ArticlePDF Available

Modulation of the DNA Scanning Activity of the Micrococcus luteus UV Endonuclease

Authors:

Abstract

Micrococcus luteus UV endonuclease incises DNA at the sites of ultraviolet (UV) light-induced pyrimidine dimers. The mechanism of incision has been previously shown to be a glycosylic bond cleavage at the 5'-pyrimidine of the dimer followed by an apyrimidine endonuclease activity which cleaves the phosphodiester backbone between the pyrimidines. The process by which M. luteus UV endonuclease locates pyrimidine dimers within a population of UV-irradiated plasmids was shown to occur, in vitro, by a processive or "sliding" mechanism on non-target DNA as opposed to a distributive or "random hit" mechanism. Form I plasmid DNA containing 25 dimers per molecule was incubated with M. luteus UV endonuclease in time course reactions. The three topological forms of plasmid DNA generated were analyzed by agarose gel electrophoresis. When the enzyme encounters a pyrimidine dimer, it is significantly more likely to make only the glycosylase cleavage as opposed to making both the glycosylic and phosphodiester bond cleavages. Thus, plasmids are accumulated with many alkaline-labile sites relative to single-stranded breaks. In addition, reactions were performed at both pH 8.0 and pH 6.0, in the absence of NaCl, as well as 25,100, and 250 mM NaCl. The efficiency of the DNA scanning reaction was shown to be dependent on both the ionic strength and pH of the reaction. At low ionic strengths, the reaction was shown to proceed by a processive mechanism and shifted to a distributive mechanism as the ionic strength of the reaction increased. Processivity at pH 8.0 is shown to be more sensitive to increases in ionic strength than reactions performed at pH 6.0.
0
1989 by The American Society
for
Biochemistry and Molecular Biology, Inc
THE
JOURNAL
OF
BIOLOGICAL CHEMISTRY
Val.
264,
No.
29, Issue
of
October
15,
pp.
17422-17427,1989
Printed
in
U.
S.
A.
Modulation
of
the
DNA
Scanning Activity
of
the
Micrococcus
luteus
UV
Endonuclease*
(Received for publication, May 8, 1989)
School of Medicine, Nashuille, Tennessee 37232
Micrococcus luteus
UV endonuclease incises DNA at
the sites of ultraviolet (UV) light-induced pyrimidine
dimers. The mechanism of incision has been previously
shown to be a glycosylic bond cleavage at the 5“pyrim-
idine
of
the dimer followed by an apyrimidinic endo-
nuclease activity which cleaves the phosphodiester
backbone between the pyrimidines. The process by
which
M. luteus
UV endonuclease locates pyrimidine
dimers within a population of UV-irradiated plasmids
was shown to occur,
in vitro,
by a processive or “slid-
ing” mechanism on non-target DNA as opposed to a
distributive or “random hit” mechanism. Form
I
plas-
mid DNA containing 25 dimers per molecule was in-
cubated with
M.
luteus
UV endonuclease in time course
reactions. The three topological forms of plasmid DNA
generated were analyzed by agarose gel electrophore-
sis.
When the enzyme encounters
a
pyrimidine dimer,
it
is
significantly more likely to make only the glyco-
sylase cleavage
as
opposed to making both the glyco-
sylic and phosphodiester bond cleavages. Thus, plas-
mids are accumulated with many alkaline-labile sites
relative to single-stranded breaks. In addition, reac-
tions were performed at both pH
8.0
and pH
6.0,
in the
absence of NaC1, as well as 25,100, and 250 mM NaC1.
The efficiency of the DNA scanning reaction was
shown to be dependent on both the ionic strength and
pH of the reaction. At low ionic strengths, the reaction
was
shown to proceed by
a
processive mechanism and
shifted to
a
distributive mechanism
as
the ionic
strength of the reaction increased. Processivity
at
pH
8.0
is
shown to be more sensitive to increases in ionic
strength than reactions performed at pH
6.0.
The process by which DNA binding proteins locate and
recognize their target sequences within large domains of DNA
has been of considerable interest in recent years. Several
proteins have been shown to locate their target recognition
sites by using a facilitated diffusion mechanism in which the
protein is able to scan
or
slide along DNA in
a
processive
manner (for reviews, see Refs. 1-3). Examples of these types
of proteins are, lac repressor (4-9), T4 endonuclease V (10-
12),
EcoRI
(13-15), BarnHI, BamHI methylase (16), and RNA
*This research was supported in part by United States Public
Health Service Grants ES 04091 and ES 00267. The costs of publi-
cation of this article were defrayed in part by the payment of page
charges. This article must therefore be hereby marked “aduertise-
ment” in accordance with
18
U.S.C. Section 1734 solely to indicate
this fact.
7
Predoctoral fellow supported by the Molecular Biology University
Scholarship and the Harold Sterling Vanderbilt Award.
**
To whom correspondence should be addressed. Tel.: 615-322-
3063.
Richard
W.
HamiltonSgll and
R.
Stephen Lloyd§ll**
From the Departments
of
$Molecular Biolocv and
II
Biochemistry and the §Center in
Molecular
Toxicology, Vanderbilt University
polymerase (17-22). An additional facilitated diffusion mech-
anism which has been proposed is a looping model (for re-
views, see Refs.
1
and 23). This model allows proteins to
transfer between separate segments of DNA. Facilitated dif-
fusion mechanisms allow proteins to locate their recognition
sequences at a rate as much as 1000-fold faster than would be
expected for simple three-dimensional diffusion (3). These
mechanisms involve electrostatic interactions between the
protein and the DNA helix and are sensitive to changes in
ionic strength (2). As ionic strength increases, the dissociation
rate of the protein with nonspecific DNA sequences also
increases. Therefore, low ionic strength conditions increase
the amount of time the protein is bound to nonspecific DNA
sequences, and, as a result, the protein is able to diffuse larger
distances along the DNA.
One of these DNA-reactive enzymes, bacteriophage T4
endonuclease V, is a pyrimidine dimer-specific endonuclease
which has been shown to scan DNA in a processive manner
both
in vitro
and
in vivo
(10-12,24,25). The
in
vitro
scanning
reaction has been shown to be sensitive to increases in ionic
strength with processivity being abolished at salt concentra-
tions higher than 50 mM NaCl at pH 8.0. In addition, site-
directed mutagenesis has been performed on portions of the
denV
gene, which encodes endonuclease V. Analyses of some
of the mutant enzymes has shown that the alteration of
certain positively charged amino acids abolishes the ability of
the enzyme to act processively on non-target DNA
in vitro
(25). Studies of these mutated enzymes
in vivo
have shown
that, unlike wild type endonuclease V, they are unable to
confer enhanced UV resistance to UV-sensitive
Escherichia
coli
hosts. Thus, the ability to scan non-target DNA
in vivo
is of biological importance.
Micrococcus luteus,
a
Gram-positive cocci, produces an en-
zyme, UV endonuclease, which has several similarities to T4
endonuclease V. Both proteins are small molecular weight
proteins (18,000 and 16,000, respectively), and neither shows
any requirement for divalent cations
or
ATP. These enzymes
have been shown to incise UV-irradiated DNA at the site of
pyrimidine dimers by cleaving the 5’-glycosylic bond followed
by a type
I
apyrimidinic endonuclease activity which cleaves
the phosphodiester backbone between the two pyrimidines of
the dimer (26-33). This two-step mechanism is unique to
these enzymes. Due to the many similarities exhibited by
these two enzymes, we have investigated the
in vitro
DNA
scanning ability of the
M. luteus
UV endonuclease and how
this activity is affected by changes in pH as well as ionic
strength.
