310-
281
-
271-
$
234-
sf
194-
FIG.
2.
Patterns
of
transcription
start
sites
with
pol
I
and
pol
II
enhancer
ele-
ments
upstream
of
the
CYCI
promoter.
Synthetic
oligonucleotides
containing
a
Gal4-binding
site
(same
as
G4-1
oligonu-
cleotide
of
ref.
22,
here
designated
GOiP)
(lane
2)
or
the
pol
I
T-rich
element
[same
as
r(dAdT)
oligonucleotide
of
Table
1,
here
designated
HOT1]
(lane
3)
were
inserted
in
the
polylinker
upstream
of
the
CYCI
promoter
in
pCZ.
The
resulting
plasmids
were
introduced
into
yeast
strain
15C
(see
Table
1),
and
RNA
was
isolated
and
subjected
to
RNase
protec-
tion
mapping
with
RNA
probe
from
pSPCTB
as
described
(32).
Markers run
in
lane
1
were
from
an
Hae
III
digest
of
4X174
restriction
fragment
DNA;
sizes
in
nucleotides
are
indicated
at
left.
prevent
activation
of
pot
II
transcription.
This
conclusion
is
supported
by
the
effect
of
inserting
residues
157-180
between
a
Gal4-binding
site
and
the
CYCI
promoter
(in
pCZGAL
polI157-180,
see
Materials
and
Methods).
Activation
by
Gal4
protein
was
reduced
from
>100-
to
only
7-fold.
The
GRF2-
and
ABFI-binding
sequences
associated
with
pot
II
promoters
exert
only
modest
effects
on
transcription
(2-
to
10-fold;
refs.
16
and
17),
so
it
was
not
surprising
to
find
that
the
GRF2-
and
ABFI-binding
sequences
from
the
pol
I
enhancer
barely
stimulated
pot
II
transcription
at
all.
The
GRF2-binding
sequence
(35SRNA2
oligonucleotide)
stimu-
lated
slightly
(2.8-fold),
as
did
a
second
GRF2-binding
se-
quence
from
immediately
upstream
of
the
pot
I
transcription
start
site
(35SRNA1
oligonucleotide,
oppositely
oriented
to
35SRNA2,
3.7-fold
stimulation).
The
ABFI-binding
se-
quence
from
the
pol
I
enhancer
(rABFI
oligonucleotide)
also
activated
pot
II
transcription
slightly
(1.8-fold).
The
DEDI
T-Rich
Element
Activates
pot
I
Transcription.
Identification
of
pot
I
enhancer
elements
that
stimulate
pol
II
transcription
led
us
to
investigate
the
possibility
that
some
pol
II
enhancers
might
activate
pol
I
transcription.
Various
pol
II
enhancers
were
placed
in
front
of
the
35S
rRNA
initiation
region
(212
bp
upstream
of
the
start
site
of
pot
I
transcription),
and
levels
of
transcription
in
yeast
harboring
the
constructions
were
determined
by
RNase
protection
mapping
of
cellular
RNA
(Table
2).
A
pair
of
T-rich
elements
from
upstream
of
the
DED)
gene
(two
copies
of
the
DED48
oligonucleotide)
was
a
potent
activator
of
pot
I
transcription
(38-fold
stimulation),
more
so
than
the
pol
I
enhancer
itself
(17-fold
stimulation).
RNase
protection
mapping
also
showed
that
transcription
stimulated
by
the
DEDI
T-rich
element
was
initiated
at
the
correct
pot
I
start
site
(34),
so
transcription
was
in
all
likelihood
due
to
pol
I.
A
pair
of
GRF2-binding
sites
upstream
of
the
ACT
gene
stimulated
pot
I
transcription
less
(3.8-fold)
when
compared
with
the
effects
of
these
sites
on
pol
II
transcription
(16).
Other
pot
II
enhancers,
such
as
a
Gal4-dependent
activation
sequence,
were
also
less
effec-
tive.
The
capacity
of
both
the
pot
I
and
DED)
T-rich
elements
to
activate
pol
I
and
pol
II
transcription
raised
the
question
whether
the
effects
of
these
elements
are
mediated
by
a
common
protein
factor.
Insertion
of
the
DEDI
T-rich
element
upstream
of
a
pot
II
promoter
activates
transcription
10-
to
8204
Biochemistry:
Lorch
et
al.
Proc.
Natl.
Acad.
Sci.
USA
87
(1990)
8205
fragment
entire
entirerev
minimal
minimalrev
adjacent
downstream
del.
upstream
del.
enhancer
residues
4-336
336-4
4-191
191-4
336-192
4-156
116-180
GRF2
ABFI
T-rich
-
-
-
GRF2
ABFI
T-rich
-
-
-
GRF2
ABFI
T-rich
-AffmW-
-
--ric
T-rich
FIG.
