ArticlePDF Available

Null mutations in the SNF3gene of Saccharomyces cerevisiaecause a diVerent phenotype than previously isolated missense mutations

Authors:

Abstract and Figures

Missense mutations in the SNF3 gene of Saccharomyces cerevisiae were previously found to cause defects in both glucose repression and derepression of the SUC2 (invertase) gene. In addition, the growth properties of snf3 mutants suggested that they were defective in uptake of glucose and fructose. We have cloned the SNF3 gene by complementation and demonstrated linkage of the cloned DNA to the chromosomal SNF3 locus. The gene encodes a 3-kilobase poly(A)-containing RNA, which was fivefold more abundant in cells deprived of glucose. The SNF3 gene was disrupted at its chromosomal locus by several methods to create null mutations. Disruption resulted in growth phenotypes consistent with a defect in glucose uptake. Surprisingly, gene disruption did not cause aberrant regulation of SUC2 expression. We discuss possible mechanisms by which abnormal SNF3 gene products encoded by missense alleles could perturb regulatory functions.
Content may be subject to copyright.
Vol.
6,
No.
11
MOLECULAR
AND
CELLULAR
BIOLOGY,
Nov.
1986,
p.
3569-3574
0270-7306/86/113569-06$02.00/0
Copyright
C
1986,
American
Society
for
Microbiology
Null
Mutations
in
the
SNF3
Gene
of
Saccharomyces
cerevisiae
Cause
a
Different
Phenotype
than
Do
Previously
Isolated
Missense
Mutations
LENORE
NEIGEBORN,
PAMELA
SCHWARTZBERG,
ROBERT
REID,
AND
MARIAN
CARLSON*
Department
of
Genetics
and
Development
and
Institute
for
Cancer
Research,
Columbia
University
College
of
Physicians
and
Surgeons,
New
York,
New
York
10032
Received
8
April
1986/Accepted
29
July
1986
Missense
mutations
in
the
SNF3
gene
of
Sacharomyces
cerevisiae
were
previously
found
to
cause
defects
in
both
glucose
repression
and
derepression
of
the
SUC2
(invertase)
gene.
In
addition,
the
growth
properties
of
snf3
mutants
suggested
that
they
were
defective
in
uptake
of
glucose
and
fructose.
We
have
cloned
the
SNF3
gene
by
complementation
and
demonstrated
linkage
of
the
cloned
DNA
to
the
chromosomal
SNF3
locus.
The
gene
encodes
a
3-kilobase
poly(A)-containing
RNA,
which
was
fivefold
more
abundant
in
cells
deprived
of
glucose.
The
SNF3
gene
was
disrupted
at
its
chromosomal
locus
by
several
methods
to
create
null
mutations.
Disruption
resulted
in
growth
phenotypes
consistent
with
a
defect
in
glucose
uptake.
Surprisingly,
gene
disruption
did
not
cause
aberrant
regulation
of
SUC2
expression.
We
discuss
possible
mechanisms
by
which
abnormal
SNF3
gene
products
encoded
by
missense
alleles
could
perturb
regulatory
functions.
Glucose
repression,
or
carbon
catabolite
repression,
is
a
global
regulatory
system
affecting
the
expression
of
many
genes.
Our
studies
of
glucose
repression
in
Saccharomyces
cerevisiae
have
focused
on
the
SUC2
gene.
Expression
of
SUC2
is
regulated
only
by
glucose
repression
and
is
modu-
lated
over
a
greater
than
200-fold
range.
The
SUC2
gene
encodes
both
secreted
and
intracellular
forms
of
invertase
via
two
mRNAs
with
different
5'
ends
(4,
7,
18).
Secreted
invertase
is
encoded
by
a
glucose-repressible
1.9-kilobase
(kb)
mRNA.
This
secreted
enzyme
is
responsible
for
the
extracellular
hydrolysis
of
sucrose
and
raffinose.
The
intra-
cellular
invertase
is
encoded
by
a
constitutive
1.8-kb
mRNA
and
has
no
obvious
physiological
function
(22).
We
previously
isolated
mutations
in
six
genes,
SNFI
through
SNF6
(sucrose
nonfermenting),
essential
for
regu-
lated
SUC2
expression
(6,
17).
The
snJf3
mutants
were
unable
to
grow
on
raffinose
and
were
defective
in
growth
on
sucrose,
but
none
showed
pleiotropic
defects
in
utilization
of
galactose
or
nonfermentable
carbon
sources.
Our
11
snf3
mutants
displayed
a
range
of
phenotypes
with
respect
to
regulation
of
SUC2
expression:
derepressed
secreted
invertase
activity
ranged
from
10
to
35%
of
the
wild-type
level.
Moreover,
all
showed
constitutive
(glucose-
insensitive)
synthesis
of
secreted
invertase,
ranging
up
to
20%
of
the
derepressed
wild-type
level
(17;
P.S.,
unpub-
lished
results).
Further
studies
suggested
that
regulation
of
SUC2
transcription
was
aberrant;
some
snJ3
mutations
af-
fected
expression
of
a
gene
fusion
in
which
the
LEU2
promoter
is
controlled
by
the
SUC2
upstream
regulatory
region
(23).
A
puzzling
phenotype
of
the
snJ3
mutants
was
that
all
were
more
defective
in
growth
on
sucrose
and
raffinose
than
would
have
been
predicted
from
their
invertase
activity.
We
suggested
previously
(17)
that
the
snf3
mutants
might
be
defective
in
uptake
of
glucose
and
fructose,
which
are
released
by
the
extracellular
hydrolysis
of
sucrose
and
raffinose.
Although
one
can
imagine
that
such
a
defect
in
glucose
uptake
could
account
for
the
invertase
constitutivity
*
Corresponding
author.
of
snf3
mutants,
it
is
not
obvious
how
such
a
defect
per
se
could
also
impair
derepression
of
secreted
invertase.
We
report
here
the
cloning
of
the
SNF3
gene.
A
3-kb
mRNA
encoded
by
SNF3
was
identified,
and
its
level
was
shown
to
be
regulated
by
glucose
repression.
The
gene
was
disrupted
at
its
chromosomal
locus
by
several
methods.
Disruption
resulted
in
phenotypes
consistent
with
a
defect
in
glucose
uptake,
but
surprisingly
did
not
cause
the
aberrant
regulation
of
invertase
expression
that
was
observed
with
the
missense
mutations.
MATERIALS
AND
METHODS
Strains
and
genetic
methods.
Table
1
lists
the
S.
cerevisiae
strains
used
in
this
study.
Standard
methods
were
used
for
genetic
analysis
(25)
and
transformation
(13).
Scoring
for
carbon
source
utilization
was
done
as
described
previously
(6),
except
that
1
p.g
of
antimycin
A
per
ml
was
added
to
rich
medium
instead
of
incubating
the
plates
anaerobically.
Com-
plementation
of
snJ3
mutations
by
plasmids
was
determined
by
testing
for
anaerobic
growth
on
synthetic
complete
me-
dium
(25)
lacking
uracil
and
containing
the
specified
carbon
source.
Isolation
of
plasmids
carrying
SNF3.
The
library
(gift
of
P.
Novick
and
M.
Rose)
contained
genomic
DNA
from
an
isogenic
derivative
of
strain
S288C
cloned
in
YCp5O.
Plas-
mids
able
to
complement
snf3-72
were
isolated
by
methods
described
previously
(8).
Subclones.
pRRC5
and
pLN19
(see
Fig.
1)
are
subclones
in
vectors
YCp5O
and
pCGS40
(11),
respectively.
pLNC2
was
constructed
by
deleting
the
BamHI
fragment
from
pPSC5.
pLN10
is
a
subclone
of
the
BamHI
fragment
in
pCGS40.
All
other
subclones
are
derivatives
of
YIp5
(3).
pLN204-30
contains
the
3.4-kb
SalI-EcoRI
fragment
cloned
in
YIp5.
pRR1,
pRR2,
and
pRR4
were
derived
from
pLN204-30
by
insertion
of
the
HIS3
BamHI
fragment.
Analysis
of
diploid
transformants
with
SNF3
gene
disrup-
tions.
Genomic
DNAs
from
transformants
were
digested
with
EcoRI
and
analyzed
by
Southern
blot
hybridization
(26)
with
radiolabeled
pLN204-9
DNA.