MATERIALS
AND
METHODS
Purification
of
M. luteus
UV
Endonuclease-Freeze-dried M. luteus
cells (ATCC 4689) were obtained from Sigma and resuspended in 500
ml of 10 mM Tris-HC1 (pH 7.5), 10 mM EDTA, and 25 mM
KCI.
Cells
17422
DNA Scanning Activity
of
M.
luteus
UV
Endonuclease
17423
were treated with lysozyme and sonicated briefly to shear the large
genomic
DNA.
Cellular debris was removed by centrifugation. The
UV
endonuclease was purified
to
homogeneity by sequentially using
the following chromatographic steps: single-stranded DNA-agarose,
Sephacryl
HR200,
and fast protein liquid chromatography Mono
S.
Details of this purification scheme are
to
be published elsewhere.
Kinetic Analysis
of
M.
luteus
UV
Endonuclease with UV-irradiated
DNA-Form
I
[3H]pBR32Z in 20 mM EDTA
(pH
8.0)
at
1
pg/d
was
irradiated by
short
wave (254
nm)
UV
light at
100
microwatts/cm*
for 245
s
in order
to
generate 25 pyrimidine dimers per plasmid
molecule. This DNA was then added
to
various buffers (described
below), at
a
final concentration of
0.1
pg/pl.
All
kinetic analyses were
performed at substrate concentrations below
the
K,,
as revealed hy
altering
the
reaction velocity by altering the number
of
pyrimidine
dimers per
molecule
(data not shown).
All
buffers contained 5%
glycerol,
0.1
mg/ml
BSA,’ and various concentrations of
NaC1.
The
pH
of
these buffers was modulated with one
of
the following:
10
mM
Tris-HC1 (pH 7.5),
20
mM
Tricine (pH
8.0),
or
20
mM
MES (pH
6.0).
M.
luteus
UV
endonuclease was added at appropriate concentrations
to
allow the reactions
to
be monitored
over
a
30-min kinetic analysis.
At the indicated time points,
1O-pl
aliquots of the reaction mixture
were removed and added to the appropriate stop buffer. The stop
buffers used contained
50%
sucrose, 2% SDS, 50
mM
Tris-HCl (pH
8),
20 mM EDTA, and bromphenol blue. Alkaline hydrolysis
was
achieved by adding
1
N
NaOH to this stop buffer
to
a final concen-
tration
of
1%.
The final pH
of
the reaction is approximately pH
11.
This pH results in the cleavage of the phosphodiester bond
at
apu-
rinic/apyrimidinic (AP) sites without denaturing the DNA. For the
analyses which were performed
using
denaturing gels, the
stop
buffer
contained 40% sucrose,
1%
SDS,
1
mM
EDTA, and
100
mM NaOH.
These conditions not only
cleave
the phosphodiester bond at
AP
sites,
but also denature the DNA. The methoxyamine stop buffer contained
200
mM
methoxyamine,
0.5%
SDS,
and 30% glycerol. Methoxyamine
treatment renders AP sites impervious to alkaline hydrolysis
(34),
and
thus
the
complete
incision activity of
the
enzyme
can
be moni-
tored utilizing denaturing agarose gels without
the
concomitant
hy-
drolysis
of
alkaline-labile sites.
In order
to
investigate the pH optimum for
the
enzyme, reactions
were performed in the following buffers at the indicated pH values:
Na+-acetate (4.5, 5.0), MES (5.0, 5.5,
6.0),
PIPES (6.0,
6.5),
MOPS
(6.5,
7.0),
HEPES
(7.0,
7.5),
Tricine
(7.5,
8.0,
8.5), CHES (8.5,
9.0),
and CAPS
(9.0).
A
wide range
of
buffers with overlapping pH values
were used
to
eliminate
the
possibility
of
buffer-specific effects. The
final concentration
for
all these buffers was 50
mM,
and the ionic
strength of the reactions was
100
mM
NaC1. Single time points were
taken after 20 min and terminated with and without the alkaline
nondenaturing treatment. These reaction products were
then
ana-
lyzed by agarose
gel
electrophoresis.
All
reaction products were resolved
by
electrophoresis in 1.2%
agarose gels. The
gels
were stained in
1
pg/ml
ethidium bromide in
electrophoresis buffer consisting
of 40
mM sodium
acetate
and
2
mM
EDTA. For the native and nondenaturing alkaline treatmehts,
the
DNA hands were visualized by exposure
to
long wave
UV
light. Form
I,
11,
and
III
DNA bands were excised from the gels and counted
by
liquid scintillation spectroscopy
(10).
In order
to
determine
the
DNA
fragment size distribution which was generated by t.he enzyme,
sam-
ples
were resolved by electrophoresis in denaturing agarose gels.
These gels were presoaked in 30
mM
NaOH and
1
mM
EDTA
overnight to
ensure complete
denaturation
of
the
DNA as well
as
alkaline hydrolysis
of
any remaining AP sites. Following the separa-
tion
of
single-stranded DNA fragments by electrophoresis, the dena-
turing
gels
were processed by the method
of
Southern (35, 36).
RESULTS
Investigation
of
the DNA Scanning Ability
of
M.
luteus
UV
Endonuclease-In order to investigate the in vitro DNA scan-
ning ability of the UV endonuclease, kinetic analysis of the
enzyme with UV-irradiation plasmid DNA was performed as
The abbreviations used are: BSA, bovine serum albumin; AP,
apurinic/apyrimidinic;
SDS,
sodium dodecyl sulfate; MES, 2-(N-
morpho1ino)ethanesulfonic
acid; PIPES,
piperazine-N,N-bis(2-eth-
anesulfonic acid),
MOPS,
3-(N-m0rpholino)propanesulfonic
acid;
HEPES,
N-2-hydroxyethylpiperazine-N”2-ethanesulfonic
acid;
CHES,
2-(cyclohexylamino)-ethanesulfonic
acid; CAPS, 3-(cyclohex-
y1amino)propanesulfonic acid.
described under “Materials and Methods.” For a reaction
mechanism in which an enzyme may function by
a
processive
nicking activity, plasmids should accumulate which have been
completely incised at all the dimer sites. Thus, one would
expect to see a significant amount of form
111
DNA generated
while there is still
a
large percentage of form
I
DNA remain-
ing. In a distributive
or
random hit mechanism, almost all the
form
I
DNA
should be converted to form
I1
DNA before
significant amounts of form
111
DNA are generated.
Fig.
1
shows the results of these analyses performed at salt
concentrations of
0,
25,
100,
and
250
mM NaC1. The rate of
loss of form I DNA is most rapid for the
25
and
100
mM NaCl
reactions and slowest
for
the
0
mM reaction, while the
250
mM reaction is intermediate. Under these conditions, no form
I11
was observed to accumulate for any of the salt concentra-
tions tested, a result which was contrary to what would be
expected if the enzyme was acting in a processive manner.