3.
Identification
of
a
region
of
the
pol
I
enhancer
that
prevents
activation
of
pol
1I
transcription.
The
fragments
indicated
(residue
numbers
as
in
Fig.
1)
were
incorporated
in
the
polylinker
of
pCZ, and
their
effects
on
transcription
were
analyzed
as
in
Table
1.
Fragments
are
designated
entire
and
minimal
for
entire
and
minimal
enhancers,
entirer,
and
minimalrv
for
the
same
fragments
in
the
opposite
orientation,
adjacent
for
the
region
immediately
downstream
of
the
minimal
enhancer,
and
downstream
del.
and
upstream
del.
for
synthetic
oligonucleotides
containing
the
minimal
enhancer
with
sequences
downstream
or
upstream
of
the
T-rich
element
deleted.
30-fold
in
vitro,
and
activation
is
inhibited
by
addition
of
free
DEDI
T-rich
element
(DED48
oligonucleotide)
to
the
reac-
tion
(Fig.
4;
ref.
15),
indicating
involvement
of
a
DEDI
T-rich-binding
factor.
By
contrast,
the
pol
I
T-rich
element
caused
no
activation
of
transcription
in
vitro,
and
the
low
level
of
transcription
observed
was
unaffected
by
addition
of
the
free
element
[r(dAdT)
oligonucleotide,
Fig.
4].
Thus,
action
of
the
pol
I
fragment
is
probably
mediated
by
a
distinct
factor,
as
appears
so
for
other
T-rich
elements
that
have
proved
ineffective
in
vitro
(15).
Consistent
with
this
inter-
pretation,
the
pol
I
element
was
a
poor
competitor
of
tran-
scription
from
a
DEDJ
T-rich
template,
and
the
inhibition
that
did
occur
(25%
of
that
with
the
DEDI
T-rich
element;
data
not
shown)
probably
reflects
low
affinity
of
the
DEDI
T-rich-binding
factor
for
the
pol
I
element.
DISCUSSION
Dissection
of
the
pol
I
enhancer
was
essential
for
revealing
the
capacity
of
one
sequence,
the T-rich
element,
to
activate
pol
II
transcription
because
another
sequence,
between
res-
idues
157-180,
was
inhibitory.
These
residues
probably
con-
tain
a
factor-binding
site,
as
deletion
of
residue
166
and
insertion
of
a
linker
greatly
diminishes
activity
of
the
en-
hancer
(18).
It
has
also
been
noted
that
residues
166-173
nearly
match
the
simian
virus
40
enhancer
core
consensus
Table
2.
Effects
of
RNA
pol
II
enhancers
on
RNA
pol
I
transcription
Relative
Sequence
transcription
A
1.0
(DED48)2
38
ACT
4.0
3GAL
3.8
The
sequences
listed
(A,
no
insertion;
DED48,
48-bp
DEDI
T-rich
element;
ACT,
two
GRF2-binding
sites;
3GAL,
three
GALA-binding
sites)
were
placed
in
front
of
a
pol
I
promoter
in
derivatives
of
pCZ
designated
pCpolIZA,
pCpolIZ-(DED48)2,
pCpolIZ-ACT,
and
pCpolIZ-3GAL.
Plasmids
were
introduced
into
yeast
strain
5C
(a
his3A200,
ura3-52);
transformants
were
isolated
and
grown
as
de-
scribed
(15)
to
an
A6w
value
of
0.5
and
RNA
was
isolated
(33);
transcripts
initiated
at
the
35S
precursor
start
site
in
the
plasmids
were
quantitated
by
hybridizing
with
an
RNA
probe,
RNase
diges-
tion,
and
gel
electrophoresis
as
described
(34),
and
then
counted
with
an
AMBIS
(San
Diego)
radioanalytic
imaging
system.
(11).
Factor-binding
between
the
T-rich
element
and
pro-
moter
may
block
activation
of
pol
II,
much
as
was
shown
for
protein-binding
sites
interposed
between
a
Gal4-dependent
activation
sequence
and
promoter
(37).
The
minimal
pol
I
enhancer,
containing
GRF2,
ABFI,
and
T-rich
elements,
but
lacking
residues
157-180,
gave
no
in-
crease
in
activation
of
pol
II
transcription
above
the
level
obtained
with
the
T-rich
element
alone
(Fig.
2).