Strains
carrying
a
disrup-
tion
gave
rise
to
a
4.3-kb
fragment
from
the
wild-type
SNF3
3569
3570
NEIGEBORN
ET
AL.
TABLE
1.
List
of
S.
cerevisiae
strainsa
Strain
Genotype
MCY501
MATa
ade2-101
gal2
SUC2
MCY638
MATa
ura3-52
his4-539
lys2-801
SUC2
MCY657
MATxL
snJ3-72
ura3-52
lys2-801
SUC2
MCY714
MATat
snJ3-217
ura3-52
SUC2
MCY1093
MATa
ura3-52
his4-539
lys2-801
SUC2
MCY1407
MATa
snj3-A4::HIS3
Ahis3
ura3-52
Iys2-801
SUC2
MCY1443
MATa
snJ3-2::HIS3
Ahis3
ura3-52
lys2-801
SUC2
MCY1257
MA
TalMA
Ta
Ahis3lAhis3
ura3-521
+
lys2-8011lys2-801
+
Iade2-101
SUC21SUC2
a
All
strains
are
isogenic
or
congenic
to
S288C.
locus
and
a
larger
fragment
from
the
disrupted
locus
(5
kb
for
the
insertions
and
4.6
kb
for
the
substitution).
Invertase
assays.
Glucose-repressed
cells
were
prepared
by
growing
cells
to
exponential
phase
in
medium
containing
2%
glucose,
and
derepressed
cells
were
obtained
by
shifting
repressed
cells
to
medium
containing
0.05%
glucose,
as
described
previously
(8).
Secreted
invertase
activity
was
assayed
in
whole
cells
(8).
Assays
for
I8-galactosidase
and
maltase.
3-Galactosidase
was
assayed
in
permeabilized
cells
(12).
For
maltase
assays,
cells
were
broken
by
vortexing
with
glass
beads
and
assayed
as
described
previously
(14)
in
the
presence
of
40
,ug
of
phenylmethylsulfonyl
fluoride
per
ml
and
1
mM
EDTA.
Protein
concentrations
were
determined
by
the
Bio-Rad
protein
assay.
RESULTS
Cloning
the
SNF3
gene.
Plasmids
carrying
the
SNF3
gene
were
isolated
from
a
library
by
complementation
of
the
defect
in
raffinose
utilization
caused
by
snf3.
A
library
of
genomic
DNA
cloned
in
the
centromere-containing
vector
YCp5O
was
used
to
transform
strain
MCY657
(snjf3-72
ura3)
to
uracil
independence.
Transformants
were
then
screened
for
anaerobic
growth
on
rich
medium
containing
raffinose.
Seven
different
plasmids
carrying
overlapping
cloned
seg-
ments
were
isolated
from
the
Raf+
transformants
by
passage
through
Escherichia
coli
cells
(8).
Figure
1
shows
the
restric-
tion
maps
of
two
of
these
plasmids,
pPSC4
and
pPSC5.
To
verify
that
the
cloned
DNA
complemented
the
snjf3
muta-
tions,
we
showed
that
transformation
of
strains
MCY657
(snJ3-72)
and
MCY714
(snJ3-217)
with
pPSC4
restored
raf-
finose
utilization
and
proper
invertase
derepression
(Table
2).
The
cloned
DNA
was
shown
to
direct
integration
of
pLN185
(Fig.
1)
to
the
SNF3
chromosomal
locus.
pLN185
was
used
to
transform
MCY638
(SNF3
ura3),
and
a
Ura+
transformant
was
crossed
to
MCY657
(snf3
ura3).
Tetrad
analysis
of
the
resulting
diploid
showed
that
the
Ura+
and
Snf1
phenotypes
cosegregated
2:2
in
19
tetrads,
demonstrat-
ing
that
pLN185
had
integrated
into
the
genome
at
a
site
tightly
linked
to
the
SNF3
locus.
Localization
of
the
SNF3
gene
on
the
cloned
DNA.
Subclones
were
constructed
in
episomal
vectors
and
tested
for
SNF3
function
by
complementation
(Fig.
1).
Each
plas-
mid
was
used
to
transform
MCY657
(snf3-72),
and
four
Ura+
COMPLEME
NTATION
H
R
RRR
S
XB
Bg
R
I.M.
a
I
I
a
I
2.
. '
-
1E
R
B
I
ie.-I
i,*
pRRC5,
pLN19
pLN10,
pLN185
pLN204
-9
pLN204-1
7
pLN204-48
pLN
204-6
pRRI
|HIS3
pRR2
pR
R4
FIG.
1.
Restriction
maps
of
SNF3
clones.
Plasmids
are
described
in
Materials
and
Methods.
Only
the
S.
cerevisiae
DNA
segment
(solid
line)
is
shown,
except
that
the
YEp24
sequence
(shaded
bar)
is
indicated
in
pPSC5
and
pPSC4.
Restriction
sites
within
the
HIS3
fragment
(open
box)
are
not
shown.
The
ability
of
each
episomal
plasmid
to
complement
the
raffinose-nonfermenting
phenotype
of
a
snf3
mutant
is
indicated.
The
wavy
arrow
indicates
the
approximate
position
and
direction
of
transcription
of
the
SNF3
RNA.
Restriction
sites:
B,
BamHI;
Bg,
BglII;
H,
HindIII;
R,
EcoRI;
S,
SaII;
X,
XhoI
PLASMID
pPSC5
pPSC4
pLNC2
+
........
1J.-,
MOL.
CELL.
BIOL.
HIS3
-1111.
SNF3
GENE
OF
S.
CEREVISIAE
3571
transformants
were
tested,
for
raffinose
utilization.
Se-
quences
within
the
4.3-kb
EcoRI
fragment
of
pLN19
were
sufficient
to
complement
the
snf3
defect.
The
inability
of
pLNC2
and
pLN10
to
complement
indicated
that
the
SNF3
gene
spanned
the
BamHI
site.
To
confirm
that
sequences
within
the
4.3-kb
EcoRI
fragment
complemented
the
defect
in
single
copy,
the
centromere-containing
plasmid
pRRC5
was
used
to
transform
strains
carrying
snJ3-72,
snJ3-217,
or
snJ3-39.
pRRC5
complemented
the
defects
in
sucrose
and
raffinose
utilization
and
regulation
of
secreted
invertase
expression
(Table
2).
Structure
and
regulated
expression
of
the
SNF3
RNA.
To
identify
the
RNA
encoded
by
SNF3,
poly(A)-containing
RNAs
from
both
glucose-repressed
and
derepressed
cultures
of
a
wild-type
(SNF3)
strain
were
examined
by
Northern
blot
hybridization
analysis.
A
3-kb
RNA
was
detected
with
three
contiguous
probes,
pLN204-17,
pLN204-48,
and
pLN204-6
(Fig.
2).
Because
the
complementation
analysis
indicated
that
the
SNF3
gene
spanned
the
BamHI
site,
this
3-kb
RNA
must
be
encoded
by
SNF3.
The
approximate
position
of
the
SNF3
RNA
is
indicated
in
Fig.
1;
the
direction
of
transcription
was
inferred
from
studies
of
gene
fusions
expressed
in
bacteria
(J.
Celenza
and
M.
Carlson,
unpublished).
A
second
RNA
was
also
detected:
probes
pLN204-9
and
pLN204-17
hybridized
to
a
4.2-kb
RNA,
which
could
not
be
encoded
in
its
entirety
by
the
SNF3
complementing
region.
Comparison
of
the
amounts
of
SNF3
RNA
in
glucose-
repressed
and
-derepressed
cells
indicated
that
the
level
of
SNF3
RNA
was
regulated
by
glucose
repression
(Fig.
2).
In
three
experiments,
the
SNF3
RNA
was
about
fivefold
more
abundant
in
derepressed
cells
than
in
glucose-repressed
cells.
In
derepressed
cells
the
SNF3
RNA
was
5-
to
10-fold
less
abundant
than
the
URA3
RNA.
Disruption
of
the
chromosomal
SNF3
locus.
The
phenotype
of
a
null
mutation
at
the
SNF3
locus
was
not
apparent
from
the
previous
isolation
of
snJ3
alleles
(17).
Therefore,
the
cloned
SNF3
gene
was
used
to
disrupt
the
chromosomal
SNF3
locus.
Insertion
and
substitution
mutations
were
con-
structed
in
pRR1,
pRR2,
and
pRR4
(Fig.