These reactions were also analyzed by determining the
distribution
of
single-stranded DNA reaction products on
denaturing agarose gels as described under “Materials and
Methods.” By utilizing denaturing gels, the size distribution
of
the DNA fragments generated by the enzyme can be
visualized. The size of the DNA fragments which were gen-
erated corresponds to the distance between incised dimer
sites. If small fragments were produced while significant
amounts of form
I
DNA still remain, this would indicate that
adjacent dimer sites have been cleaved and that the enzyme
was acting in a processive manner. Also, the alkaline condi-
tions which are present in the denaturing gels will cleave any
AP sites which were generated by the glycosylase activity and
not completely incised by the enzyme. The results of this type
of analysis are shown in Fig.
2.
Panel A shows the loss of
form
I
and form
I1
DNA
for time course reactions performed
at
25
mM and
100
mM NaC1. The DNA fragment size distri-
bution which was generated is not evident in panel
A
but
is
present in
a
longer exposure (panels
B
and
C).
The major
difference between the two reactions shown in panel A is that
for the
25
mM NaCl reaction the form
I
and form
I1
DNA are
lost at approximately the same rate throughout the time
course, whereas in the
100
mM NaCl reaction, the form
I
DNA is rapidly converted to form
I1
DNA, followed by a slow
loss in form
I1
DNA. Panel
B
shows the distribution of
fragment size generated for both the
25
and the
100
mM NaCl
reactions
at
the same
(20
min) time point. This shows that
for the
25
mM reaction (lane
B),
much smaller fragments are
being generated than those observed
for
the
100
mM reaction
(lane
C).
In addition, it should be pointed out that there is
still form
I
DNA remaining in the
25
mM reaction, while in
the
100
mM reaction all the form
I
DNA
has been incised to
generate form
I1
DNA. Panel
C
shows the distribution of
fragments generated for the
25
mM and
100
mM NaCl reac-
tions
at
the point where equal amounts
of
form
I
DNA are
remaining for both reactions. This type of comparison allows
the examination of how many single-stranded breaks (AP
endonuclease and alkaline hydrolysis) are generated per en-
zyme-DNA encounter. As seen in panel
C,
the
25
mM NaCl
reaction generated
a
distribution
of
smaller fragments relative
to that observed at the
100
mM NaCl reaction. This suggests
that for each enzyme-DNA encounter in the
25
mM NaCl
reaction, the enzyme is acting
at
the sites of several if not all
the dimers present in that molecule. By comparison, the
100
mM NaCl reaction has generated only form
I1
DNA suggesting
that for each enzyme DNA encounter, it is generating only
one
or
a very few single-stranded beaks.
Modulation
of
DNA
Scanning in Vitro by Ionic Strength-
Due to the apparent discrepancy between the fragment size
17424
DNA
Scanning Activity
of
M.
luteus
UV
Endonuclease
liK
lOOmM
NaCl
40
20
0
0
10
20
30
1
time (min)
OmM NaCl
il;-El
:E
m
FIG.
1.
Kinetic analysis
of
M.
lu-
teus
UV Endonuclease on UV-irra-
diated DNA.
Form
I
[:'H]pRR322 in 20
a
0
40
mM EDTA at
1
pg/pl was irradiated by
short wave (254 nm)
UV
light at
100
E
20
microwatts/cm'
for
245
s
in order to
e
generate 25 pyrimidine dimers per mol-
$0
ecule. This DNA was incubated with pu-
time tmin)
rified
M.
luteus
UV
endonuclease in time
E
2!
a
z
z
0
C
al
.-
c
e
0
10
20
30
40
P
course reactions. The three topological
forms of DNA were separated by agarose
gel electrophoresis and quantitated by
liquid scintillation counting. Reactions
were performed in
10
mM Tris-HCI (pH
T.5),
S%
glycerol,
0.1
mg/ml BSA, and
the indicated concentration of NaCI.
Symbols
used are: form
I.
0:
form
11,
0,
form
111,
m.
250mM
NaCl
rn
25mM
NaCl
40
20
C
e
$0
0
10 20
30
40
time (min)
a
z
0
0
10
20
30
40
time (min)
Panel
A
Panel
B
Panel
C
AB
25mM
time
(min)
0
5
10
15
20
30
lOOmM
time
(min)
0
5
10 15 20
30
ABC
-
form
I1
form
I
form
I1
form
I
form
II
m
Y
form
I
form
II
4
form
I
FIG.
2.
Analysis
of
the distribution
of
single-stranded breaks produced
by
M.
luteus
UV endonuclease
within a population
of
UV-irradiated DNA.
Panel
A,
kinetic analysis reactions identical with those shown in
Fig.
1
were performed, and the reaction products were separated on denaturing agarose gels. Reactions were
performed in
10
mM Tris-HCI (pH i.5), 5% glycerol,
0.1
mg/ml BSA, and the indicated NaCl concentration.
Panel
H,
identical time points from
panel
A
were selected for comparison of the 25 mM and
100
mM NaCl reactions.
Lane
A,
0
min.
Lane
B,
25 mM, 20 min.
Lane
c,
100
mM, 20 min.
Panel
C,
time points with the same amount of form
I
DNA remaining for both the 25 mM and
100
mM NaCl reactions
(panel
A)
were selected
for
side-by-side comparison.
Lane
A,
25 mM NaCI.
Lane
H,
100
mM NaC1.
generated in the denaturing gels (Fig.
2)
and the lack of form
I11
generated in the native agarose gels (Fig.
I),
it was hy-
pothesized that the AP endonuclease activity was much less
active than the glycosylase activity under these conditions
(Tris-HC1, pH
7.5).
A difference in the activity levels has
been reported previously
(37).
If this were the case, many
apyrimidinic sites would be created by the enzyme because
the phosphodiester bond between the pyrimidines would not
be incised to generate
a
complete break. In order to test this
hypothesis, we performed time course reactions identical with
those described in Fig.
1.
These reactions were terminated
with SDS and then treated with alkali to hydrolyze the
phophodiester bond at any remaining AP sites. Fig.
3
shows
a
kinetic analysis of the reaction products obtained under
these conditions. All the salt concentrations tested showed an
increase in the amount of form
I11
DNA that was generated.
However, the
0
and
25
mM NaCl reactions show
a
significant
accumulation of form
I11
DNA while there is still form
I
DNA
remaining; by contrast, the
100
and
250
mM reactions do not
start to accumulate significant amounts of form
I11
DNA until
nearly
all
the form
I
DNA has been converted to form
11.
This
is highly consistent with what would be expected for a pro-
DNA Scanning Activity
of
M.
luteus
UV
Endonuclease
17425
FIG.
3.
Kinetic analysis
of
M.
lu-
teus
UV
endonuclease using alka-
line hydrolysis.
Reactions identical
with those described for Fig.
1
were per-
OmM
NaCl
1i;bl
z
0
40
g
20
z
$0
-
0
10
20 30
40
time (min)
lOOmM
NaCl
z
n
40
g
20
n
0
10
20
30
40
time (min)
formed. After terminating the reactions
with
SDS,
alkaline hydrolysis of the re-
25mM
NaCl
250mM
NaCl
maining apyrimidinic sites was achieved
by adding NaOH to pH
11.
Symbols are;
form
I,
0;
form
11,.
form
111,
..
~1:f"-g
~
1f-fy"y
4
60
40
-
c
5
20
c
a
no
K
n
0
10
20
30
40
time (min)
0
10
20
30
40
time (min)
cessive reaction mechanism occurring at the lower
(0
and
25
mM) salt concentrations and a more distributive mechanism
occurring for the higher
(100
and
250
mM) salt concentrations.