Both
GRF2
and
ABFI
sites
exert
synergistic
effects
when
located
adja-
cent
to
T-rich
elements
in
other
constructs
(16,
17),
but
in
the
case
of
a
GRF2
site,
the
effect
has
been
shown
to
be
strongly
distance-dependent,
declining
by
85%
when
spaced
from
a
T-rich
element
only
half
that
in
the
pol
I
enhancer.
The
distance-dependence
of
the
ABFI
effect
as
well
as
possible
synergism
between
GRF2
and
ABFI
elements
have
not
been
investigated.
100
n
DED48
c
80-
r(dAdT)
00
fi60-
2
40-
20
.
0
200
400 600
800
oligonucleotide
(ng)
FIG.
4.
Effects
of
T-rich
elements
from
upstream
of
the
DEDI
gene
(DED48
oligonucleotide)
and
from
the
pol
I
enhancer
[r(dAdT)
oligonucleotide,
Fig.
1]
on
pol
II
transcription
in
vitro.
Templates
for
transcription
were
p(DED48)2CG-,
with
two
copies
of
the
DED48
oligonucleotide
(48-bp
DEDI
T-rich
element)
in
front
of
the
yeast
CYCI
promoter
(35),
and
pr(dAdT)CG-,
with
the
r(dAdT)
oligonu-
cleotide.
pr(dAdT)CG-
was
constructed
by
insertion
of
r(dAdT)
between
the
Xho
I
and
HindI.1
sites
of
pGAL4CG-
(35).
Transcrip-
tion
reactions
with
p(DED48)2CG-
(E)
contained
200
ng
of
template
and
the
amounts
of
DED48
oligonucleotide
indicated,
whereas
reactions
with
pr(dAdT)CG-
(*)
contained
380
ng
of
template
and
r(dAdT)
oligonucleotide
[levels
of
transcription
from
pr(dAdT)CG-
were
divided
by
1.8
to
correct
for
the
greater
quantity
of
DNA
used].
Procedures
were
as
described
(36).
relative
transcription
1.7
0.5
0.8
0.7
0.3
38
2.3
Biochemistry:
Lorch
et
al.
Proc.
Natl.
Acad.
Sci.
USA
87
(1990)
The
evidence
presented
here
for
reciprocal
actions
of
pol
I
and
pol
II
enhancers
can
be
interpreted
in
two
ways.
(i)
pol
I
and
pol
II
enhancers
may
function
by
similar
mechanisms.
Stimulatory
proteins
that
bind
to
these
enhancers,
such
as
GRF2,
ABFI,
and
T-rich
binding
factors,
may
interact
with
common
components
of
the
pol
I
and
pol
II
transcription
machineries.
The
selection
of
one
or
the
other
type
of
polymerase
at
a
particular
promoter
would
be
determined
by
additional
sequences.
For
example,
the pol
II-inhibitory
component
of
the
pol
I
enhancer
described
above,
or
the
TATA
element
specific
to
pol
II
promoters,
might
fulfill
this
role.
(ii)
An
alternative
interpretation
would
be
that
different
proteins
bind
to
the
same
enhancer
sequence,
depending
on
whether
it
is
located
in
front
of
a
pol
I
or
pol
II
promoter.
The
detailed
mechanisms
of
transcriptional
activation
might
then
be
different
at
the
two
types
of
promoter.
Although
there
is
precedent
for
multiple
transcription
factors
recognizing
the
same
DNA
sequence
in
higher
cells
(38),
a
single
factor
appears
to
bind
a
particular
sequence
in
most
cases.
The
question
of
whether
one
or
multiple
factors
are
involved
in
the
present
examples
could
be
addressed
with
the
use
of
mutant
binding
sequences,
such
as
those
described
for
ABFI
(17)
and
GRF1
(31).
We
thank
Dr.
S.
Roeder
for
providing
pSES5.
This
research
was
supported
by
National
Institutes
of
Health
Grant
GM36659
to
R.D.K.
1.
Allison,
L.
A.,
Moyie,
M.,
Shales,
M.
&
Ingles,
C.
J.
(1985)
Cell
42,
599-610.
2.
Memet,
S.,
Saurin,
W.
&
Sentenac,
A.
(1988)
J.
Biol.
Chem.
263,
10048-10051.
3.
Sentenac,
A.
(1985)
Crit.
Rev.
Biochem.
18,
31-91.
4.
Woychik,
N.
A.,
Liao,
S.-M.,
Kolodziej,
P.
A.
&
Young,
R.
A.
(1990)
Genes
Dev.
4,
313-323.
5.
Dynan,
W.
S.
&
Tjian,
R.
(1985)
Nature
(London)
316,
774-
778.
6.
