ArticlePDF Available

Reversal of a Mutator Activity by a Nearby Fidelity-Neutral Substitution in the RB69 DNA Polymerase Binding Pocket

Authors:

Abstract and Figures

Phage RB69 B-family DNA polymerase is responsible for the overall high fidelity of RB69 DNA synthesis. Fidelity is compromised when conserved Tyr567, one of the residues that form the nascent polymerase base-pair binding pocket, is replaced by alanine. The Y567A mutator mutant has an enlarged binding pocket and can incorporate and extend mispairs efficiently. Ser565 is a nearby conserved residue that also contributes to the binding pocket, but a S565G replacement has only a small impact on DNA replication fidelity. When Y567A and S565G replacements were combined, mutator activity was strongly decreased compared to that with Y567A replacement alone. Analyses conducted both in vivo and in vitro revealed that, compared to Y567A replacement alone, the double mutant mainly reduced base substitution mutations and, to a lesser extent, frameshift mutations. The decrease in mutation rates was not due to increased exonuclease activity. Based on measurements of DNA binding affinity, mismatch insertion, and mismatch extension, we propose that the recovered fidelity of the double mutant may result, in part, from an increased dissociation of the enzyme from DNA, followed by the binding of the same or another polymerase molecule in either exonuclease mode or polymerase mode. An additional antimutagenic factor may be a structural alteration in the polymerase binding pocket described in this article.
Content may be subject to copyright.
Reversal of a Mutator Activity by a Nearby
Fidelity-Neutral Substitution in the RB69 DNA
Polymerase Binding Pocket
Anna Trzemecka
1
, Agata Jacewicz
1
, Geraldine T. Carver
2
,
John W. Drake
2
and Anna Bebenek
1
1
Department of Molecular Biology, Institute of Biochemistry and Biophysics, Polish Academy of Sciences,
Pawinskiego 5a, 02-104 Warsaw, Poland
2
Laboratory of Molecular Genetics, National Institute of Environmental Health Sciences, Research Triangle Park,
NC 27709, USA
Received 29 July 2010;
received in revised form
27 September 2010;
accepted 28 September 2010
Available online
13 October 2010
Edited by M. Gottesman
Keywords:
Exo;
Pol
Phage RB69 B-family DNA polymerase is responsible for the overall high
fidelity of RB69 DNA synthesis. Fidelity is compromised when conserved
Tyr567, one of the residues that form the nascent polymerase base-pair
binding pocket, is replaced by alanine. The Y567A mutator mutant has an
enlarged binding pocket and can incorporate and extend mispairs
efficiently. Ser565 is a nearby conserved residue that also contributes to
the binding pocket, but a S565G replacement has only a small impact on
DNA replication fidelity. When Y567A and S565G replacements were
combined, mutator activity was strongly decreased compared to that with
Y567A replacement alone. Analyses conducted both in vivo and in vitro
revealed that, compared to Y567A replacement alone, the double mutant
mainly reduced base substitution mutations and, to a lesser extent,
frameshift mutations. The decrease in mutation rates was not due to
increased exonuclease activity. Based on measurements of DNA binding
affinity, mismatch insertion, and mismatch extension, we propose that the
recovered fidelity of the double mutant may result, in part, from an
increased dissociation of the enzyme from DNA, followed by the binding of
the same or another polymerase molecule in either exonuclease mode or
polymerase mode. An additional antimutagenic factor may be a structural
alteration in the polymerase binding pocket described in this article.
© 2010 Elsevier Ltd. All rights reserved.
Introduction
The DNA polymerases (gp43 s encoded by gene 43)
of the related bacteriophages T4 and RB69 synthesize
DNA with high fidelity, inserting one wrong nucle-
otide per 10
5
replicated bases.
1
These polymerases
belong to the B-familythe same family that is
populated by the eukaryotic replicative polymerases
α,δ,andɛ. For RB69 DNA polymerase, crystal
structures are available for the apo enzyme,
2
abinary
editing complex with DNA bound in the exonuclease
domain,
3
and ternary complexes (replicating com-
plex) with DNA and dNTP bound in the polymerase
site.