Since most incisions at dimer sites are the glycosylic bond
scission rather than the combined glycosylase/AP incision,
the processive reaction refers only to the scanning of the
enzyme on non-target DNA which facilitates dimer binding
and glycosylic bond scission. This is in contrast to the T4
endonuclease V enzyme in which the complete nicking reac-
tion is processive.
Quantitation
of
Glycosylase Versus AP Endonuclease Actiu-
ity-In order to quantitate the difference between the activi-
ties of the glycosylase and the AP endonuclease, the -In of
the rate of form I loss for both the native and alkaline-treated
reactions were compared. Under distributive conditions
(100
mM NaCl), the -In of the rate of form I loss should equal the
average number of single-stranded breaks generated. By com-
parison, we were able to calculate that under these conditions
the glycosylase is approximately 3 to
5
times more active than
the AP endonuclease. This ratio appears to be constant over
the salt concentrations which were evaluated in these exper-
iments (data not shown). It should be noted that this value
varied over time and with different enzyme preparations
suggesting that the AP endonucleolytic activity is more labile
over time than is the glycosylase activity.
Quantitation
of
pH
Optimum in Vitro-In order to investi-
gate the possibility that the glycosylase and the associated
AP endonuclease might have different pH optimum in uitro,
reactions similar to those previously described for the UV
endonuclease were performed across a wide pH range utilizing
several different buffer systems. These results are shown in
Fig. 4. Panel
A
shows the percentage of form I DNA remaining
after
20
min for both native and alkaline-treated reactions.
All reactions were performed under conditions which result
in a random incision of dimers at pH
8.0
(100
mM NaCl).
There is an apparent pH optimum of between
6
and
8
for the
glycosylase activity, with a slightly larger loss of form I DNA
observed for the alkaline-treated samples. More striking are
the differences in the amount of form I11 DNA generated
during these reactions as shown in Panel
B.
Modulation
of
DNA
Scanning in Vitro
by
Changes in
pH-
The large amount of form I11 DNA generated at pH
6
led us
to postulate that the enzyme may be acting processively at
this lower pH in spite of the high
(100
mM NaC1) salt
conditions. In order to investigate this possibility, kinetic
analyses, identical with those previously described, were con-
ducted with the exception that the reactions were performed
at both pH
6
and pH
8
for the same salt concentrations
(0,
25,
100,
and
250
mM NaC1). The reaction products were
analyzed using denaturing agarose gels. Fig.
5
shows a com-
parison of the reaction products obtained for the
100
mM
NaCl reaction and both pH
6
and pH
8.
At pH
6,
the size of
the DNA fragments obtained is much smaller than for the
same reaction performed at pH
8.
A
qualitatively similar effect
was observed for the
25
mM NaCl reactions (data not shown).
There is no observable effect seen for the
0
and
250
mM NaCl
concentrations. This is due to the mechanistic limitations
placed on the reactions by the extreme ionic conditions which
cannot be modulated by changes in pH (data not shown).
Analysis
of
AP Endonuclease Activity Utilizing Methoxy-
amine-In addition to the above analysis, aliquots from the
same time courses were terminated with SDS and treated
with
100
mM methoxyamine which has the effect of rendering
any apyrimidinic sites impervious to alkaline hydrolysis (34).
This allows the DNA fragment size generated by the AP
endonuclease activity to be analyzed during denaturing aga-
rose gels. Fig.
6
shows side-by-side identical time points in
the presence and absence of methoxyamine for both the pH
6
and pH
8
reactions. It can be clearly seen for both the pH
6
and pH
8
reactions that many AP sites are being resolved
by the alkaline treatment in the absence of methoxyamine
and that these AP sites must be the result of the enzyme
making only the glycosylic bond cleavage at the site of a
dimer. These data also suggest that the AP endonuclease has
a pH optimum around
6.
This effect was observed for all NaCl
concentrations tested (data not shown).
DISCUSSION
Previous work (37) had suggested that the glycosylase ac-
tivity of M. luteus UV endonuclease is more active than the
associated AP endonucleolytic activity in uitro. Our results
strongly support this observation as shown by the differences
17426
DNA Scanning Activity
of
M.
luteus
UV
Endonuclease
Panel
A
80
100
80
60
40
20
0
4
5
6
7
8
9
10
PH
Panel
B
4
5
6
7
8
9
10
PH
FIG. 4.
Determination
of
the pH optimum
for
the glycosy-
lase
and
AP
endonuclease activities
of
M.
luteus
UV
endonu-
clease.
In order to investigate the pH optimum for the enzyme,
reactions similar to those described in Fig.
1
were performed in the
following buffers at the pH values indicated by parentheses: Na+-
acetate (4.5, 5.0). MES (5.0, 5.5, 6.0), PIPES (6.0, 6.5), MOPS
(6.5,
7.0),
HEPES
(7.0,
7.5), Tricine
(7.5,
8.0, 8.5), CHES (8.5,
9.0),
and
CAPS
(9.0).
The final concentration for these buffers was
50
mM,
and the ionic strength of the reactions was
100
mM NaCI. Single time
points were taken after
20
min and terminated with and without the
alkaline nondenaturing treatment. These reaction products were then
analyzed by agarose gel electrophoresis and liquid scintillation count-
ing.
Panel
A
(form
I
DNA):
0,
(-)-alkaline treatment;
.,
(+)-alkaline
treatment.
Panel
R
(form
I11
DNA):
0,
(-)-alkaline treatment;
B,
(+)-alkaline treatment.
in the native and alkaline-treated time course reactions (see
Figs.
1,3,
and 6). We suggest that this is due to the apparent
lability of the AP endonuclease over time. This is very similar
to what has been observed for endonuclease V
(31,33)
where
the AP endonuclease displays
a
greater fragility than the
glycosylase activity. In addition to the apparent lability of the
AP endonuclease, our results suggest that this activity has
a
pH
6
lOOmM
pH
8
lOOmM
time (min)
0
5
10
15
20
30
time (min)
0
5
10
15
20
30
FIG.
5.
Kinetic analysis
of
M.
luteus
UV
endonuclease
at
pH
6
and
8.
Time course reactions similar to those described in Fig.
1
were performed using either 20 mM MES (pH 6.0) or
20
mM Tricine
(pH 8.0) and the indicated NaCl concentration. The reaction products
were analyzed using denaturing agarose gels.
pH optimum around 6.0,
as
shown by the methoxyamine-
treated reactions in Fig. 6,
as
well
as
the peak of form
I
loss
observed for the nonalkaline-treated reactions shown in
panel
A,
Fig. 4. The glycosylase activity,
as
judged by the loss of
form
I
DNA after alkaline hydrolysis
(panel
A,
Fig. 4), appears
to have a fairly wide pH optimum of 6.0-8.0. These are both
very consistent with the pH optimum observed for endonu-
clease V, 6.0-8.5 for the glycosylase, and
a
more narrow
optimum of around 6 for the AP endonuclease
(33).
Taken
together, these results illustrate even further the many simi-
larities between the T4 endonuclease V and the
M.
luteus
UV
endonuclease.
Our results support the following model for the DNA scan-
ning ability of
M.
luteus
UV endonuclease on UV-irradiated
DNA
in
uitro.
For reaction conditions which enhance protein-
DNA interactions
(0-25
mM NaCl (pH 8.0) or 0-100 mM
NaCl (pH 6.0), UV endonuclease binds to non-target DNA
and is able to diffuse along this DNA, until it encounters
a
pyrimidine dimer. After binding to that site, the enzyme
makes
a
glycosylase cut in the DNA
at
the site of the dimer.