Grummt,
1.
(1989)
in
Nucleic
Acids
and
MolecularBiology,
eds.
Eckstein,
F.
&
Lilley,
M.
J.
(Springer
Verlag,
Berlin),
Vol.
3,
pp.
148-163.
7.
Mestel,
R.,
Yip,
M.,
Holland,
J.
P.,
Wang,
E.,
Kang,
J.
&
Holland,
M.
J.
(1989)
Mol.
Cell.
Biol.
9,
1243-1254.
8.
Sollner-Webb,
B.
&
Tower,
J.
(1986)
Annu.
Rev.
Biochem.
55,
801-830.
9.
Jones,
M.
H.,
Learned,
R.
M.
&
Tjian,
R.
(1988)
Proc.
Natl.
Acad.
Sci.
USA
85,
669-673.
10.
Elion,
E.
A.
&
Warner,
J.
R.
(1984)
Cell
39,
663-673.
11.
Elion,
E.
A.
&
Warner,
J.
R.
(1986)
Mol.
Cell.
Biol.
6,
2089-
2097.
12.
Schreck,
R.,
Carey,
M.
F.
&
Grummt,
I.
(1989)
EMBO
J.
8,
3011-3017.
13.
Kempers-Veenstra,
A.
E.,
van
Heerikhuizen,
H.,
Musters,
W.,
Klootwijk,
J.
&
Planta,
R.
J.
(1984)
EMBO
J.
3,
1377-1382.
14.
Mestel,
R.,
Yip,
M.,
Holland,
J.
P.,
Wang,
E.,
Kang,
J.
&
Holland,
M.
J.
(1989)
Mol.
Cell.
Biol.
9,
1243-1254.
15.
Lue,
N.
F.,
Buchman,
A.
R.
&
Kornberg,
R.
D.
(1989)
Proc.
Natl.
Acad.
Sci.
USA
86,
486-490.
16.
Chasman,
D.
I.,
Lue,
N.
F.,
Buchman,
A.
R.,
LaPointe,
J.
W.,
Lorch,
Y.
&
Kornberg,
R.
D.
(1990)
Genes
Dev.
4,
503-514.
17.
Buchman,
A.
R.
&
Kornberg,
R.
D.
(1990)
Mol.
Cell.
Biol.
10,
887-897.
18.
Stewart,
S.
E.
&
Roeder,
G.
S.
(1989)
Mol.
Cell.
Biol.
9,
3464-3472.
19.
Chasman,
D.
I.,
Leatherwood,
J.,
Carey,
M.,
Ptashne,
M.
&
Kornberg,
R.
D.
(1989)
Mol.
Cell.
Biol.
9,
4746-4749.
20.
Ito,
H.,
Kukuda,
Y.,
Murata,
K.
&
Kimura,
A.
(1983)
J.
Bacteriol.
153,
163-168.
21.
Morrow,
B.
E.,
Johnson,
S.
P.
&
Warner,
J.
R.
(1989)
J.
Biol.
Chem.
264,
9061-9068.
22.
Wang,
H.,
Nicholson,
P.
R.
&
Stillman,
D.
J.
(1990)
Mol.
Cell.
Biol.
10,
1743-1753.
23.
Fedor,
M.
J.,
Lue,
N.
F.
&
Kornberg,
R.
D.
(1988)
J.
Mol.
Biol.
204,
109-138.
24.
Skryabin,
K.
G.,
Eldarov,
M.
A.,
Larionov,
V.
L.,
Bayev,
A.
A.,
Klootwijk,
J.,
de
Regt,
C.
H.
F.,
Veldman,
G.
M.,
Planta,
R.
J.,
Georgiev,
0.
1.
&
Hadjiolov,
A.
A.
(1984)
Nu-
cleic
Acids
Res.
12,
2955-2968.
25.
Swanson,
M.
E.,
Yip,
M.
&
Holland,
M.
J.
(1985)
J.
Biol.
Chem.
260,
9905-9915.
26.
Struhl,
K.
(1985)
Nucleic
Acids
Res.
13,
8587-8601.
27.
Struhl,
K.
(1985)
Proc.
Natl.
Acad.
Sci.
USA
82,
8419-8423.
28.
Riedel,
N.,
Wise,
J.
A.,
Swerdlow,
H.,
Mak,
A.
&
Guthrie,
C.
(1986)
Proc.
Natl.
Acad.
Sci.
USA
83,
8097-8101.
29.
Ares,
M.
(1986)
Cell
47,
49-59.
30.
Gallwitz,
D.,
Donath,
C.
&
Sander,
C.
(1983)
Nature
(London)
306,
704-707.