4
Because of the availability of crystal structures in
so many conformations, RB69 polymerase is a good
structural model for studying the mechanisms re-
sponsible for high base selectivity.
RB69 DNA polymerase achieves its fidelity by the
coordinated action of two activities: a polymerase
(Pol) activity that is responsible for correct
*Corresponding author. E-mail address:
aniab@ibb.waw.pl
Abbreviations used: dsDNA, double-stranded DNA;
ssDNA, single-stranded DNA.
doi:10.1016/j.jmb.2010.09.058 J. Mol. Biol. (2010) 404, 778793
Contents lists available at www.sciencedirect.com
Journal of Molecular Biology
journal homepage: http://ees.elsevier.com.jmb
0022-2836/$ - see front matter © 2010 Elsevier Ltd. All rights reserved.
nucleotide selection and an exonuclease (Exo)
activity that removes mismatched nucleotides from
the primer terminus. As with all polymerases that
replicate DNA with high fidelity, binding the correct
incoming dNTP introduces conformational changes
in both the enzyme and the DNA. The fingers
subdomain undergoes the largest conformational
change, moving toward the palm subdomain to
enclose the incoming dNTP and the complementary
template nucleotide and thus to generate a binding
pocket for the nascent base pair.
47
In the RB69 polymerase, the active Pol binding site
is formed by several highly conserved residues
located in the fingers and palm. Arg482 and Lys560
in the fingers are involved in hydrogen bonding to
the phosphate groups of the incoming dNTP.
Hydrophobic interactions with the nascent base
pair are formed by Leu415 and Tyr416 on the minor
groove side and by Tyr567 and Gly568 on the
template side. Asn564 and Ser565 form the rear wall
of the binding pocket and ensure a coplanar base
arrangement of the nascent base pair (M. Wang et al.,
unpublished results).
1,4,8
All of these residues play important roles in RB69
DNA replication fidelity. Leu415 and Tyr567 are
critical for substrate discrimination. L415F/G mutants
have elevated rates for a variety of base substitution
errors and single-base deletions in repetitive and
nonrepetitive sequences.
8
Replacing Tyr567 by Ala,
Ser, or Thr confers a strong mutator phenotype,
particularly for base substitution errors.
1
Tyr416 is
responsible for discriminating between ribose moiety
and deoxyribose moiety in incoming nucleotides,
9
while Leu561 discriminates sterically against mis-
matches in the nascent base-pair binding pocket.
10
Thus, the size and shape of the polymerase binding
pocket play crucial roles in replication fidelity.
Insertion of a wrong nucleotide compromises the
rate of primer extension. The polymerase stalls,
altering the balance between extension by the poly-
merase and excision by the mismatch-editing exonu-
clease. Newly incorporated mismatches reduce the
efficiency of subsequent nucleotide insertions and
extensions by 10
2
-fold to 10
6
-fold,
6,7,11,12
with the
magnitude depending on the mismatch and the
polymerase.
Polymerases with compromised fidelity due to
alterations at the Pol site may display reduced
partitioning to the Exo site and thus favor mismatch
extension over proofreading, notably with the T4
Pol
L412M
and RB69 Pol
L415F/G
mutants.
8,13
The same
mechanism probably applies to the RB69 Pol
Y567A
mutant, as previously proposed.
1
Conversely, T4
polymerase mutants (such as T4
A737V
) that display
an antimutator phenotype, mainly reducing
A·TG·C rates, tend to display a decreased ability
to translocate after nucleotide incorporation, thus
favoring the formation of exonuclease complexes
and increasing the efficacy of proofreading.