1)
and
were
then
introduced
into
the
genome
at
the
SNF3
locus
by
the
method
of
Rothstein
(21).
pRR1,
pRR2,
and
pRR4
DNAs
were
cleaved
with
EcoRI
and
SaiI,
and
the
fragments
were
used
to
transform
the
diploid
strain
MCY1257
(Ahis3lAhis3
SNF31SNF3)
to
histidine
prototrophy;
a
diploid
was
used
in
TABLE
2.
Secreted
invertase
activity
in
strains
carrying
SNF3
plasmids
Relevant
Secreted
invertase
activitya
genotype
Plasmid
Repressed
Derepressed
snj3-72
YCp50
1
40
snj3-72
pPSC4
<1
200
snJ3-72
pRRC5
<1
160
snj3-217
YCp50
<1
25
snf3-217
pPSC4
<1
310
snj3-217
pRRC5
<1
220
snj3-39
YCp5O
50
80
snf3-39
pRRC5
6
110
SNF3
pCGS40
<1
220
SNF3
pLN19
<1
190
a
Micromoles
of
glucose
released
per
minute
per
100
mg
(dry
weight)
of
cells.
Values
are
the
average
from
assays
of
two
transformants,
and
standard
errors
were
<15%.
Cultures
were
grown
in
synthetic
complete
medium
(25)
lacking
uracil;
assay
values
are
therefore
not
directly
comparable
to
those
shown
in
Table
3,
for
which
cultures
were
grown
in
rich
medium.
M
R
D
so
-
-
SNF3
_~~-URA3
FIG.
2.
Northern
blot
analysis
of
the
SNF3
RNA.
Poly(A)-
containing
RNAs
were
prepared
(24)
from
cells
of
strain
MCY501
grown
under
repressing
(lane
R)
or
derepressing
(lane
D)
conditions,
separated
by
electrophoresis
in
a
O.9o
agarose
gel
containing
formaldehyde
(15),
and
transferred
to
nitrocellulose
(27).
RNAs
homologous
to
32P-labeled
(20)
pLN204-48
DNA
were
detected
by
hybridization
(5)
and
autoradiography.
The
relative
abundance
of
RNAs
was
estimated
with
a
Joyce-Loebl
Chromoscan
3
densitometer.
Lane
M,
Restriction
fragment
markers.
case
SNF3
proved
to
be
an
essential
gene.
Recombination
events
occurring
on
either
side
of
the
HIS3
gene
would
result
in
the
replacement
of
the
wild-type
sequence
with
the
mutation.
To
verify
the
presence
of
the
appropriate
mutation
at
the
SNF3
locus
of
one
homolog,
genomic
DNA
from
His'
transformants
was
analyzed
by
Southern
blot
hybridization
(see
Materials
and
Methods).
This
analysis
also
confirmed
that
SNF3
is
a
unique
gene.
Diploids
carrying
each
disruption
were
subjected
to
tetrad
analysis.
In
each
case
all
four
spores
of
seven
tetrads
were
viable;
thus,
disruption
of
SNF3
was
not
lethal
in
haploids.
Histidine
prototrophy
cosegregated
2:2
with
inability
to
utilize
raffinose,
indicating
that
disruption
conferred
a
raf-
finose-nonfermenting
phenotype.
The
mutants
also
grew
more
slowly
on
sucrose
than
did
the
wild
type.
These
new
alleles
have
been
designated
as
follows:
the
insertion
derived
from
pRR1,
snJ3-J::HIS3;
the
insertion
from
pRR2,
snf3-
2::HIS3;
and
the
substitution,
snJ3-A4::HIS3.
Each
mutation
was
recessive to
the
wild-type
SNF3
allele
and
failed
to
complement
snj3-72
and
snJ3-217.
pRRC5
complemented
snJ3-A4::HIS3.
These
results
prove
that
we
have
cloned
the
authentic
SNF3
gene
and
disrupted
the
SNF3
locus.
Disruption
of
SNF3
does
not
cause
aberrant
regulation
of
SUC2
expression.
The
three
disruptants
were
assayed
for
secreted
invertase
activity
under
glucose-repressing
and
derepressing
conditions
(Table
3).
Surprisingly,
no
signifi-
cant
defect
in
regulation
of
SUC2
expression
could
be
detected:
secreted
invertase
was
glucose
repressed
and
was
derepressed
to
wild-type
levels
in
response
to
glucose
dep-
rivation.
For
comparison,
Table
3
also
shows
the
aberrant
regulation
of
secreted
invertase
expression
(low-level
constitutivity
and
reduced
derepression)
observed
in
snf3
missense
mutants.
Heteroallelic
diploids
of
genotype
snJ3-A4::HIS31snf3-39
and
snf3-A4:
:HIS31snf3-217
exhibited
defective
invertase
VOL.
6,
1986
3572
NEIGEBORN
ET
AL.
TABLE
3.
Secreted
invertase
activity
in
snj3
mutants
Secreted
invertase
activitya
Relevant
genotype
Repressed
Derepressed
SNF3
<1
230
snJ3-l::HIS3
<1
170
snj3-2::HIS3
<1
220
snJ3-A4::HIS3
<1
210
snf3-39b
40
70
snI3_72b
14
50
snJ3_142b
7
40
snI3-217b
5
25
SNF31SNF3
<1
200
sn]3-A4::HIS31SNF3
<1
210
snJ3-391SNF3
2
160
snJ3-2171SNF3
<1
180
snf3-A4::
HIS31snJ3-39
40
73
snj3-A4::HIS31snJ3-217
5
24
snf3-A4::
HIS31snJ3-A4::
HIS3
3
160
snf3-2171snj3-217
6
29
snf3-391snf3-39
38
68
a
Micromoles
of
glucose
released
per
minute
per
100
mg
(dry
weight)
of
cells.
Values
are
the
average
of
at
least
two
determinations.
Cultures
were
grown
in
rich
medium
(YEP).
b
Values
are
from
Neigeborn
and
Carlson
(17).
regulation
(Table
3)
and
growth
properties
similar
to
those
of
mutants
homozygous
for
snJ3-39
and
snJ3-217,
respectively.
The
finding
that
sn]3-A4::HIS3
is
recessive
to
snf3-39
and
snJ3-217
suggests
that
the
missense
alleles
encode
SNF3
gene
products
with
abnormal
function.
In
addition,
the
phenotypic
resemblance
of
these
diploids
to
the
parent
missense
mutant
argues
that
the
snj3
mutation
is
solely
responsible
for
the
observed
phenotype.
Both
snJ3-39
and
snJ3-217
are
recessive
to
the
wild-type
SNF3
allele
with
respect
to
growth
phenotypes
and
regulation
of
secreted
invertase
synthesis
(Table
3).
snJ3
mutants
are
defective
in
utilization
of
low
concentra-
tions
of
glucose.
The
snf3
null
mutants
were
defective
in
sucrose
and
raffinose
utilization
despite
normal
derepression
2%o
Glucose
a
c
of
secreted
invertase.
These
findings
suggested
that
the
snJ3
mutations
caused
defects
in
the
utilization
of
the
low
con-
centrations
of
glucose
and
fructose
that
result
from
the
extracellular
hydrolysis
of
sucrose
and
raffinose.
We
there-
fore
examined
the
growth
of
single
colonies
of
snJ3
null
mutants
on
rich
medium
containing
2
or
0.1%
glucose.
Colonies
of
snf3-2::HIS3
and
snJ3-A4::HIS3
mutants
grew
more
slowly
than
wild-type
colonies
on
medium
with
0.1%
glucose
(Fig.
3).
No
difference
in
colony
size
was
discernible
on
plates
containing
2%
glucose.
This
reduced
ability
to
grow
on
low
concentrations
of
glucose
would
be
consistent
with
a
defect
in
glucose
uptake.
Effects
of
snf3
mutations
on
expression
of
a
GALIO-lacZ
gene
fusion
and
maltase.
Evaluation
of
the
growth
properties
of
snf3
missense
(17)
and
null
mutants
provided
no
evidence
for
effects
of
snf3
mutations
on
the
expression
of
other
glucose-repressible
genes
besides
SUC2;
these
mutants
were
able
to
grow
on
galactose,
glycerol,
and
ethanol.