31.
Buchman,
A.
R.,
Kimmerly,
W.
J.,
Rine,
J.
&
Kornberg,
R.
D.
(1988)
Mol.
Cell.
Biol.
8,
210-225.
32.
Lue,
N.
F.
&
Kornberg,
R.
D.
(1987)
Proc.
Natl.
Acad.
Sci.
USA
84,
8839-8843.
33.
Gallwitz,
P.,
Perrin,
F.
&
Seidel,
R.
(1981)
Nucleic
Acids
Res.
9,
6339-6350.
34.
Lue,
N.
F.
&
Kornberg,
R.
D.
(1990)
J.
Biol.
Chem.,
in
press.
35.
Lue,
N.
F.,
Flanagan,
P.,
Sugimoto,
K.
&
Kornberg,
R.
D.
(1989)
Science
246,
661-664.
36.
Lue,
N.
F.,
Flanagan,
P.
M.,
Kelleher,
R.
J.,
III,
Edwards,
A.
M.
&
Kornberg,
R.
D.
(1990)
Methods
Enzymol.,
in
press.
37.
Brent,
R.
&
Ptashne,
M.
(1984)
Nature
(London)
312,
612-615.
38.
Schaffner,
W.
(1989)
Trends
Genet.
5,
37-39.
39.
Ju,
Q.,
Morrow,
B.
E.
&
Warner,
J.
R.
(1990)
Mol.
Cell.
Biol.
10,
in
press.
8206
Biochemistry:
Lorch
et
al.
... As for orientation, some studies have defined the stringent requirement that a DNA sequence can be considered an enhancer only if it can enhance expression no matter it is inverted or not [96][97][98]. However, many examples have shown that inverting an enhancer can in fact substantially alter the expression of the target gene [98][99][100][101][102][103][104]. Nonetheless, this issue seems to be not commonly considered in our surveyed articles, with none of them specifying the orientation explicitly. ...
... The full expression level is usually defined based on a certain reference situation, such as the original genomic context. However, there could be silencers and insulators in the original sequence, the deletion of which can lead to an expression level of the target gene even higher than the natural context, thereby contradicting the definition of 'full' expression [99,142]. Similarly, when multiple enhancers can regulate the same gene, it is unclear whether full expression should be defined based on the enhancers that actively regulate the gene in the natural context, or when all of them are made active artificially. ...
Article
Since the 1st discovery of transcriptional enhancers in 1981, their textbook definition has remained largely unchanged in the past 37 years. With the emergence of high-throughput assays and genome editing, which are switching the paradigm from bottom-up discovery and testing of individual enhancers to top-down profiling of enhancer activities genome-wide, it has become increasingly evidenced that this classical definition has left substantial gray areas in different aspects. Here we survey a representative set of recent research articles and report the definitions of enhancers they have adopted. The results reveal that a wide spectrum of definitions is used usually without the definition stated explicitly, which could lead to difficulties in data interpretation and downstream analyses. Based on these findings, we discuss the practical implications and suggestions for future studies.
... 31 In this respect the yeast rDNA enhancer may be unique among the eukaryotic Pol I stimulating elements. Attempts to define specific sub-elements in this rather long enhancer, using either deletion analysis in minigenes 43,72 or mutational analysis of enhancer sequences in templates transcribed by Pol I in vitro, 35,36 have not resulted in unambiguous conclusions. Only small deletions or mutations at the 3 -end of the enhancer segment appear to cause a drastic reduction in its stimulatory activity. ...
... 31 In this respect the yeast rDNA enhancer may be unique among the eukaryotic Pol I stimulating elements. Attempts to define specific sub-elements in this rather long enhancer, using either deletion analysis in minigenes 43,72 or mutational analysis of enhancer sequences in templates transcribed by Pol I in vitro, 35,36 have not resulted in unambiguous conclusions. Only small deletions or mutations at the 3 -end of the enhancer segment appear to cause a drastic reduction in its stimulatory activity. ...
Article
We have identified a terminator for transcription by RNA polymerase I in the genes coding for rRNA of the yeast Saccharomyces cerevisiae. The terminator is located 108 bp downstream of the 3' end of the mature 25S rRNA and shares several characteristics with previously studied polymerase I terminators in the vertebrates. For example, the yeast terminator is orientation dependent, is inhibited by its own sequence, and forms RNA 3' ends 17 +/- 2 bp upstream of an essential protein binding site. The recognition sequence for binding of the previously cloned REB1 protein (Q. Ju, B. E. Morrow, and J. R. Warner, Mol. Cell. Biol. 10:5226-5234, 1990) is an essential component of the terminator. In addition, the efficiency of termination depends upon sequence context extending at least 12 bp upstream of the REB1 site.