14,15
In
preliminary screens, we observed that the RB69
S565G replacement had only weak impacts on
fidelity; however, surprisingly, when the Y567A
and S565G replacements were combined within the
highly conserved B-motif KX
3
NSXYG, the powerful
mutator activity of the Pol
Y567A
mutation was
strongly reduced. To better understand the role of
Ser565 in fidelity, we characterized some of the
biochemical properties of RB69 polymerase mutants
containing S565G replacement, Y567A replacement,
or both replacements. We also investigated the
impact of these replacements using fidelity assays
both in vivo and in vitro. Our results led us to
propose that introducing the S565G replacement
lowers the DNA binding affinity of the polymerase
and increases its ability to dissociate from the
primer/template, thus increasing the opportunity
for proofreading by the same or another polymerase
molecule (proofreading in trans) and contributing to
the antimutagenic impact of S565G on Y567A.
Results
Bacteriophage RB69 gp43 Y567 and S565 are
highly conserved residues within the B-family of
DNA polymerases and are located in the region
involved in dNTP binding. We showed previously
that Y567 plays a crucial role in base selection:
when Y567 was replaced with A, S, or T, the
mutant polymerase displayed severely decreased
fidelities both in vivo and in vitro.
1
To study the role
of S565, we replaced it with G in both Exo
+
and
Exo
backgrounds. We also examined the combi-
natorial effect of the double replacement Y567A/
S565G. Most of the assays were performed using
Exo
+
derivatives because of the surprising anti-
mutator effect of the double replacement in this
background.
Polymerase activity in vivo and specific activity
in vitro
In assays in vivo, we often used a hybrid system in
which T4, whose own gp43 is inactivated by amber
mutations at codons 202 and 386 (T4 43amam), is
replicated by wild-type or mutant RB69 gp43 s
expressed from plasmids.
1
Polymerase activity in
vivo was determined only for Exo
+
derivatives.
DNA synthesis by T4 gp43 was totally blocked by a
43amam mutation and performed instead by one of
the RB69 DNA polymerases expressed from a
plasmid. DNA synthesis was measured in infected
Escherichia coli cells at a time when host DNA
synthesis had ceased completely.
1
DNA synthesis
was moderate in the presence of S565G either alone
(30% of wild-type activity) or in combination with
Y567A (40% of wild-type activity), while the Y567A
mutant retained 50% activity.
779Suppression of a Mutator DNA Polymerase
We also measured polymerase-specific activity in
vitro for the RB69 DNA polymerases using activated
DNA as substrate. The specific activities for the Exo
+
versions of Pol
Y567A
, Pol
Y567A/S565G
, and Pol
S565G
were 80%, 72%, and 47%, respectively, of the Pol
+
Exo
+
value.
Fidelity in vivo
In rII reversion assays (Table 1), RB69 Pol
S565G
has
almost negligible impact on replication fidelity for
both base substitutions and frameshifts; the average
change in mutation rates was a 2-fold increase
compared to the wild-type enzyme. Similarly, in rI
forward mutation assays (Table 2), Pol
S565G
in Exo
+
background had a mutation rate elevated by a factor
of 5 compared to the wild-type enzyme; in Exo
background, no significant difference (rate decreased
0.8-fold) was observed between parental enzymes
and mutant enzymes. Surprisingly, the double
mutant (Pol
S565G/Y567A
Exo
+
) had strongly decreased
mutator activities compared to the Pol
Y567A
Exo
+
mutant in rII reversion assays: the second replace-
ment (at S565) suppressed most of the mutator traits
of the Y567A replacement. The most striking results
were the 90-fold decrease in base substitution
mutation rates at the UV375 A·T site and the 60-fold
decrease in base substitution mutation rates at the
UV256 G·CA·T site, while 1 and +1 frameshift
mutations were reduced by 50-fold and 6-fold,
respectively, at other sites (Table 1). In rI forward
mutation assays, the Pol
S565G/Y567A
double mutant
decreased mutation rates by 40-fold compared to the
single Pol
Y567A
mutant (Table 2). In Exo
background,
the Pol
S565G/Y567A
mutant did not support the
production of viable T4 43amam progeny, perhaps
because of a lethally high mutation rate from the
combination of the mutator effects of the Y567A
replacement and the Exo deficiency
1,16
acombina-
tion that the S565G mutation may reduce insufficiently
to restore viability.