The enzyme may either catalyze the phosphodiester bond
scission, then continue to scan additional non-target DNA,
or it may be released from the dimer site without completing
the phosphodiester bond scission thus leaving the alkaline-
labile site. The enzyme continues to scan the remaining DNA
for any additional dimers. Eventually all alkaline-labile sites
are converted to single-stranded breaks by the enzyme. How-
ever, from these data, we are unable to directly show that this
secondary AP endonuclease activity is acting in
a
processive
manner. Thus, depending on the lability
of
the AP endonu-
clease activity, many single-stranded breaks might require
a
secondary encounter with
a
dimer site after initially making
the glycosylic cleavage before the AP endonuclease cleavage
can be made. However, the analysis of the glycosylase activity
shows the enzyme to be scanning non-target DNA. At low
(0
and
25
mM NaCl) ionic strength conditions, the enzyme is
able to remain associated with
a
given plasmid long enough
for all glycosylic bond scissions to be made
at
the dimer sites.
At higher (100 and
250
mM NaCl) salt concentrations, espe-
cially at pH
8,
the increased ionic strength probably has
increased the dissociation constant of the enzyme with non-
DNA Scanning Activity
of
M.
luteus
UV
Endonuclease
17427
pH
6
lOOmM
(-)
methoxyamine
0
5
10 15
20
30
time (min)
I1
I
pH
8
lOOmM
(-)
methoxyamine
0
5
10
15
20
30
time
(min)
pH
6
lOOmM
(+)
methoxyamine
time (min)
0
5
10 15
20
30
The smaller fragment size obtained at pH
6
suggests that the
enzyme is acting sequentially
at
adjacent dimer sites within
the population of DNA. If the dissociation rate of the enzyme
for non-target DNA is significantly less
at
pH
6
than at pH
8,
the enzyme would remain bound longer and have
a
higher
probability of diffusing
a
greater distance along the DNA to
act at adjacent dimer sites; the enzyme would be acting more
processively. Thus, the pH of the reaction
in vitro
appears to
modulate the degree of processivity exhibited by the
M.
luteus
UV
endonuclease. We hypothesize that this effect may be due
to the titration of certain amino acid residues of the enzyme,
possibly histidine, to a net positive charge. This increase in
the number of positively charged residues may result in an
overall increase in the affinity of the enzyme for the negatively
charged phosphodiester backbone of non-target DNA. This
hypothesis is supported by the mutagenesis data described
previously for endonuclease
V
(25).
To
our knowledge, this
pH-dependent modulation
of
processivity has not been de-
scribed for any other DNA scanning enzymes.
REFERENCES
1.
Ptashne, M.
(1986)
Nature
322,697-701
2.
Lohman,
T.
M.
(1986)
CRC
Crit.
Reu.
Biochem.
19,191-245
3.
von Hippel,
P.
H., and Berg,
0.
G.
(1989)
J.
Rhl.
Chem.
264,675-678
(+)
methoxyamine
4.
Riggs, A.
D.,
Bourgeois,
S.,
and Cohn,
M.
(1970)
J.
Mol.
Riol.
53,401-417
5.
Berg,
0.
G.,
Winter, R. B., and von Hippel,
P.
H.
(1981)
Blochemutry
20,
15
2o
3o
6.
Berg,
0.
G., Winter, R. B., and von Hippel,
P.
H.
(1982)
Trends
Biochem.
7.
Winter, R. B., and von Hippel,
P.
H.
(1981)
Biochemistry
20.6948-6960
8.
Winter, R. B., Berg,
0.
G.,
and von Hippel,
P.
H.
(1981)
Biochemistry
20,
9.
Barkley,
M.
D.
(1981)
Biochemistry
20,3833-3842
pH
8
lOOmM
time (min)
fi929-6948
Sci.
7.52-55
6961-6977
10.
Llovd.
R.
S..
Hanawalt.
P.
C.. and Dodson. M. L.
(1980)
Nucleic Acids
Res.
FIG.
6.
Kinetic analysis
of
M.
luteus
UV
endonuclease
(2)-
methoxyamine.
Reactions identical with those described for Fig.
5
were performed. After terminating the reactions, methoxyamine
(100
mM) was added to render any remaining apyrimidinic sites impervious
to alkaline hydrolysis. Reactions were performed at the indicated pH
and NaCl concentrations. The reaction products were analyzed using
denaturing agarose gels.
target DNA to the point where it does not remain associated
with a plasmid long enough for all dimers to be completely
incised. Instead, the enzyme acts by
a
more distributive or
random hit mechanism in which the length of non-target
DNA that it is able scan before dissociating is shorter than
the interdimer distance for that plasmid. We also propose
that this dissociation rate can be modulated
in
vitro
by
changes in the pH of the reaction. This is clearly shown by
the difference in the size of the DNA fragments generated by
the enzyme at pH
6
versus
pH
8
at
100
mM NaCl (see Fig.
5).
11.
12.
13.
14.
15.
16.
17.
18.
8,
g113-5127
Biochemistry
25,5751-5755
Ganesan, A.
K.,
Seawell,
P.
C., Lewis, R.
J.,
and Hanawalt,
P.
C.
(1986)
Jack. W.
E..
Terrv. B.
J..
and Modrich.
P.
(1982)
Proc.
Natf.
Acad. Sci.
U.
Gruskin,
E.
A., and Lloyd, R.
S.
(1986)
J.
Biof.
Chem.
261,9607-9613
S.
A.
79,'4010:4014
'
Lansowski.
J..
Alves.
J..
Pineuod.
A..
and Maass.
G.
(1983)
Nucleic Acids
~
Res.
11,'50i-510
Ehhrecht, H.-J., Pinguod,
A.,
Urbanke, C., Maass,
G.,
and Gualeni, C.
Nardone,
G.,
George,
J.,
and Chirkjian,
J.
G.
(1986)
J.
Biof.
Chem.
261,
Belinstev, B.
N.,
Zauriev,
S.
K.,
and Shemyakin. M.
F.
(1980)
Nucleic Acids
Hannon, R., Richards,
E.
G.,
and Gald, H.
J.
(1980)
EMBO
J.
5, 3313-
..
.,
..
...
(1985)
J.
Riol.
Chem.
260,6160-6166
12128-12133
Res.
8,1391-1403
X119
19.
ParkiC.
S.,
Wu,
F.
Y.-H., and Wu, C.-W.
(1982)
J.
Biof.
Chem.
257,6950-
20.
Roe, J.-H., and Record, M.
T.,
Jr.
(1985)
Biochemistry
24,4721-4726
21.
Singer, P., and Wu, C.-W.
(1987)
J.
Riol.
Chem.
262.14178-14189
22.
Singer,
P.
T.,
and Wu, C.-W.
(19M)
J.
Biol.
Chem.
263,4208-4214
23.
Wang,
J.
C., and Giaever,
G.
N.
(1988)
Science
240,300-304
24.
Gruskin,
E.
A,,
and Lloyd, R.
S.
(1988)
J.
Riol.
Chem.
263, 12728-12737
25. Dowd,
D.
R., and Lloyd, R.
S.
(1989)
J.
Mol.
Riol.,
in press
26.
Haseltine, W. A,, Gorden, L.