Article
In Saccharomyces cerevisiae, the rRNA genes are organized as a tandem array of head-to-tail repeats. An enhancer of rRNA transcription is present just at the end of each transcription unit, 2 kb away from the next one. This enhancer is unusual for S. cerevisiae in that it acts both upstream and downstream of, and even across, genes. The role of the enhancer in the nutritional regulation of rRNA transcription was studied by introducing a centromere plasmid carrying two rRNA minigenes in tandem, flanking a single enhancer, into cells. Analysis of the transcripts from the two minigenes showed that the enhancer was absolutely required for the stimulation of transcription of rRNA that occurs when cells are shifted from a poor carbon source to a good carbon source. While full enhancer function is provided by a 45-bp region at the 3' end of the 190-bp enhancer, some activity was also conferred by other elements, including both a T-rich stretch and a region containing the binding sites for the proteins Reb1p and Abf1p. We conclude that the enhancer is composed of redundant elements and that it is a major element in the regulation of rRNA transcription.
Chapter
Regulation of transcription of the rRNA genes may be central to the intricate process of ribosome biosynthesis in response to environmental conditions. In Saccharomyces cerevisiae the rRNA genes are organized in a tandem array of ~150 units on chromosome XII (Petes, 1979; Warner, 1989). The genes encoding 17S, 5.8S and 26S rRNA are arranged in a prerRNA operon, which is transcribed by RNA polymerase I (Pol I) in the nucleolus (cf. Fig 1).
Chapter
In many systems a correlation can be drawn between an increase in the translational capacity of cells and their subsequent embryological development, growth, or response to hormone or other stimulus. In order to distinguish a causal relationship from a coincidental one between the induction of specific protein products and the preceding ribosome accumulation, one would like to be able to separate the events. Such intervention is not feasible in vivo. Therefore, a cell-free system was developed in which isolated intact cockerel liver nuclei were transcriptionally active, in the presence or absence of fractionated or unfractionated active translational machinery isolated from rabbit reticulocytes.
Chapter
For a meeting on Protein Synthesis in Yeast, it seems fitting to focus on the chief agent of protein synthesis, the ribosome, and particularly on the control of the synthesis of this agent. Ultimately, the control of ribosome synthesis is the control of protein synthesis, and the control of protein synthesis is the control of growth. Let us examine what little we know about the control of ribosome synthesis in yeast, and attempt to pose the questions that will occupy us for the duration of this millenium.
Article
The beta A1- and beta A3-crystallins are major polypeptides in the lenses of vertebrates. We present evidence that a single beta A3/A1 gene encodes these two proteins in the chicken. The beta A3/A1 gene has been sequenced and its functional promoter identified in transfection experiments. The chicken beta A3/A1 gene has the same structure as the human orthologue: six exons with standard splice sites and two alternative start codons from which the protein products are apparently translated. Northern analysis revealed an abundant 0.9-kb transcript in the lenses of 1-2-day-old chickens and no detectable transcripts in the rest of the eye, brain, heart, kidney, liver or skeletal muscle. The 5'-flanking sequence of the chicken beta A3/A1 gene is very similar to that of the human and mouse genes, suggesting conservation of important putative regulatory sequences in addition to the TATA box. A thymidine-rich element (bp -218 to -163) and a potential AP-1-binding site (bp -264 to -258) are present within the chicken 5'-flanking region. A DNA fragment from -382 to +22 of the chicken beta A3/A1 gene is sufficient to promote expression of the bacterial cat gene in transfected chicken primary lens epithelial cells, but not in transfected dermal fibroblasts.(ABSTRACT TRUNCATED AT 250 WORDS)
Article
Two yeast/Escherichia coli shuttle vectors have been constructed to analyze promoter structures in Saccharomyces cerevisiae: the multicopy vector, pYLZ-2, and the centromere-based vector, pYLZ-6. Both plasmids contain the coding region of lacZ from E. coli lacking the N-terminal eight amino acids. The truncated reporter gene is preceded by a short polylinker (MCS) suitable for the insertion of promoter fragments. The vectors allow for the study of expression from complete promoters containing UAS and TATA elements, transcriptional start point(s) and the original context of the ATG start codon of a yeast gene. A yeast terminator fragment has been inserted 3' of the lacZ coding region. It contains the transcription termination region of the convergently transcribed yeast genes, GCY1 and PFY1, together with sequences corresponding to the mapped 3'-ends of the respective mRNAs. As an example, reporter activity was measured with promoter fragments from three yeast genes (GCY1, PFY1 and LEO1). The results demonstrate the efficiency of the plasmids for studying constitutive and regulated transcription, both at high and low levels of expression.