We sequenced independent rI mutants from experi-
ments in which T4 43amam phage was replicated by
the RB69 polymerase mutants Pol
S565G
Exo
+
and
Pol
S565G/Y567A
Exo
+
, respectively. The kinds of muta-
tions made by these polymerases are shown in
Table 3, together with the kinds obtained previous-
ly for the Pol
+
and Pol
Y567A
enzymes.
1,16
Mutations
produced by the Pol
S565G
Exo
+
polymerase were
very similar to those produced by the Pol
+
Exo
+
polymerase; the complex mutations were almost all
GCGCTA at the 146148 hot spot.
1618
There
were no strong differences between the kinds of
mutations produced by Pol
S565G/Y567A
Exo
+
poly-
merases and the kinds of mutations produced by
Pol
Y567A
Exo
+
polymerases, each making predom-
inantly base substitutions. GA predominated
among transition mutations, with as much as 70%
localized at a hot spot at rI base 247. CTand
CA mutations at another hot spot at rI base 202
were also present in the spectra produced by both
Table 1. Mutation rates for RB69 Exo
+
DNA polymerase mutants in rII reversion assays
rIItester mutation
tester mutation
Mutation
scored
Reversion rate× 10
8
Pol
+
[17] Pol
S565G
[67] Pol
Y567A
[7] Pol
S565G/Y567A
[1314]
131 +1 9.0 (1) 4.3 (0.5) 37 (4.0) 5.6 (0.6)
UV232 1 1.0 (1) 3.3 (3.3) 210 (210) 4.4 (4.5)
UV256 G·CA·T 6.4 (1) 15 (2.4) 7200 (1100) 130 (20)
UV375 A·Tany 1.6 (1) 0.5 (0.3) 2600 (1700) 28 (18)
Values in brackets are the numbers of stocks providing the median rate. Values in parentheses below each rate are relative to those for
Pol
+
. The entries for Pol
+
and Pol
Y567A
were adapted from a previous article.
1
Table 2. Forward mutation rates of RB69 DNA
polymerase mutants in vivo
Polymerase
Number of
stocks
Mutation
rate× 10
5
Relative
rate
Pol
+
Exo
+
7 0.43 1
Pol
S565G
Exo
+
7 2.1 5
Pol
Y567A
Exo
+
7 210 490
Pol
S565G/Y567A
Exo
+
11 5.5 13
Pol
+
Exo
7 220 510
Pol
S565G
Exo
7 180 430
The entries for Pol
+
Exo
+
, Pol
+
Exo
, and Pol
Y567A
Exo
+
were
adapted from previous articles.
1,14
Table 3. Classes of rI mutations generated by Exo
+
RB69
DNA polymerase mutants
Polymerase
Pol
+
Pol
S565G
Pol
Y567A
Pol
S565G/Y567A
Number of mutations 103 53 79 45
A·TG·C 7 0 25 5
G·CA·T 22 6 43 27
Transitions 29 6 68 32
A·TT·A 0 2 0 0
A·TC·G 2 0 2 3
G·CT·A 34 22 8 7
Transversions 36 24 10 10
±1 10 6 1 2
±2 7 2 0 0
Complex
GCGCTA
21 15 0 1
The Pol
+
and Pol
Y567A
entries were adapted from previous
articles.
1,17
780 Suppression of a Mutator DNA Polymerase
Pol
Y567A
Exo
+
and Pol
S565G/Y567A
Exo
+
polymerases
(Fig. 1). Only AG mutations were underrepre-
sented in this spectrum, with only one appearing,
while 11 were found in the Pol
Y567A
Exo
+
spectrum
(Table 3). Thus, the decreased mutation rates in the
double mutant hardly affected specificity but
merely lowered the overall mutation rate.