For
a
more
sensitive
assay,
we
examined
the
effects
of
snJ3
mutations
on
expression
of
the
GAL1O-lacZ
gene
fusion
on
the
episomal
plasmid
pRY123
(28).
snJ3
mutants
transformed
with
pRY123
were
grown
under
inducing
and
repressing
conditions
and
assayed
for
3-galactosidase
activity
(Table
4).
GAL1O-lacZ
expression
was
inducible
to
within
two-
or
threefold
the
wild-type
level
in
all
snf3
mutants.
Small
defects
in
glucose
repression
were
observed
in
the
snf3-72,
snf3-217,
and
snf3-A4::HIS3
mutants.
The
snJ3-39
mutant,
however,
showed
dramatic
insensitivity
to
glucose,
produc-
ing
nearly
1,000-fold
more
,B-galactosidase
than
the
wild
type
in
the
presence
of
2%
glucose
and
80-fold
more
in
the
presence
of
5%
glucose.
We
also
examined
the
effect
of
snf3-39
on
expression
of
the
glucose-repressible
enzyme
maltase.
Our
strains
are
derived
from
the
maltose-nonfermenting
strain
S288C
and
were
therefore
transformed
with
aplasmid
that
comple-
mented
the
defect,
pM1.2F
(gift
of
M.
Charron
and
C.
Michels).
Assays
of
maltase
activity
showed
that
the
snf3-39
mutant
was
inducible
to
normal
levels,
but
exhibited
16-fold
more
maltase
activity
under
glucose-repressing
conditions
0.13/
Glucose
b
c
FIG.
3.
Growth
of
snJ3
mutants
on
0.1%
glucose.
Strains
were
streaked
for
single
colonies
on
solid
rich
medium
containing
either
2
or
0.1%
glucose
(without
antimycin),
and
the
plates
were
incubated
at
30°C
for
2
days.
(a)
MCY1407
(snJ3-A4::HIS3);
(b)
MCY1443
(snf3-2::HIS3);
(c)
wild-type
strain
MCY1093
(SNF3).
MOL.
CELL.
BIOL.
a
SNF3
GENE
OF
S.
CEREVISIAE
3573
than
did
the
wild
type
(Table
4).
Thus,
snJ3-39,
which
was
the
allele
that
caused
the
most
severe
defect
in
glucose
repression
of
SUC2,
also
caused
defects
in
glucose
repres-
sion
of
two
other
genes.
Effects
of
multiple
copies
of
SNF3
on
invertase
expression.
To
assess
the
effect
of
multiple
copies
of
the
SNF3
gene
on
SUC2
expression,
wild-type
strain
MCY1093
was
trans-
formed
with
the
multicopy
plasmid
pLN19
(Fig.
1).
Trans-
formants
were
grown
under
conditions
of
glucose
repression
and
derepression
with
selection
for
maintenance
of
the
plasmid.
No
significant
alteration
of
SUC2
expression
was
observed
(Table
2).
DISCUSSION
We
have
cloned
the
SNF3
gene
and
identified
the
3-kb
RNA
that
it
encodes.
The
level
of
this
RNA
was
regulated
by
glucose
repression.
The
unexpected
finding
of
this
study
was
that
disruption
of
the
SNF3
gene
resulted
in
a
phenotype
different
from
those
of
previously
isolated
missense
mutants.
The
11
missense
mutants
showed
aberrant
regulation
of
SUC2
expression
in
response
to
glucose
availability,
includ-
ing
defects
in
both
repression
and
derepression.
In
contrast,
disruption
of
the
gene
caused
no
perturbation
in
the
regula-
tion
of
SUC2
by
glucose
repression.
Thus,
regulation
was
normal
in
mutants
carrying
a
disrupted
SNF3
locus,
which
presumably
encoded
no
functional
product,
and
was
aber-
rant
in
missense
mutants.
Regulation
of
two
other
glucose-
repressible
genes,
a
GAL1O-lacZ
fusion
gene
and
the
maltase
gene,
was
strikingly
defective
in
cells
carrying
the
snf3-39
mutation.
To
explain
these
results,
we
suggest
that
these
missense
mutants
each
produce
an
abnormal
SNF3
gene
product
that
is
reasonably
stable
within
the
cell,
and
we
further
propose
that
regulation
is
normal
in
the
absence
of
the
SNF3
gene
product
but
that
abnormal
SNF3
products
can
cause
aber-
rant
regulation.
The
variations
in
phenotypes
among
the
missense
mutants
would
then
result
from
differences
in
the
mutant
gene
products.
The
absence
of
any
detectable
regu-
latory
defect
in
snJ3
null
mutants
suggests
that
the
SNF3
gene
is
not
essential
for
normal
regulation
of
SUC2
expres-
sion.
A
possible
explanation
for
the
regulatory
anomalies
in
the
missense
mutants
is
that
the
SNF3
product
interacts
physically
with
one
or
more
proteins
that
play
regulatory
roles
in
glucose
repression
and
that
mutant
SNF3
proteins
TABLE
4.
Expression
of
a
GAL1O-lacZ
fusion
and
maltase
in
snJ3
mutantsa
1-Galactosidase
activityb
Maltase
activityc
Relevant
Gal-Gly-Glu
Gal-Gly-Glu
Mal-Glu
genotype
(5%),
(2%),
Ga[-Gly,
(5%)'
Mal,
repressed
repressed
repressed
SNF3
0.17
0.37
990
70
8,100
snj3-39
14
350
390
1,100
7,500
snf3-72
0.14
3.9
860
NDd
ND
snjf3-217
0.18
1.2
640
ND
ND
sn13-A4::
HIS3
0.09
2.6
580
ND
ND
aCultures
were
grown
to
mid-log
phase
in
synthetic
complete
medium
lacking
uracil
and
containing
the
indicated
carbon
sources:
Gal,
2%
galactose;
Gly,
3%
glycerol;
Glu,
2
or
5%
glucose,
as
indicated;
Mal,
2%
maltose.
b
Units
of
activity,
normalized
to
the
OD6wo
of
the
culture,
were
calculated
as
described
by
Miller
(16).
Values
are
the
average
of
assays
of
two
transformants,
and
standard
errors
were
<25%.
c
Nanomoles
of
p-nitrophenyl-a-D-glucopyranoside
cleaved
per
minute
per
milligram
of
protein.
Values
are
the
average
of
assays
of
two
transformants,
and
standard
errors
were
<15%.
d
ND,
Not
determined.
perturb
the
function
of
such
neighboring
proteins.
If
that
is
indeed
the
case,
it
may
be
possible
to
obtain
mutations
in
the
genes
encoding
such
proteins
by
pseudoreversion
of
snf3
missense
mutants.
The
snf3
missense
mutations,
which
clearly
cause
a
change
of
function
rather
than
a
simple
loss
of
function,
are
nonetheless
recessive
to
the
wild-type
allele.
This
finding
suggests
that
the
wild-type
SNF3
gene
product
either
func-
tionally
displaces
the
mutant
product
or
perhaps
interacts
with
the
mutant
product
to
restore
normal
function.
There
is
a
simple
explanation
for
the
fact
that
none
of
our
snJ3
missense
mutations
conferred
the
null
phenotype.
In
our
original
mutant
searches,
raffinose-nonfermenting
mu-
tants
were
screened
for
defects
in
derepression
of
secreted
invertase,
and
mutants
showing
normal
derepression
of
invertase
were
not
retained
for
further
study.
The
defect
of
snJ3
null
mutants
in
sucrose
and
raffinose
utilization,
despite
normal
levels
of
secreted
invertase,
and
their
defect
in
utilization
of
0.1%
glucose
suggest
that
SNF3
is
essential
for
glucose
and
fructose
uptake.
Subsequent
analysis
of
the
kinetics
of
glucose
uptake
has
shown
that
snf3
mutants
are
defective
in
the
high-affinity
glucose
uptake
system
(L.
Bisson,
L.
Neigeborn,
M.
Carlson,
and
D.
Fraenkel,
manuscript
in
preparation),
which
is
normally
derepressed
in
response
to
glucose
limitation
(2).
One
pos-
sibility
is
that
SNF3
encodes
a
structural
component
of
the
uptake
system.
We
speculate
that
the
aberrant
glucose
repression
in
snf3
missense
mutants
results
from
a
defect
in
sensing
or
signal-
ling
the
availability
of
glucose
in
the
environment.