Article
Full-text available
GRF2, an abundant yeast protein of Mr approximately 127,000, binds to the GAL upstream activating sequence (UASG) and creates a nucleosome-free region of approximately 230 bp. Purified GRF2 binds to sequences found in many other UASs, in the 35S rRNA enhancer, at centromeres, and at telomeres. Although GRF2 stimulates transcription only slightly on its own, it combines with a neighboring weak activator to give as much as a 170-fold enhancement. This effect of GRF2 is strongly distance-dependent, declining by 85% when 22 bp is interposed between the GRF2 and neighboring activator sites.
Article
Full-text available
Ribosomal gene transcription requires the functional interplay of at least two promoter elements, the upstream control element (UCE) and the start site proximal core, which operate in concert to promote efficient and accurate transcription initiation by RNA polymerase I (pol I). Because this bipartite organization of the rDNA promoter is formally analogous to the organization of a typical pol II promoter, we have examined whether transcriptional activation by upstream activating sequences is brought about by similar molecular mechanisms for both classes of genes. We have replaced the UCE of the mouse rDNA promoter by three different pol II activating sequences (the yeast GAL4 binding sites, the target sequence of the enhancer binding protein E2 from bovine papilloma virus type 1 and the octamer motif), and measured the template activity of these chimeric promoters in the presence of the trans-activating proteins either in a cell free transcription system or in vivo after transfection into mouse cells. In the context of the pol I promoter none of these transcriptional activators enhanced rDNA transcription. The results indicate that activation by UCEs is not interchangeable between genes transcribed by RNA pol I and II, respectively, and suggest that different molecular mechanisms mediate the synergistic action of distant control sequences of different classes of genes.
Article
Full-text available
A thymidine-rich sequence upstream of the DED1 gene of Saccharomyces cerevisiae activated transcription of the CYC1 promoter by RNA polymerase II in vitro. Activation was inhibited by an excess of an oligonucleotide with the same but not a closely related thymidine-rich sequence, pointing to the involvement of a specific thymidine-rich element-binding factor. The extent of activation was as great as 30-fold and showed a similar distance and orientation dependence and a similar effect of deletions in vitro as in vivo.
Article
Full-text available
Sequences within the spacer region of yeast rRNA cistrons stimulate synthesis of the major 35S rRNA precursor in vivo 10- to 30-fold (E. A. Elion and J. R. Warner, Cell 39:663-673, 1984). Spacer sequences that mediate this stimulatory activity are located approximately 2.2 kilobases upstream from sequences that encode the 5' terminus of the 35S rRNA precursor. By utilizing a centromere-containing plasmid carrying a 35S rRNA minigene, a 160-base-pair region of spacer rDNA was identified by deletion mapping that is required for efficient stimulation of 35S rRNA synthesis in vivo. A 22-base-pair sequence, previously shown to support RNA polymerase I-dependent selective initiation of transcription in vitro, was located 15 base pairs upstream from the 3' boundary of the stimulatory region. A 77-base pair region of spacer DNA that mediates transcriptional terminator activity in vivo was identified immediately downstream from the 5' boundary of the stimulatory region. Deletion mutations extending downstream from the 5' boundary of the 160-base-pair stimulatory region simultaneously interfere with terminator activity and stimulation of 35S rRNA synthesis from the minigene. The terminator region supported termination of transcripts initiated by RNA polymerase I in vivo. The organization of sequences that support terminator and promoter activities within the 160-base-pair stimulatory region is similar to the organization of rDNA gene promoters in higher organisms. Possible mechanisms for spacer-sequence-dependent stimulation of yeast 35S rRNA synthesis in vivo are discussed.
Article
A DNA-binding protein has been identified from extracts of the budding yeast Saccharomyces cerevisiae which binds to sites present in the promoter regions of a number of yeast genes transcribed by RNA polymerase II, including SIN3 (also known as SDI1), SWI5, CDC9, and TOP1. This protein also binds to a site present in the enhancer for the 35S rRNA gene, which is transcribed by RNA polymerase I, and appears to be identical to the previously described REB1 protein (B. E. Morrow, S. P. Johnson, and J. R. Warner, J. Biol. Chem. 264:9061-9068, 1989). When oligonucleotides containing a REB1-binding site are placed between the CYC1 upstream activating sequence and TATA box, transcription by RNA polymerase II in vivo is substantially reduced, suggesting that REB1 acts as a repressor of RNA polymerase II transcription. The in vitro levels of the REB1 DNA-binding activity are reduced in extracts prepared from strains bearing a mutation in the SIN3 gene. A greater reduction in REB1 activity is observed if the sin3 mutant strain is grown in media containing galactose as a carbon source.