Fidelity in vitro
To learn more about the role of Ser565 in fidelity, we
conducted studies in vitro using the lacZαDNA
template in gapped M13mp2 DNA (Table 4). The
lacZαmutant frequencies for the Pol
+
Exo
+
and
Pol
S565G
Exo
+
mutants were 10
4
and 4×10
4
,
respectively, with the values indistinguishable from
each other, similar to the 5× 10
4
to 7× 10
4
frequen-
cies typically obtained with gapped DNA,
16,19
and
representing only the mutations that accumulated
during the growth of the M13 phage stock. The lacZα
mutant frequency for the Pol
Y567A
Exo
+
mutant in
earlier studies was 55× 10
4
, and we observed a 5-fold
decrease in mutation frequency when the DNA was
copied by the Pol
S565G/Y567A
Exo
+
polymerase. This
frequency (10× 10
4
) was close to that of the Pol
+
Exo
+
polymerase. These results were fully consistent
with those obtained in our fidelity studies conducted
in vivo.
The lacZαmutant frequency for the Pol
S565G
Exo
polymerase was 14×10
4
, modestly lower that the
29× 10
4
frequency obtained previously with the Pol
+
Exo
enzyme. The Pol
S565G/Y567A
Exo
polymerase
was more proficient in proofreading some errors than
was the Pol
Y567A
Exo
polymerase (Table 4;89×10
4
versus 164× 10
4
). However, the Pol
S565G/Y567A
Exo
polymerase retained a high mutation frequency,
which probably resulted in lethally mutated progeny
during growth in vivo.
The contribution of S565G to the fidelity of
Pol
S565G/Y567A
As previously reported,
16
the RB69 Pol
Y567A
Exo
+
polymerase is a strong base substitution mutator in
vitro, mostly for transition mutations (especially for
TC produced by T·G mispairs) but also for
frameshift mutations in homopolymeric runs,
while frameshift mutations in runs were almost
absent in vivo.ThePol
S565G/Y567A
Exo
+
double
mutant also produced base substitutions, again
PolS565G/Y567A Exo+
T A T A
ATGGCCTTAAAAGCAACAGCACTTTTTGCCATGCTAGGATTGTCATTTGTTTTATCTCCATCGATTGAAGCGAATGTCGATCCTCATTTTGATAAATTTA 100
PolY567A C T + C T C G
Exo+ C
A
A +
T T A + T CTA C
TGGAATCTGGTATTAGGCACGTTTATATGCTTTTTGAAAATAAAAGCGTAGAATCGTCTGAACAATTCTATAGTTTTATGAGAACGACCTATAAAAATGA 200
C AG A C G C A C
C A G C
C
AAAAA
T AAA
AT AAA
AT C A
AT C A A
ATAC A AT G
CCCGTGCTCTTCTGATTTTGAATGTATAGAGCGAGGCGCGGAGATGGCACAATCATACGCTAGAATTATGAACATTAAATTGGAGACTGAATGA 294
ATACA A G T C A AT T G G G
AT C G T AT T G
T C A G
T C AAAAA
TT AAAAA
TTT AAAAA
AAAAAA
Fig. 1. rI mutational spectra for the RB69 Pol
S565G/Y567A
Exo
+
and Pol
Y567A
Exo
+
polymerases. The sequence
represents the wild-type rI strand complementary to the coding strand, with every 10th base in boldface. Upper-case
letters indicate base substitutions. The addition of a single base is indicated by +.The underlined three-base
mutation at residues 146148 is a hot spot for complex mutations. The entries for Pol
Y567A
were adapted from a
previous article.