A
role
for
SNF3
in
glucose
uptake
is
consistent
with
the
notion
that
abnormal
SNF3
products
could
affect
membrane
or
mem-
brane-associated
proteins
with
sensory
or
signalling
func-
tions.
In
bacteria,
a
protein
of
the
phosphoenolpyruvate-
sugar
phosphotransferase
system,
enzyme
IIIGIc,
is
involved
in
glucose
uptake
and
also
activates
adenylate
cyclase,
and
therefore
plays
a
role
in
the
regulation
of
expression
of
catabolite-repressible
genes
(see
reference
19
for
a
review).
In
S.
cerevisiae
there
is
evidence
for
interaction
between
the
high-affinity
glucose
uptake
system
and
hexokinase
PII(B),
which
is
putatively
a
regulatory
protein
for
glucose
repres-
sion
(9,
10):
the
functioning
of
the
high-affinity
glucose
uptake
system
is
dependent
on
the
presence
of
hexose
kinases
(1).
Previous
genetic
evidence
that
mutations
in
other
trans-acting
genes
required
for
regulated
SUC2
ex-
pression,
SNFI,
SNF2,
SNF4,
and
SSN6,
are
epistatic
to
snJ3
missense
mutations
(17)
also
suggests
that
the
regula-
tory
defect
in
snf3
mutants
occurs
early
in
the
regulatory
circuitry.
ACKNOWLEDGMENTS
We
thank
P.
Novick
and
M.
Rose
for
providing
the
library,
R.
West
for
pRY123,
and
M.
Charron
and
C.
Michels
for
pM1.2F.
This
work
was
supported
by
Public
Health
Service
grant
GM-34095
from
the
National
Institutes
of
Health
and
an
Irma
T.
Hirschl
Research
Career
Award
to
M.C.
LITERATURE
CITED
1.
Bisson,
L.
F.,
and
D.
G.
Fraenkel.
1983.
Involvement
of
kinases
in
glucose
and
fructose
uptake
by
Saccharomyces
cerevisiae.
Proc.
Natl.
Acad.
Sci.
USA
80:1730-1734.
2.
Bisson,
L.
F.,
and
D.
G.
Fraenkel.
1984.
Expression
of
kinase-
dependent
glucose
uptake
in
Saccharomyces
cerevisiae.
J.
Bacteriol.
159:1013-1017.
3.
Botstein,
D.,
S.
C.
Falco,
S.
E.
Stewart,
M.
Brennan,
S.
Scherer,
D.
T.
Stinchcomb,
K.
Struhl,
and
R.
W.
Davis.
1979.
Sterile
host
VOL.
6,
1986
3574
NEIGEBORN
ET
AL.
yeast
(SHY):
a
eukaryotic
system
of
biological
containment
for
recombinant
DNA
experiments.
Gene
8:17-24.
4.
Carlson,
M.,
and
D.
Botstein.
1982.
Two
differentially
regulated
mRNAs
with
different
5'
ends
encode
secreted
and
intracellular
forms
of
yeast
invertase.
Cell
28:145-154.
5.
Carlson,
M.,
J.
L.
Celenza,
and
F.
J.
Eng.
1985.
Evolution
of
the
dispersed
SUC
gene
family
of
Saccharomyces
by
rearrange-
ments
of
chromosome
telomeres.
Mol.
Cell.
Biol.
5:2894-2902.
6.
Carlson,
M.,
B.
C.
Osmond,
and
D.
Botstein.
1981.
Mutants
of
yeast
defective
in
sucrose
utilization.
Genetics
98:25-40.
7.
Carlson,
M.,
R.
Taussig,
S.
Kustu,
and
D.
Botstein.
1983.
The
secreted
form
of
invertase
in
Saccharomyces
cerevisiae
is
synthesized
from
mRNA
encoding
a
signal
sequence.
Mol.
Cell.
Biol.
3:439-447.
8.
Celenza,
J.
L.,
and
M.
Carlson.
1984.
Cloning
and
genetic
mapping
of
SNFI,
a
gene
required
for
expression
of
glucose-
repressible
genes
in
Saccharomyces
cerevisiae.
Mol.
Cell.
Biol.
4:49-53.
9.
Entian,
K.-D.,
and
K.-U.
Frolich.
1984.
Saccharomyces
cerevi-
siae
mutants
provide
evidence
of
hexokinase
Pll
as
a
bifunc-
tional
enzyme
with
catalytic
and
regulatory
domains
for
trigger-
ing
carbon
catabolite
repression.
J.
Bacteriol.
158:29-35.
10.
Entian,
K.-D.,
F.
Hilberg,
H.
Opitz,
and
D.
Mecke.
1985.
Cloning
of
hexokinase
structural
genes
from
Saccharomyces
cerevisiae
mutants
with
regulatory
mutations
responsible
for
glucose
repression.
Mol.
Cell.
Biol.
5:3035-3040.
11.
Goff,
C.
G.,
D.
T.
Moir,
T.
Kohno,
T.
C.
Gravius,
R.
A.
Smith,
E.
Yamasaki,
and
A.
Taunton-Rigby.
1984.
Expression
of
calf
prochymosin
in
S.
cerevisiae.
Gene
27:35-46.
12.
Guarente,
L.
1983.
Yeast
promoters
and
lacZ
fusions
designed
to
study
expression
of
cloned
genes
in
yeast.
Methods
Enzymol.
101C:181-191.
13.
Hinnen,
A.,
J.
B.
Hicks,
and
G.
R.
Fink.
1978.
Transformation
of
yeast.
Proc.
Natl.
Acad.
Sci.
USA
75:1929-1933.
14.
Khan,
N.
A.,
and
N.
R.
Eaton.
1967.
Purification
and
character-
ization
of
maltase
and
a-methylglucosidase
from
yeast.
Biochim.
Biophys.
Acta
146:173-180.
15.
Maniatis,
T.,
E.
F.
Fritsch,
and
J.
Sambrook.
1982.
Molecular
cloning:
a
laboratory
manual.
Cold
Spring
Harbor
Laboratory,
Cold
Spring
Harbor,
N.Y.
16.
Miller,
J.
H.
1972.
Experiments
in
molecular
genetics.
Cold
Spring
Harbor
Laboratory,
Cold
Spring
Harbor,
N.Y.
17.
Neigeborn,
L.,
and
M.
Carlson.
1984.
Genes
affecting
the
regulation
of
SUC2
gene
expression
by
glucose
repression
in
Saccharomyces
cerevisiae.
Genetics
108:845-858.
18.
Perlman,
D.,
H.
0.
Halvorson,
and
L.
E.
Cannon.
1982.
Presecretory
and
cytoplasmic
invertase
polypeptides
encoded
by
distinct
mRNAs
derived
from
the
same
structural
gene
differ
by
a
signal
sequence.
Proc.
Natl.
Acad.
Sci.
USA
79:781-785.
19.
Postma,
P.
W.,
and
J.
W.
Lengeler.
1985.
Phosphoenol-
pyruvate:carbohydrate
phosphotransferase
system
of
bacteria.
Microbiol.
Rev.
49:232-269.
20.
Rigby,
P.
W.
J.,
M.
Dieckmann,
C.
Rhodes,
and
P.
Berg.
1977.
Labeling
deoxyribonucleic
acid
to
high
specific
activity
in
vitro
by
nick
translation
with
DNA
polymerase
I.
J.
Mol.
Biol.
113:237-251.
21.
Rothstein,
R.
J.
1983.
One-step
gene
disruption
in
yeast.
Meth-
ods
Enzymol.
101C:202-210.
22.
Sarokin,
L.,
and
M.
Carlson.
1984.
Upstream
region
required
for
regulated
expression
of
the
glucose-repressible
SUC2
gene
of
Saccharomyces
cerevisiae.
Mol.
Cell.
Biol.
4:2750-2757.
23.
Sarokin,
L.,
and
M.
Carlson.
1985.
Upstream
region
of
the
SUC2
gene
confers
regulated
expression
to
a
heterologous
gene
in
Saccharomyces
cerevisiae.
Mol.
Cell.
Biol.
5:2521-2526.
24.
Sarokin,
L.,
and
M.
Carlson.
1985.
Comparison
of
two
yeast
invertase
genes:
conservation
of
the
upstream
regulatory
re-
gion.
Nucleic
Acids
Res.
13:6089-6103.
25.
Sherman,
F.,
G.
R.
Fink,
and
C.
W.
Lawrence.