K.,
Lundan, C. P., Grafstrom.
R.
H., Shaper,
27.
Gordon, L.
K.,
and Haseltine,
W.
A.
(1980)
J.
Rtof.
Chem.
255, 12047-
29.
Seawell,
P.
C., Smith, C. A,, and Ganesen, A.
K.
(1980)
J.
Virol.
35, 790-
28.
Radany,
E.
H., and Friedberg,
E.
C.
(1980)
Nature
286,182-185
30.
McMillan,
S.,
Edenberg, H. J., Radany,
E.
H., Friedberg, R. C., and
31.
Nakaheuou.
Y..
and Sekieuchi.
M.
(1981)
Proc. Natf. Acad. Sci.
U.
S.
A.
6056
N.
L., and Grossman, L.
(1980)
Nature
285,634-641
12050
797
Friedberg.
E.
C.
(1981)
J.
Virof.
40,211-223
78,2'7i212746
204-210
257,2556-2562
.,
.
32.
Warner, H. R., Christensen, L. M., and Persson, M.-L.
(1981)
J.
Virof.
40,
33.
Nakabeppu, Y., Yamashita,
K.,
and Sekiguchi, M.
(1982)
J.
Biol.
Chem.
34.
Liuzzi,
M.,
Weinfield, M., and Paterson, M. C.
(1987)
Biochemistry
26,
35.
Southern,
E.
M.
(1975)
J.
Mol.
Biof.
98,503-517
36.
Wahl,
G.
M., Stern, M., and Stark,
G.
K.
(1975)
Proc.
Natf. Acad. Sci.
U.
37.
Grafstrom, R. H., Park, L., and Grossman, L.
(1982)
J.
Biof.
Chem.
257,
:3:115-:3321
S.
A.
76,3683-3687
13465-13474
... Characterization of an 18-kDa protein has shown that the UV endonuclease prefers thyminecontaining dimers over cytosine-containing dimers (under conditions of substrate excess), double-stranded over singlestranded DNA, and apyrimidinic sites at the site of glycosylase action to simple apurinic or apyrimidinic residues (2). The M. luteus UV endonuclease locates pyrimidine dimers, at least in vitro, by a processive sliding mechanism on nontarget DNA (8). The efficiency of this scanning is dependent on both ionic strength and pH. ...
... Its propensity not to dissociate prior to the ␤-elimination step may explain why the M. luteus enzymesubstrate complex was trapped more readily with NaBH 4 than the T4 endonuclease V imino intermediate. Other investigators have reported that the N-glycosylase activity of the M. luteus UV endonuclease is up to an order of magnitude greater than its apparent AP lyase activity, even in partially purified preparations that may be contaminated by multiple AP endonucleases (2,7,8). These findings are difficult to reconcile with our own unless, as noted by Hamilton and Lloyd (8), the AP lyase activity is more labile over time than is the N-glycosylase activity. ...
Article
Full-text available
Although Micrococcus luteus UV endonuclease has been reported to be an 18-kDa enzyme with possible homology to the 16-kDa endonuclease V from bacteriophage T4 (Gordon, L. K., and Haseltine, W. A.(1980) J. Biol. Chem. 255, 12047-12050; Grafstrom, R. [Abstract] H., Park, L., and Grossman, L.(1982) J. Biol. Chem. 257, 13465-13474), this study describes three independent purification schemes in which M. luteus UV damage-specific or pyrimidine dimer-specific nicking activity was associated with two proteins of apparent molecular masses of 31 and 32 kDa. An 18-kDa contaminant copurified with the doublet through many of the chromatographic steps, but it was determined to be a homolog of Escherichia coli ribosomal protein L6. Edman degradation analyses of the active proteins yielded identical NH2-terminal amino acid sequences. The corresponding gene (pdg, pyrimidine dimer glycosylase) was cloned. The protein bears strong sequence similarities to the E. coli repair proteins endonuclease III and MutY. Nonetheless, traditionally purified M. luteus protein acted exclusively on cis-syn thymine dimers; it was unable to cleave site-specific oligonucleotide substrates containing a trans-syn -I,), or Dewar thymine dimer, a 5,6-dihydrouracil lesion, or an A:G or A:C mismatch. The UV endonuclease incised cis-syn dimer-containing DNA in a dose-dependent manner and exhibited linear kinetics within that dose range. Enzyme activity was inhibited by the presence of NaCN or NaBH4 with NaBH4 additionally being able to trap a covalent enzyme-substrate product. These last findings confirm that the catalytic mechanism of M. luteus UV endonuclease, like those of other glycosylase/AP lyases, involves an imino intermediate.
... c, which is more likely to be electrostatic interactions rather than intercalation in experiments of Barton et al. Although they also demonstrate the dsDNA mediated electrochemistry of several iron-sulfur proteins[16], the salt concentration of the testing buffer they employ is much higher, to the extent that can prevent electrostatic interactions between proteins and DNA[30]. Therefore, there must be some other binding patterns between dsDNA and these proteins. ...
Article
Full-text available
In this work, we reported the investigation on the interaction between DNAstrands self-assembled at gold electrodes and an electron transfer protein, cytochrome c. Weobserved that cytochrome c exhibited well-defined electrochemistry in both double-strandedand single-stranded DNA films. This suggested that the electron transfer reaction ofcytochrome c arose possibly due to the electron hopping along DNA strands rather thanwiring along the double helix. We also compared the heterogeneous electron transfer rate ofcytochrome c with that of a ruthenium complex, which further confirmed this mechanism.
Chapter
To store the genetic information and serve as the genetic link between generations, the nucleotide sequence of DNA must be faithfully maintained despite the numerous physical or chemical insults discussed in the previous chapter. To that end, all organisms from bacteria to mammals are endowed with DNA repair mechanisms, and even some viral genomes carry their own DNA repair enzymes.
Article
Full-text available
This review will present a current understanding of mechanisms for the initiation of base excision repair (BER) of oxidatively-induced DNA damage and the biological consequences of deficiencies in these enzymes in mouse model systems and human populations.
Article
DNA repair proteins conduct a genome-wide search to detect and repair sites of DNA damage wherever they occur. Human alkyladenine DNA glycosylase (AAG) is responsible for recognizing a variety of base lesions, including alkylated and deaminated purines, and initiating their repair via the base excision repair pathway. We have investigated the mechanism by which AAG locates sites of damage using an oligonucleotide substrate containing two sites of DNA damage. This substrate was designed so that AAG randomly binds to either of the two lesions. AAG-catalyzed base excision creates a repair intermediate, and the subsequent partitioning between dissociation and diffusion to the second site can be quantified from the rates of formation of the different products. Our results demonstrate that AAG has the ability to slide for short distances along DNA at physiological salt concentrations. The processivity of AAG decreases with increasing ionic strength to become fully distributive at high ionic strengths, suggesting that electrostatic interactions between the negatively charged DNA and the positively charged DNA binding surface are important for nonspecific DNA binding. Although the amino terminus of the protein is dispensable for glycosylase activity at a single site, we find that deletion of the 80 amino-terminal amino acids significantly decreases the processivity of AAG. These observations support the idea that diffusion on undamaged DNA contributes to the search for sites of DNA damage.