1
Table 4. Fidelities of RB69 polymerases in lacZαforward
mutation tests in vitro
Pol Exo MF× 10
4
Mutations
Base
substitutions 1 Other
Pol
+
+ 8.4 22 20 2
Pol
S565G
+ 3.6 ——
Pol
Y567A
+ 55 147 110 37
Pol
S565G/
Y567A
+ 10.4 100 55 31 14
Pol
+
29 159 93 40 26
Pol
S565G
14 48 24 15 9
Pol
Y567A
164 154 120 28 6
Pol
S565G/
Y567A
89 103 75 25 3
MF, mutation frequency. Mutations are numbers detected by
sequencing; mutants from the Pol
S565G
Exo
+
polymerase were not
sequenced because the mutant frequency was at the historical
background for uncopied DNA. Othermutations consisted of
larger deletions and a few complex mutations.
The data for Pol
+
and Pol
Y567A
were adapted from a previous
article.
17
781Suppression of a Mutator DNA Polymerase
mostly transition mutations. This polymerase also
produced frameshift mutations and large deletions
not seen in the Pol
Y567A
Exo
+
spectrum (Table 5).
Among 12 large deletions, 11 (of 27371 nucleotides)
occurred between direct repeats. Most of the
frameshift mutations occurred at runs of two or
more repeated bases.
The roster of mutational classes showed a
significant decrease in T C transitions with the
double replacement compared to that with
Pol
Y567A
alone. There were at least five sequence
contexts where T C mutations either were absent
or appeared only once in the double-mutant
spectrum but were strongly represented in the
single-mutant spectrum (Fig. 2a). Overall, base
substitution rates decreased by 14-fold in the
double mutant, with the decrease in T·dGMP
rates being as much as 35-fold compared to the
Pol
Y567A
polymerase (Table 5). The other transition
mutation rates decreased 7-fold for C·dATP, 8-fold
for G·dTMP, and 5-fold for A·dCMP. Transversion
mutations were less frequent in the spectra of both
polymerases. A·dAMP rates decreased 14-fold,
G·dGMP rates decreased 7-fold, and C·dTMP rates
decreased 6-fold; for the remaining mismatches, the
mutation rates decreased 3-fold to 1.5-fold (Table 5).
Frameshift mutation rates decreased about 8-fold in
thedoublemutantcomparedtothesinglemutant.
We did not sequence lacZαmutants from the
Pol
S565G
Exo
+
polymerase background because the
mutation frequency (4× 10
4
) for this polymerase was
at the level of the historical background frequency
(5.7×10
4
to 6.2×10
4
) for unfilled template DNA.
This low mutation frequency probably recorded the
intrinsic mutation frequency resulting from phage
M13 replication and made further analysis of the
Pol
S565G
Exo
+
polymerase unreliable.
The contribution of proofreading to the fidelity of
the Pol
S565G/Y567A
Exo
+
polymerase
The lacZαmutation frequency for the Pol
S565G/Y567A
Exo
polymerase (89× 10
4
) was 6-fold higher that
that for the Pol
S565G
Exo
polymerase (14× 10
4
)and
about 2-fold lower than that for the Pol
Y567A
Exo
mutant (164× 10
4
). The Pol
S565G/Y567A
Exo
poly-
merase had reduced nucleotide selectivity compared
with the parental Pol
+
Exo
polymerase, but was less
error prone than the Pol
Y567A
Exo
polymerase.
Proofreading improved fidelity 8-fold in the
Pol
S565G/Y567A
Exo
+
polymerase but only 3-fold in
the Pol
Y567A
Exo
+
polymerase (Table 4). Proofreading
in the Pol
S565G/Y567A
double mutant removed both
base mismatches and frameshift mutations. The error
rates for all 12 possible base mismatches over all
detectable sites in the lacZαtemplate and for Exo
+
and
Exo
single and double Pol mutants are presented in
Table 5.InExo
background, the spectra for the single
and double mutants showed no striking differences,
and the distribution of mutations along the lacZα
sequence revealed no significant differences between
the two enzymes (Fig. 2b and c).