1978.
Methods
in
yeast
genetics.
Cold
Spring
Harbor
Laboratory,
Cold
Spring
Harbor,
N.Y.
26.
Southern,
E.
M.
1975.
Detection
of
specific
sequences
among
DNA
fragments
separated
by
gel
electrophoresis.
J.
Mol.
Biol.
98:503-517.
27.
Thomas,
P.
S.
1980.
Hybridization
of
denatured
RNA
and
small
DNA
fragments
transferred
to
nitrocellulose.
Proc.
Natl.
Acad.
Sci.
USA
77:5201-5205.
28.
West,
R.
W.,
Jr.,
R.
R.
Yocum,
and
M.
Ptashne.
1984.
Saccha-
romyces
cerevisiae
GALI-GALIO
divergent
promoter
region:
location
and
function
of
the
upstream
activating
sequence,
UASG.
Mol.
Cell.
Biol.
4:2467-2478.
MOL.
CELL.
BIOL.
... Disruption or deletion of SNF3 of S. cerevisiae leads to defects in glucose and fructose uptake [9] and the inability of growth on raffinose and sucrose [9]. The rgt2 snf3 double mutant of S. cerevisiae is growth-defective on 2% glucose [10,11]. ...
... Disruption or deletion of SNF3 of S. cerevisiae leads to defects in glucose and fructose uptake [9] and the inability of growth on raffinose and sucrose [9]. The rgt2 snf3 double mutant of S. cerevisiae is growth-defective on 2% glucose [10,11]. ...
... These results were unexpected. In S. cerevisiae, the snf3 mutant caused defects in the uptake of glucose and fructose and was unable to grow in raffinose and sucrose [9]. The rgt2 snf3 double mutant of S. cerevisiae was growth defective in 2% glucose [10,11]. ...
Article
Full-text available
Background Glucose repression in yeast leads to the sequential or diauxic utilization of mixed sugars and reduces the co-utilization of glucose and xylose from lignocellulosic biomasses. Study of the glucose sensing pathway helps to construct glucose repression-released yeast strains and enhance the utilization of lignocellulosic biomasses. Results Herein, the glucose sensor/receptor repressor (SRR) pathway of Kluyveromyces marxianus which mainly consisted of KmSnf3, KmGrr1, KmMth1, and KmRgt1 was studied. The disruption of KmSNF3 led to a release of glucose repression, enhanced xylose consumption and did not result in deficient glucose utilization. Over-expression of glucose transporter gene restored the mild decrease of glucose utilization ability of Kmsnf3 strain to a similar level of the wildtype strain but did not restore glucose repression. Therefore, the repression on glucose transporter is parallel to glucose repression to xylose and other alternative carbon utilization. KmGRR1 disruption also released glucose repression and kept glucose utilization ability, although its xylose utilization ability was very weak with xylose as sole carbon source. The stable mutant of KmMth1-ΔT enabled the release of glucose repression irrespective that the genetic background was Kmsnf3, Kmmth1, or wildtype. Disruption of KmSNF1 in the Kmsnf3 strain or KmMTH1-ΔT overexpression in Kmsnf1 strain kept constitutive glucose repression, indicating that KmSNF1 was necessary to release the glucose repression in both SRR and Mig1-Hxk2 pathway. Finally, overexpression of KmMTH1-ΔT released the glucose repression to xylose utilization in S. cerevisiae. Conclusion The glucose repression-released K. marxianus strains constructed via a modified glucose SRR pathway did not lead to a deficiency in the utilization ability of sugar. The obtained thermotolerant, glucose repression-released, and xylose utilization-enhanced strains are good platforms for the construction of efficient lignocellulosic biomass utilization yeast strains.
... Gpr1 transmits its signal through the large G-protein Gpa2 [5][6][7][9][10][11] while Cdc25 and Sdc25 activate the small G-proteins Ras1 and Ras2. The transceptors Snf3 and Rgt2 recruit the protein kinases Yck1 and Yck2 as well as the transcription corepressors Mth1 and Std1 [45][46][47][48][49]. GPCR: G-protein coupled receptor; GAP: GTPase activating protein; GEF: guanine nucleotide exchange factor; AC: adenylyl cyclase; PDE: phosphodiesterase; TF: transcription factor; CKI: casein kinase I; PKA: protein kinase A. ...
... Although they resemble glucose transporters, Snf3 and Rgt2 appear to have lost their transporter function and instead serve exclusively as receptor or "transceptor" proteins. Following glucose addition [45][46][47], Snf3 and Rgt2 recruit the Type I casein kinases Yck1 and Yck2 as well as the transcription corepressors Mth1 and Std1 [48,49]. Subsequent phosphorylation of these factors results in their ubiquitination and degradation [50][51][52]; this derepresses genes encoding hexose transporters and promotes the uptake of the newly available sugars [47,[53][54][55][56][57][58][59][60][61][62]. ...
Article
Full-text available
The yeast Saccharomyces cerevisiae has long been used to produce alcohol from glucose and other sugars. While much is known about glucose metabolism, relatively little is known about the receptors and signaling pathways that indicate glucose availability. Here, we compare the two glucose receptor systems in S. cerevisiae. The first is a heterodimer of transporter-like proteins (transceptors), while the second is a seven-transmembrane receptor coupled to a large G protein (Gpa2) that acts in coordination with two small G proteins (Ras1 and Ras2). Through comprehensive measurements of glucose-dependent transcription and metabolism, we demonstrate that the two receptor systems have distinct roles in glucose signaling: the G-protein-coupled receptor directs carbohydrate and energy metabolism, while the transceptors regulate ancillary processes such as ribosome, amino acids, cofactor and vitamin metabolism. The large G-protein transmits the signal from its cognate receptor, while the small G-protein Ras2 (but not Ras1) integrates responses from both receptor pathways. Collectively, our analysis reveals the molecular basis for glucose detection and the earliest events of glucose-dependent signal transduction in yeast.
... Both genes are expressed at very low levels: about 100-to 300-fold lower than the HXT1 to HXT4 genes (114). Consistent with its proposed role as a high-affinity glucose sensor, SNF3 transcription is repressed at high concentrations of glucose (91,103,109). Rgt2 is proposed to function as a low-affinity glucose sensor, and consistent with this role, its expression is independent of the glucose concentration (114). ...
... However, several subsequent pieces of evidence led to the view that Snf3 has a regulatory rather than a metabolic role in glucose transport. First, SNF3 is expressed at a very low level relative to other glucose transporter genes (about 300-fold less than HXT1) (12,103,114). Second, SNF3 has a negative effect on the growth of an hxt1-hxt4⌬ strain on intermediate levels (0.5%) of glucose, rather than the positive effect that would be expected for a glucose transporter (72). Third, analysis of transport kinetics in a snf3⌬ mutant suggested that the decrease in high-affinity glucose uptake is not due to loss of a single transporter (30). ...
Article
Glucose, the most abundant monosaccharide in nature, is the principal carbon and energy source for nearly all cells. The first, and rate-limiting, step of glucose metabolism is its transport across the plasma membrane. In cells of many organisms glucose ensures its own efficient metabolism by serving as an environmental stimulus that regulates the quantity, types, and activity of glucose transporters, both at the transcriptional and posttranslational levels. This is most apparent in the baker’s yeast Saccharomyces cerevisiae, which has 20 genes encoding known or likely glucose transporters, each of which is known or likely to have a different affinity for glucose. The expression and function of most of these HXT genes is regulated by different levels of glucose. This review focuses on the mechanisms S. cerevisiae and a few other fungal species utilize for sensing the level of glucose and transmitting this information to the nucleus to alter HXT gene expression. One mechanism represses transcription of some HXT genes when glucose levels are high and works through the Mig1 transcriptional repressor, whose function is regulated by the Snf1-Snf4 protein kinase and Reg1-Glc7 protein phosphatase. Another pathway induces HXT expression in response to glucose and employs the Rgt1 transcriptional repressor, a ubiquitin ligase protein complex (SCF Grr1 ) that regulates Rgt1 function, and two glucose sensors in the membrane (Snf3 and Rgt2) that bind glucose and generate the intracellular signal to which Rgt1 responds. These two regulatory pathways collaborate with other, less well-understood, pathways to ensure that yeast cells express the glucose transporters best suited for the amount of glucose available.