Article
A general mechanism by which proteins locate their target sites within large domains of DNA is a one-dimensional facilitated diffusion process in which the protein scans DNA in a nonspecifically bound state. An electrostatic contribution to this type of mechanism has been previously established. This study was designed to question whether other characteristics of a protein's structure might contribute to the scanning mechanism of target site location. In this regard, T4 endonuclease V was shown to establish an ionic strength dependent monomer-dimer equilibrium in solution. A protein dimer interaction site was postulated to exist along a putative alpha-helix containing amino acid residues 54-62. The conservative substitutions of Phe-60----Leu-60 and Phe-59, Phe-60----Leu-59, Leu-60 resulted in mutant enzymes which remained in the monomeric state independent of the ionic strength of the solution. The target site location mechanism of these mutants has also been altered. Under conditions where wild-type endonuclease V processively scans nontarget DNA, the target location mechanism of the monomeric mutant proteins was shifted toward a less processive search. This decrease in the processivity of the mutants was especially surprising because the nontarget DNA binding affinity was found to be significantly increased. Thus, an additional component of the endonuclease V DNA scanning mechanism appears to be the formation of a stable endonuclease V dimer complex.
Article
Full-text available
Several DNA-interactive proteins, including the DNA repair enzyme T4 endonuclease V, have been shown to locate their target recognition sites utilizing an electrostatically mediated facilitated diffusion mechanism. Previous work indicates that a decrease in the affinity of endonuclease V for nontarget DNA results in an increased nontarget dissociation rate. This study was designed to investigate the effect of an increase in the affinity of endonuclease V for nontarget DNA. Using a working structural model of the enzyme as a guide, the electrostatic character of endonuclease V was altered. Substitution of Thr-7 with Lys-7 resulted in an enzyme with wild type in vitro characteristics. Mutations which increased the positive charge along a proposed solvent-exposed alpha-helical face had significant effects. The mutants Ala-30, Val-31----Lys-30, Leu-31 and Asn-37----Lys-37 displayed wild type in vitro apurinic-specific and dimer-specific nicking activities. Although the processive dimer-specific nicking rate of the Lys-37 mutant resembled that of wild type, the rate of the Lys-30, Leu-31 mutant was reduced by 60%. In addition, the salt concentration range over which these mutants processively nick dimer-containing DNA has been greatly expanded. Both mutants are shown to have an increased affinity for nontarget DNA.
Article
Full-text available
Facilitated diffusion along nontarget DNA is employed by numerous DNA-interactive proteins to locate specific targets. Until now, the biological significance of DNA scanning has remained elusive. T4 endonuclease V is a DNA repair enzyme which scans nontarget DNA and processively incises DNA at the site of pyrimidine dimers which are produced by exposure to ultraviolet (UV) light. In this study we tested the hypothesis that there exists a direct correlation between the degree of processivity of wild type and mutant endonuclease V molecules and the degree of enhanced UV resistance which is conferred to repair-deficient Eshcerichia coli. This was accomplished by first creating a series of endonuclease V mutants whose in vitro catalytic activities were shown to be very similar to that of the wild type enzyme. However, when the mechanisms by which these enzymes search nontarget DNA for its substrate were analyzed in vitro and in vivo, the mutants displayed varying degrees of nontarget DNA scanning ranging from being nearly as processive as wild type to randomly incising dimers within the DNA population. The ability of these altered endonuclease V molecules to enhance UV survival in DNA repair-deficient E. coli then was assessed. The degree of enhanced UV survival was directly correlated with the level of facilitated diffusion. This is the first conclusive evidence directly relating a reduction of in vivo facilitated diffusion with a change in an observed phenotype. These results support the assertion that the mechanisms which DNA-interactive proteins employ in locating their target sites are of biological significance.
Article
The purpose of this study was to determine the mechanism by which uracil DNA glycosylase locates uracil residues within double-stranded DNA. Using reaction conditions that contained low salt concentrations, the addition of uracil DNA glycosylase to plasmid DNAs containing multiple, randomly incorporated uracils resulted in the accumulation of form III DNA while unreacted form I DNA was still present. These data suggested that the enzyme utilizes a one-dimensional DNA-scanning mechanism such that this linear DNA arose by the accumulation of many single-strand breaks within the plasmid prior to enzyme dissociation. Reactions containing higher concentrations of uracil DNA glycosylase revealed a further accumulation of form III DNA after all form I DNA had been lost. These results suggested a partial (1.5-2 kb) enzyme processivity since the enzyme does not incise at all uracil bases on the DNA molecule prior to dissociation from that DNA. Since DNA scanning is regulated by electrostatic interactions, the processivity of the enzyme was evaluated through kinetic analyses of incision at various salt concentrations. At NaCl concentrations (< 50 mM), a significant amount of form III DNA accumulated while there were still unreacted form I DNAs present. In contrast, the accumulation of form III DNA was delayed at higher salt concentrations and the overall accumulation of form III DNA was less than that monitored at lower salt concentrations. DNAs were also analyzed by denaturing agarose gel electrophoresis in order to measure the average distance between strand breaks. Southern hybridizations showed a greater accumulation of breaks in DNAs that were reacted with the uracil DNA glycosylase at the lower salt concentrations, confirming a partial processivity for the enzyme.
Article
Full-text available
The action of the dimer-specific endonuclease V of bacteriophage T4 was studied on UV-irradiated, covalently-closed circular DNA. Form I ColEl DNA preparations containing average dimer frequencies ranging from 2.5 to 35 pyrimidine dimers per molecule were treated with T4 endonuclease V and analysed by agarose gel electrophoresis. At all dimer frequencies examined, the production of form III DNA was linear with time and the double-strand scissions were made randomly on the ColEl DNA genome. Since the observed fraction of form III DNA increased with increasing dimer frequency but the initial rate of loss of form I decreased with increasing dimer frequency, it was postulated that multiple single-strand scissions could be produced in a subset of the DNA population while some DNA molecules contained no scissions. When DNA containing an average of 25 dimers per circle was incubated with limiting enzyme concentrations, scissions appeared at most if not all dimer sites in some molecules before additional strand scissions were produced in other DNA molecules. The results support a processive model for the interaction of T4 endonuclease V with UV-irradiated DNA.
Article
Full-text available
We describe a technique for transferring electrophoretically separated bands of double-stranded DNA from agarose gels to diazobenzyloxymethyl-paper. Controlled cleavage of the DNA in situ by sequential treatment with dilute acid, which causes partial depurination, and dilute alkali, which causes cleavage and separation of the strands, allows the DNA to leave the gel rapidly and completely, with an efficiency independent of its size. Covalent attachment of DNA to paper prevents losses during subsequent hybridization and washing steps and allows a single paper to be reused many times. Ten percent dextran sulfate, originally found to accelerate DNA hybridization in solution by about 10-fold [J.G. Wetmur (1975) Biopolymers 14, 2517-2524], accelerates the rate of hybridization of randomly cleaved double-stranded DNA probes to immobilized nucleic acids by as much as 100-fold, without increasing the background significantly.