Table 5. RB69 DNA polymerase mutation specificities in lacZαforward mutation tests in vitro
Mutations
Pol
Y567A
Exo
+
Pol
S565G/Y567A
Exo
+
Pol
+
Exo
Pol
S565G
Exo
Pol
Y567A
Exo
Pol
S565G/Y567A
Exo
147 100 159 48 154 103
nμnμnμnμnμnμ
Base substitutions 110 47 55 3.4 93 17 24 4.4 120 160 75 78
AG (A·dCMP) 2 6 3 1.3 1 1 2 2.5 0 97 7
GA (G·dTMP) 7 18 6 2.2 8 9 2 2.1 54 420 20 122
CT (C·dAMP) 17 38 17 5.4 45 42 9 8.5 12 82 1 37
TC (T·dGMP) 76 150 15 4.3 13 11 4 3.4 37 230 33 158
Transitions 102 60 41 3.5 67 17 17 4.3 103 187 61 85
AC (A·dGMP) 0 3 2 0.9 1 1 0 1.4 1 10 1 8
AT (A·dAMP) 2 5 0 0.35 1 1 1 1.2 2 15 2 12
TG (T·dCMP) 0 200.35 0 101075 29
TA (T·dTMP) 2 7 5 2.5 0 101.5 8 85 4 33
GT (G·dAMP) 1 2 4 1.3 15 14 4 3.8 2 14 1 5
GC (G·dGMP) 2 6 2 0.8 7 8 1 1.2 4 34 1 6
CA (C·dTMP) 1 3 1 0.47 1 1 0 1.4 0 10 1 7
CG (C·dCMP) 0 5.8 0 0.9 1 3 1 2.6 0 19 0 15
Transversions 8 2.9 14 0.7 26 4.1 7 1.1 17 19 15 13
± 1 37 10 31 1.3 40 5 15 1.8 28 24 25 17
(residues 2436) 0 55 12 96 23 540 8 189 5 850 3 401
Other 0 55 2 16 3 71 1 27 1 170 0 134
Mutation rates (μ) are incorporated per 10
6
nucleotides. values were calculated as if one mutant had been detected. Mutation rates for
the Pol
+
Exo
+
and Pol
S565G
Exo
+
polymerases are not displayed because of the lack of signals significantly above the historical
background. Values for Pol
Y567A
Exo
+
, Pol
+
Exo
, and Pol
Y567A
Exo
were adapted from a previous article.
16
The number of detectable
sites is not defined for deletions N1 nt, and their mutation rates are therefore not normalized to the number of opportunities and thus only
appear to be larger than other normalized values.
782 Suppression of a Mutator DNA Polymerase
Exonuclease activities of single and double Pol
mutants
One possible explanation for the observed de-
crease in mutation rates in the Pol
S565G/Y567A
mutant
would be increased exonuclease activity. We there-
fore conducted exonuclease assays on substrates
with either a correct base pair or a T·dGMP
mismatch at the 3terminus to compare the various
polymerase variants. However, both Pol
S565G
and
Pol
S565G/Y567A
polymerases were less efficient than
the wild-type and Pol
Y567A
polymerases in
Fig. 2. lacZαmutational spectra. The 53sequence of the viral template strand of the lacZαsequence in M13mp2 is
shown from position 84 through position + 197, where + 1 is the first transcribed base and every 10th base is underlined.
Upper-case letters indicate base substitutions. The deletion of a single base is indicated by Δ,whereas the addition of a
single base is indicated by +.Large deletions and complex mutations are not shown. (a) Spectra for Pol
Y567A
Exo
+
(above the lacZαsequence) and Pol
S565G/Y567A
Exo
+
(below the lacZαsequence). The five lacZαbase triplets in red are
positions where significant differences were observed between the two spectra. (b) Spectra for Pol
Y567A
Exo
(above) and
Pol
S565G/Y567A
Exo
(below). (c) Spectra for Pol
S565G/567A
Exo
+
(above) and Pol
S565G/Y567A
Exo
(below). The spectra for
Pol
Y567A
were adapted from a previous article.