... Among the 426 direct targets of SomA were two genes have potential roles in governing glucose uptake and utilization in A. fumigatus, including snf3 (AFUB_030220) and hxk2 (AFUB_089570) (Fig. 3G). The orthologue of snf3 in S. cerevisiae encoding a SomA Regulates Biofilm and Cell Wall ® major facilitator superfamily monosaccharide transporter responsible for glucose uptake (26). The orthologue of hxk2 in S. cerevisiae encoding a putative hexokinase plays a role in glucose phosphorylation (27). ...
Article
Full-text available
The cell wall is essential for fungal viability and is absent from human hosts; thus, drugs disrupting cell wall biosynthesis have gained more attention. Caspofungin is a member of a new class of clinically approved echinocandin drugs to treat invasive aspergillosis by blocking β-1,3-glucan synthase, thus damaging the fungal cell wall. Here, we demonstrate that caspofungin and other cell wall stressors can induce galactosaminogalactan (GAG)-dependent biofilm formation in the human pathogen Aspergillus fumigatus . We further identified SomA as a master transcription factor playing a dual role in both biofilm formation and cell wall homeostasis. SomA plays this dual role by direct binding to a conserved motif upstream of GAG biosynthetic genes and genes involved in cell wall stress sensors, chitin synthases, and β-1,3-glucan synthase. Collectively, these findings reveal a transcriptional control pathway that integrates biofilm formation and cell wall homeostasis and suggest SomA as an attractive target for antifungal drug development.
... Although they are homologous to glucose transporters, Rgt2 and Snf3 appear to have lost their transporter function and instead serve exclusively as receptor or "transceptor" proteins. Following glucose addition (45)(46)(47), Rgt2 and Snf3 recruit the transcription corepressors Mth1 and Std1 (48,49), which are then phosphorylated by Yck1 and Yck2, ubiquitinated, and degraded (50)(51)(52). The destruction of Mth1 and Std1 derepresses genes encoding hexose transporters and promotes the uptake of the newly available sugars (47,(53)(54)(55)(56)(57)(58)(59)(60)(61)(62). ...
Article
Full-text available
The yeast Saccharomyces cerevisiae employs multiple pathways to coordinate sugar availability and metabolism. Glucose and other sugars are detected by a G protein coupled receptor, Gpr1, as well as a pair of transporter-like proteins, Rgt2 and Snf3. When glucose is limiting however, an ATP-driven proton pump (Pma1) is inactivated leading to a marked decrease in cytoplasmic pH. Here we determine the relative contribution of the two sugar sensing pathways to pH regulation. Whereas cytoplasmic pH is strongly dependent on glucose abundance, and is regulated by both glucose-sensing pathways, ATP is largely unaffected and therefore cannot account for the changes in Pma1 activity. These data suggest the pH is a second messenger of the glucose sensing pathways. We show further that different sugars differ in their ability to control cellular acidification, in the manner of inverse agonists. We conclude that the sugar sensing pathways act via Pma1 to invoke coordinated changes in cellular pH and metabolism. More broadly, our findings support the emerging view that cellular systems have evolved the use of pH signals as a means of adapting to environmental stresses such as those caused by hypoxia, ischemia, and diabetes.
Article
Addition of glucose to Saccharomyces cerevisiae cells grown on a nonfermentable carbon source triggers a cyclic AMP (cAMP) signal, which induces a protein phosphorylation cascade. In a yeast strain lacking functional RAS1 and RAS2 genes and containing a bcy mutation to suppress the lethality of RAS deficiency, the cAMP signal was absent. Addition of dinitrophenol, which stimulates in vivo cAMP synthesis by lowering intracellular pH, also did not enhance the cAMP level. A bcy control strain, with functional RAS genes present, showed cAMP responses similar to those of a wild-type strain. In disruption mutants containing either a functional RAS1 gene or a functional RAS2 gene, the cAMP signal was not significantly different from the one in wild-type cells, indicating that RAS function cannot be a limiting factor for cAMP synthesis during induction of the signal. Compared with wild-type cells, the cAMP signal decreased in intensity with increasing temperature in a ras2 disruption mutant. When the mutant RAS2Val-19, which carries the equivalent of the human H-rasVal-12 oncogene, was grown under conditions in which RAS1 expression is repressed, the cAMP signal was absent. The oncogene product is known to be deficient in GTPase activity. However, the amino acid change at position 19 (or 12 in the corresponding human oncogene product) might also have other effects, such as abolishing receptor interaction. Such an additional effect probably provides a better explanation for the lack of signal transmission than the impaired GTPase activity. When the RAS2Val-19 mutant was grown under conditions in which RAS1 is expressed, the cAMP signal was present but significantly delayed compared with the signal in wild-type cells. This indicates that oncogenic RAS proteins inhibit normal functioning of wild-type RAS proteins in vivo and also that in spite of the presence of the RAS2(Val-19) oncogene, adenyl cyclase is not maximally stimulated in vivo. Expression of only the RAS(Val-19) gene product also prevented most of the stimulation of cAMP synthesis by dinitrophenol, indicating that lowered intracellular pH does not act directly on adenyl cyclase but on a step earlier in the activation pathway of the enzyme. The results obtained with the control bcy strain, the RAS2(Val-19) strain under conditions in which RAS1 is expressed, and with dinitrophenol show that the inability of the oncogene product to mediate the cAMP signal is not due to feedback inhibition by the high protein kinase activity in strains containing the RAS2(Val-19) oncogene. Hence, the present results show that the RAS protein in S. cerevisiae are involved in the transmission of the glucose-induced cAMP signal and that the oncogenic RAS protein is unable to act as a signal transducer. The RAS protein in S. cerevisiae apparently act similarly to the Gs proteins of mammalian adenyl cyclase, but instead of being involved in hormone signal transmission, they function in a nutrient-induced signal transmission pathway.
Article
In Saccharomyces cerevisiae, TRK1 and TRK2 are required for high- and low-affinity K+ transport. Among suppressors of the K+ transport defect in trk1 delta trk2 delta cells, we have identified members of the sugar transporter gene superfamily. One suppressor encodes the previously identified glucose transporter HXT1, and another encodes a new member of this family, HXT3. The inferred amino acid sequence of HXT3 is 87% identical to that of HXT1, 64% identical to that of HXT2, and 32% identical to that of SNF3. Like HXT1 and HXT2, overexpression of HXT3 in snf3 delta cells confers growth on low-glucose or raffinose media. The function of another new member of the HXT superfamily, HXT4 (previously identified by its ability to suppress the snf3 delta phenotype; L. Bisson, personal communication), was revealed in experiments that deleted all possible combinations of the five members of the glucose transporter gene family. Neither SNF3, HXT1, HXT2, HXT3, nor HXT4 is essential for viability. snf3 delta hxt1 delta hxt2 delta hxt3 delta hxt4 delta cells are unable to grow on media containing high concentrations of glucose (5%) but can grow on low-glucose (0.5%) media, revealing the presence of a sixth transporter that is itself glucose repressible. This transporter may be negatively regulated by SNF3 since expression of SNF3 abolishes growth of hxt1 delta hxt2 delta hxt3 delta hxt4 delta cells on low-glucose medium. HXT1, HXT2, HXT3, and HXT4 can function independently: expression of any one of these genes is sufficient to confer growth on medium containing at least 1% glucose. A synergistic relationship between SNF3 and each of the HXT genes is suggested by the observation that SNF2 hxt1 delta hxt2 delta hxt3 delta hxt4 delta cells and snf3 delta HXT1 HXT2 HXT3 HXT4 cells are unable to grow on raffinose (low fructose) yet SNF3 in combination with any single HXT gene is sufficient for growth on raffinose. HXT1 and HXT3 are differentially regulated. HXT1::lacZ is maximally expressed during exponential growth whereas HXT3::lacZ is maximally expressed after entry into stationary phase.