Article
Full-text available
In 1970 Riggs et al. (1) reported that Escherichia coli lac repressor binding to λ DNA in vitro seemed to find its target (operator) site on the DNA at a rate as much as 1000-fold faster than the upper limit estimated for a diffusion-controlled process involving macromolecules of this size. This observation startled and intrigued many physically oriented molecular biologists and biochemists and initiated a flurry of theoretical and experimental papers seeking to offer an explanation. However, scrutiny of the older literature reveals that scientists, ranging from mathematicians to biologists, had long been concerned with how systems of various sorts might transcend the rate limits set by three-dimensional diffusion control (2). Such problems are now of interest at many different levels. The pure physical chemist feels that an understanding of such phenomena might provide new insight into what happens when molecules meet and rearrange in the course of forming and passing through the transition state complex. The enzyme mechanician hopes that the secrets of some of the astonishing increases in rates achieved in enzyme-catalyzed reactions may be revealed by a study of these rate accelerations. And the cell biologist who studies macromolecular interactions and assembly processes is intrigued by the possibility that these systems may reveal opportunities for acceleration of intracellular rates beyond the limits set by the relatively slow diffusion of macromolecules in the cytoplasm. In this minireview we propose to touch on recent progress in all of these areas but will focus primarily on a problem that has engaged our attention over the past few years, i.e. how do protein regulators of gene expression at the transcriptional level find their regulatory DNA targets at speeds that appear to be faster than diffusion controlled?
Article
T4 endonuclease V [endodeoxyribonuclease (pyrimidine dimer), EC 3.1.25.1)], which is involved in repair of UV-damaged DNA, has been purified to apparent physical homogeneity. Incubation of UV-irradiated poly(dA).poly(dT) with the purified enzyme preparations resulted in production of alkali-labile apyrimidinic sites, followed by formation of nicks in the polymer. The activity to produce alkali-labile sites was optimal in a relatively broad pH range (pH 6.0-8.5), whereas the activity to form nicks had a narrow optimum near pH 6.5. By performing a limited reaction with T4 endonuclease V at pH 8.5, irradiated polymer was converted to an intermediate form that carried a large number of alkali-labile sites but only a few nicks. The intermediate was used as substrate for the assay of apurinic/apyrimidinic DNA endonuclease activity [endodeoxyribonuclease (apurinic or apyrimidinic, EC 2.1.25.2]. The two activities, a pyrimidine dimer DNA glycosylase and an apurinic/apyrimidinic DNA endonuclease, were copurified and found in enzyme preparations that contained only a 16,000-dalton polypeptide. An enzyme fraction from cells infected with bacteriophage T4v1, a mutant that is sensitive to UV radiation, was defective in both glycosylase and endonuclease activities. Moreover, occurrence of an amber mutation in the denV gene caused a simultaneous loss of the two activities, and suppression of the mutation rendered both activities partially active. These results strongly suggested that a DNA glycosylase specific for pyrimidine dimers and an apurinic/apyrimidinic DNA endonuclease reside in a single polypeptide chain coded by the denV gene of bacteriophage T4. Because the two activities exhibited different thermosensitivity, it was further suggested that conformation of the active sites for these activities may be different.
Article
A sensitive membrane filter assay for the lac repressor-operator complex has been used to study the kinetics of repressor and operator interaction. The rate of dissociation of the complex is very slow and follows first-order kinetics with a half-life of 19 minutes in our standard buffer (0.05 m ionic strength). The rate of dissociation (kb) is rather insensitive to temperature or pH, but becomes more rapid at high ionic strength. In 0.2 m ionic strength buffer the dissociation half-life is five to six minutes. The rate of association of repressor and operator is very fast and follows second-order kinetics with the rate constant for association (ka) being 7 × 109m−1 sec−1 in our standard buffer. The rate of association is only slightly affected by temperature or pH, but becomes slower with increasing ionic strength. It is concluded that the association of repressor and operator is aided by electrostatic attraction between the phosphate groups of DNA and a positively charged binding site on the lac repressor. Over a wide range of ionic strength, the ratio has been found to agree well with the equilibrium constant. The binding of repressor to operator results in an unfavorable enthalpy change; the driving force for the binding reaction comes from a large entropy increase. Binding mechanisms are discussed.
Article
The Escherichia coli lac repressor (R) locates its DNA target site, the operator (O), by a two-step mechanism. In the first step a repressor-DNA (RD) complex is formed at a non-operator site in an ordinary diffusion-controlled reaction. This non-specifically bound repressor then ‘slides’ along the DNA in a one-dimensional diffusion process. A series of intramolecular (within the ‘domain’ of the DNA molecule) dissociation-association steps, interspersed by sliding, eventually results in target location. Similar mechanisms may be used by other genome regulatory proteins in locating specific DNA binding sites.
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.
Article
The UV endonucleases [endodeoxyribonuclease (pyrimidine dimer), EC 3.1.25.1] from Micrococcus luteus and bacteriophage T4 possess two catalytic activities specific for the site of cyclobutane pyrimidine dimers in UV-irradiated DNA: a DNA glycosylase that cleaves the 5'-glycosyl bond of the dimerized pyrimidines and an apurinic/apyrimidinic (AP) endonuclease that thereupon incises the phosphodiester bond 3' to the resulting apyrimidinic site. We have explored the potential use of methoxyamine, a chemical that reacts at neutral pH with AP sites in DNA, as a selective inhibitor of the AP endonuclease activities residing in the M. luteus and T4 enzymes. The presence of 50 mM methoxyamine during incubation of UV- (4 kJ/m2, 254 nm) treated, [3H]thymine-labeled poly(dA).poly(dT) with either enzyme preparation was found to protect completely the irradiated copolymer from endonucleolytic attack at dimer sites, as assayed by yield of acid-soluble radioactivity. In contrast, the dimer-DNA glycosylase activity of each enzyme remained fully functional, as monitored retrospectively by release of free thymine after either photochemical- (5 kJ/m2, 254 nm) or photoenzymic- (Escherichia coli photolyase plus visible light) induced reversal of pyrimidine dimers in the UV-damaged substrate. Our data demonstrate that the inhibition of the strand-incision reaction arises because of chemical modification of the AP sites and is not due to inactivation of the enzyme by methoxyamine. Our results, combined with earlier findings for 5'-acting AP endonucleases, strongly suggest that methoxyamine is a highly specific inhibitor of virtually all AP endonucleases, irrespective of their modes of action, and may therefore prove useful in a wide variety of DNA repair studies.
Article
The rate of formation of transcriptionally competent open complexes between Escherichia coli RNA polymerase (RNAP) and the lambda PR promoter is extraordinarily sensitive to the nature and concentration of the electrolyte ions in the solution. The pseudo-first-order time constant of open complex formation tau obsd, determined in excess RNAP at 25 degrees C as a function of NaCl concentration, is proportional to the concentration product [Na+]12 [RNAP]-1. Consequently, tau obsd is far more sensitive to changes in the salt concentration than to changes in the concentration of RNAP. The origin of this effect is the release of the thermodynamic equivalent of 12 monovalent ions in the process of closed complex formation at the lambda PR promoter. In more complex ionic mixtures, ion-specific stoichiometric effects on tau obsd are observed. These are not ionic strength effects but are instead both valence and species specific. Both the association and dissociation rate constants of RNAP at the lambda PR promoter are strongly salt dependent, varying (in NaCl) as [Na+]-12 and [Na+]8, respectively. Consequently, the equilibrium constant characterizing open complex formation at this promoter varies with [Na+]-20. Electrostatic interactions and counterion release are the major contributors to the binding free energy driving open complex formation in a dilute salt solution. Since the in vivo ionic environment of E. coli (and other cells) is highly variable, these large salt effects are almost certainly of physiological significance. Variations in the intracellular concentrations of inorganic and organic ions, including polyamines, must exert both global and also promoter-specific regulatory effects on the initiation of transcription, as well as on numerous other protein-nucleic acid interactions.