Article
RNA polymerases I, II, and III share three subunits that are immunologically and biochemically indistinguishable. The Saccharomyces cerevisiae genes that encode these subunits (RPB5, RPB6, and RPB8) were isolated and sequenced, and their transcriptional start sites were deduced. RPB5 encodes a 25-kD protein, RPB6, an 18-kD protein, and RPB8, a 16-kD protein. These genes are single copy, reside on different chromosomes, and are essential for viability. The fact that the genes are single copy, corroborates previous evidence suggesting that each of the common subunits is identical in RNA polymerases I, II, and III. Furthermore, immunoprecipitation of RPB6 coprecipitates proteins whose sizes are consistent with RNA polymerase I, II, and III subunits. Sequence similarity between the yeast RPB5 protein and a previously characterized human RNA polymerase subunit demonstrates that the common subunits of the nuclear RNA polymerases are well conserved among eukaryotes. The presence of these conserved and essential subunits in all three nuclear RNA polymerases and the absence of recognizable sequence motifs for DNA and nucleoside triphosphate-binding indicate that the common subunits do not have a catalytic role but are important for a function shared by the RNA polymerases such as transcriptional efficiency, nuclear localization, enzyme stability, or coordinate regulation of rRNA, mRNA, and tRNA synthesis.
Article
ABFI (ARS-binding protein I) is a yeast protein that binds specific DNA sequences associated with several autonomously replicating sequences (ARSs). ABFI also binds sequences located in promoter regions of some yeast genes, including DED1, an essential gene of unknown function that is transcribed constitutively at a high level. ABFI was purified by specific binding to the DED1 upstream activating sequence (UAS) and was found to recognize related sequences at several other promoters, at an ARS (ARS1), and at a transcriptional silencer (HMR E). All ABFI-binding sites, regardless of origin, provided weak UAS function in vivo when examined in test plasmids. UAS function was abolished by point mutations that reduced ABFI binding in vitro. Analysis of the DED1 promoter showed that two ABFI-binding sites combine synergistically with an adjacent T-rich sequence to form a strong constitutive activator. The DED1 T-rich element acted synergistically with all other ABFI-binding sites and with binding sites for other multifunctional yeast activators. An examination of the properties of sequences surrounding ARS1 left open the possibility that ABFI enhances the initiation of DNA replication at ARS1 by transcriptional activation.
Article
Transcription of the yeast CYC1 promoter fused to a sequence lacking guanosine residues provided a rapid, sensitive assay of initiation by RNA polymerase II in yeast extracts. Initiation was enhanced by yeast and mammalian activator proteins. The adenoviral major late promoter fused to the G-minus sequence was transcribed in yeast extracts with an efficiency comparable to that observed in HeLa extracts, showing that promoters as well as transcription factors are functionally interchangeable across species. Initiation occurred at different sites, approximately 30 and 63 to 69 base pairs downstream of the TATA element of the adenoviral promoter in HeLa and yeast extracts, respectively, distances characteristic of initiation in the two systems in vivo. A component of the transcription system and not the promoter sequence determines the distance to the initiation site.
Article
Fusion proteins known to activate transcription in vivo were tested for the ability to stimulate transcription in vitro in a recently developed Saccharomyces cerevisiae RNA polymerase II transcription system. One fusion protein, whose activation domain was derived from the herpesvirus transcriptional activator VP16, gave more than 100-fold stimulation in the in vitro system. The order of effects of the various proteins was the same for transcription in vitro and in vivo, suggesting that the natural mechanism of activation is preserved in vitro.
Article
Guidelines for submitting commentsPolicy: Comments that contribute to the discussion of the article will be posted within approximately three business days. We do not accept anonymous comments. Please include your email address; the address will not be displayed in the posted comment. Cell Press Editors will screen the comments to ensure that they are relevant and appropriate but comments will not be edited. The ultimate decision on publication of an online comment is at the Editors' discretion. Formatting: Please include a title for the comment and your affiliation. Note that symbols (e.g. Greek letters) may not transmit properly in this form due to potential software compatibility issues. Please spell out the words in place of the symbols (e.g. replace “α” with “alpha”). Comments should be no more than 8,000 characters (including spaces ) in length. References may be included when necessary but should be kept to a minimum. Be careful if copying and pasting from a Word document. Smart quotes can cause problems in the form. If you experience difficulties, please convert to a plain text file and then copy and paste into the form.