Article
The SNF3 gene of Saccharomyces cerevisiae encodes a high-affinity glucose transporter that is homologous to mammalian glucose transporters. Point mutations affecting the function of the transporter were recovered from the genomes of four snf3 mutants and characterized. Two of the mutations introduced a charged amino acid into the first and second predicted membrane-spanning regions, respectively. The analogs of a bifunctional SNF3-lacZ fusion containing these two mutations were constructed, and the mutant fusion proteins were not localized to the plasma membrane, as judged by immunofluorescence microscopy. The third mutation produced a valine-to-isoleucine substitution in hydrophobic region 8, and the corresponding mutant fusion protein was correctly localized. The finding that this conservative change causes a transport defect is consistent with the possibility that this transmembrane region, which could exist as an amphipathic alpha-helix, forms part of the glucose channel through the membrane. The fourth snf3 allele harbored an ochre mutation midway through the coding sequence. We have also constructed mutations in the cloned SNF3 gene. A major difference between the yeast SNF3 protein and mammalian glucose transporters is the presence in the SNF3 protein of an additional 303 amino acids at the C terminus. Analysis of a series of C-terminal deletions and fusions to lacZ showed that this C-terminal region is important, but not essential, for transport function. We also report the genetic mapping of the SNF3 locus on the left arm of chromosome IV.
Article
Two novel genes affecting hexose transport in the yeast Saccharomyces cerevisiae have been identified. The gene HXT1 (hexose transport), isolated from plasmid pSC7, was sequenced and found to encode a hydrophobic protein which is highly homologous to the large family of sugar transporter proteins from eucaryotes and procaryotes. Multicopy expression of the HXT1 gene restored high-affinity glucose transport to the snf3 mutant, which is deficient in a significant proportion of high-affinity glucose transport. HXT1 was unable to complement the snf3 growth defect in low copy number. The HXT1 protein was found to contain 12 putative membrane-spanning domains with a central hydrophilic domain and hydrophilic N- and C-terminal domains. The HXT1 protein is 69% identical to GAL2 and 66% identical to HXT2, and all three proteins were found to have a putative leucine zipper motif at a consensus location in membrane-spanning domain 2. Disruption of the HXT1 gene resulted in loss of a portion of high-affinity glucose and mannose transport, and wild-type levels of transport required both the HXT1 and SNF3 genes. Unexpectedly, expression of beta-galactosidase activity by using a fusion of the lacZ gene to the HXT1 promoter in a multicopy plasmid was maximal during lag and early exponential phases of growth, decreasing approximately 100-fold upon further entry into exponential growth. Deletion analysis of pSC7 revealed the presence of another gene (called ORF2) capable of suppressing the snf3 null mutant phenotype by restoring high-affinity glucose transport and increased low-affinity transport.
Thesis
This thesis describes work on aspects of specific regulated Saccharomyces cerevisiae promoters, namely, the growth phase and medium dependent regulation of the yeast heat-shock promoter element or HSE (Chapter 3); the promoter of the polyubiquitin gene (Chapter 4) and the promoter of the mating factor alpha, MFαl gene (Chapter 5). Chapter 3 describes the use of temperature upshift as a convenient way of inducing heterologous gene expression using a plasmid in which the regulatory elements from a yeast heat-shock promoter were present in a modified CYC1 promoter lacZ fusion in place of the CYC1 promoter regulatory region. Protein induction levels of up to 50 fold were seen on temperature upshift of a logarithmically growing yeast culture from 23°C to 39°C. Heat inducibility of the HSE was maximal at 39°C and lost at stationary phase of growth. The potential problem of protein degradation was tackled by using protease deficient strains which increased β-galactosidase accumulation 2-fold. Studies with the yeast heat shock promoter UBI4 (Chapter 4), chosen because of its potential use as a growth phase dependent promoter, showed that this promoter is primarily controlled (not as previously reported by intracellular cAMP levels) by carbon catabolite repression and this control is exerted probably through the HAP 2/3/4 regulatory system and as such is the first gene for a non-mitochondrial component shown to be controlled by this system. The yeast MFαl promoter was studied (Chapter 5). Results indicate that this promoter is regulated by the growth rate of the host cell, its activity rising as the cellular growth rate falls. Therefore maximum expression levels from this promoter can be achieved by growing cells at a low growth rate under respiratory conditions, a situation that will maximize both cell biomass and protein expression levels.
Article
Full-text available
The regulatory hexokinase PII mutants isolated previously (K.-D. Entian and K.-U. Fröhlich, J. Bacteriol. 158:29-35, 1984) were characterized further. These mutants were defective in glucose repression. The mutation was thought to be in the hexokinase PII structural gene, but it did not affect the catalytic activity of the enzyme. Hence, a regulatory domain for glucose repression was postulated. For further understanding of this regulatory system, the mutationally altered hexokinase PII proteins were isolated from five mutants obtained independently and characterized by their catalytic constants and bisubstrate kinetics. None of these characteristics differed from those of the wild type, so the catalytic center of the mutant enzymes remained unchanged. The only noticeable difference observed was that the in vivo modified form of hexokinase PII, PIIM, which has been described recently (K.-D. Entian and E. Kopetzki, Eur. J. Biochem. 146:657-662, 1985), was absent from one of these mutants. It is possible that the PIIM modification is directly connected with the triggering of glucose repression. To establish with certainty that the mutation is located in the hexokinase PII structural gene, the genes of these mutants were isolated after transforming a hexokinaseless mutant strain and selecting for concomitant complementation of the nuclear function. Unlike hexokinase PII wild-type transformants, glucose repression was not restored in the hexokinase PII mutant transformants. In addition mating experiments with these transformants followed by tetrad analysis of sporulated diploids gave clear evidence of allelism to the hexokinase PII structural gene.
Article
A yeast strain which synthesizes activatable calf prochymosin (also known as prorennin) has been constructed by transformation with a vector carrying the methionyl-prochymosin coding sequence attached to efficient yeast transcriptional promoter and terminator sequences. Cloned preprochymosin cDNA was altered by restriction endonuclease cleavage and addition of a synthetic oligonucleotide to yield a DNA sequence encoding methionyl-prochymosin. This methionyl-prochymosin gene was ligated to a yeast chromosomal fragment containing the GAL1 promoter, and the construction was placed in an Escherichia coli-Saccharomyces cerevisiae shuttle vector with or without a transcriptional terminator DNA fragment from the yeast SUC2 gene. In yeast the two constructions result in equal amounts of prochymosin protein and mRNA. The prochymosin from yeast is activatable to chymosin by incubation at low pH and exhibits milk-clotting activity indistinguishable from calf chymosin.
Article
A system of biological containment for recombinant DNA experiments in Saccharomyces cerevisiae (Brewer's/Baker's yeast) is described. The principle of containment is sterility: the haploid host strains all contain a matingtype-non-specific sterile mutation. The hosts also contain four auxotrophic mutations suitable for selection for the various kinds of vectors used. All vectors are derivatives of pBR322 which can be selected and maintained in both yeast and Escherichia coli. The system has recently been certified at the HV2 level by the National Institutes of Health.
Article
A stable leu2- yeast strain has been transformed to LEU2+ by using a chimeric ColE1 plasmid carrying the yeast leu2 gene. We have used recently developed hybridization and restriction endonuclease mapping techniques to demonstrate directly the presence of the transforming DNA in the yeast genome and also to determine the arrangement of the sequences that were introduced. These studies show that ColE1 DNA together with the yeast sequences can integrate into the yeast chromosomes. This integration may be additive or substitutive. The bacterial plasmid sequences, once integrated, behave as a simple Mendelian element. In addition, we have determined the genetic linkage relationships for each newly introduced LEU2+ allele with the original leu2- allele. These studies show that the transforming squences integrate not only in the leu2 region but also in several other chromosomal locations.
Article
Circular (e.g. simian virus 40) and linear (e.g. λ phage) DNAs have been labeled to high specific radioactivities (>108 cts/min per μg) in vitro using deoxynucleoside [α-32P]triphosphates (100 to 250 Ci/mmol) as substrates and the nick translation activity of Escherichia coli DNA polymerase I. The reaction product yields single-stranded fragments about 400 nucleotides long following denaturation. Because restriction fragments derived from different regions of the nick-translated DNA have nearly the same specific radioactivity (cts/min per 10[su3] bases), we infer that nicks are introduced, and nick translation is initiated, with equal probability within all internal regions of the DNA. Such labeled DNAs (and restriction endonuclease fragments derived from them) are useful probes for detecting rare homologous sequences by in situ hybridization and reassociation kinetic analysis.
Article
This paper describes a method of transferring fragments of DNA from agarose gels to cellulose nitrate filters. The fragments can then be hybridized to radioactive RNA and hybrids detected by radioautography or fluorography. The method is illustrated by analyses of restriction fragments complementary to ribosomal RNAs from Escherichia coli and Xenopus laevis, and from several mammals.