17
783Suppression of a Mutator DNA Polymerase
degrading double-stranded DNA (dsDNA) with
either a correct primer terminus base pair or an
incorrect primer terminus base pair (Fig. 3a and b).
The exonuclease activity on single-stranded DNA
(ssDNA) was indistinguishable among the Pol
+
,
Pol
Y567A
, and Pol
S565G/Y567A
polymerases, but was
slightly reduced in the Pol
S565G
polymerase (Fig. 3c).
DNA binding activity
Decreased exonuclease activity may result from a
change in the dynamic equilibrium between the
formation of complexes at the Pol active site and the
formation of complexes at the Exo active site.
13
A
mismatch at the 3terminus usually increases Exo
activity by weakening Pol binding to the substrate.
7
Using gel retardation assays, we measured the
strength of dsDNA binding for Pol
+
,Pol
Y567A
,
Pol
S565G
, and Pol
S565G/Y567A
polymerases in Exo
+
backgrounds with either a normal terminal base pair
or a terminal T·dGMP mismatch using the same
substrates as for the exonuclease assays (Fig. 4aand
b). K
d(DNA)
values were then determined from
reciprocal plots of enzyme concentration (Table 6).
The Pol
S565G
polymerase had the highest dissociation
constants when binding correctly paired DNA
(K
d(DNA)
=38.9 nM) or the mismatched substrate (K
d
(DNA)
=132 nM). These dissociation constants were
about 6-fold higher than for the Pol
+
polymerase and
3-fold higher than for the Pol
Y567A
polymerase on
normal dsDNA, and 3-fold higher for the Pol
+
polymerase and 5-fold higher for the Pol
Y567A
polymerase on dsDNA with the mismatch. The
Pol
S565G/Y567A
polymerase also bound less strongly
to both substrates compared to the Pol
+
and Pol
Y567A
polymerases (K
d(DNA)
=29.9 nM for normal dsDNA
and K
d(DNA)
=79.9 nM for mismatched dsDNA,
compared with K
d(DNA)
=6.9 nM and K
d(DNA)
=
13.2 nM for Pol
+
and Pol
Y567A
polymerases on
correctly paired dsDNA, and K
d(DNA)
=38.2 nM and
K
d(DNA)
=25.7 nM for Pol
+
and Pol
Y567A
polymerases
on mismatched dsDNA, respectively).
Weaker DNA binding in the Pol site does not
correlate with decreased dsDNA exonuclease activ-
ity for the Pol
S565G
or Pol
S565G/Y567A
polymerase.
Thus, both enzymes probably dissociate from either
normal or mismatched primer/template substrates
more often that do either the Pol
+
polymerase or the
Pol
Y567A
polymerase.
Primer-extension activities
The error rate of the Pol
S565G/Y567A
Exo
+
polymer-
ase is decreased for almost all mispairs compared to
the Pol
Y567A
Exo
+
polymerase. These differences are
also observed in Exo
background and thus may not
simply reflect partitioning ratios between the Pol sites
and the Exo sites. Analysis of mutation spectra for the
Pol
S565G/Y567A
Exo
+
and Pol
Y567A
Exo
+
polymerases
showed that, among transition mutations, the biggest
difference was in extending T·dGMP mismatches
(Table 5). Therefore, we used standing-start and
running-start assays to measure the abilities of the
Pol
+
,Pol
Y567A
, and Pol
S565G/Y567A
polymerases (all in
Exo
+
backgrounds) to insert dGTP opposite T and to
extend the T·dGMP mismatch.
The Pol
+
and Pol
S565G
polymerases misinsert
dGTP inefficiently opposite a template T (Fig. 5)
and also extend a T·dGMP mismatch inefficiently
(Fig. 6). The Pol
Y567A
polymerase both generates
and extends T·dGMP mismatches more efficiently
(Figs. 5 and 6). The Pol
S565G/Y567A
polymerase is
Fig. 2 (continued)
784 Suppression of a Mutator DNA Polymerase
much less efficient in both insertion and extension
(Figs. 5 and 6). The same pattern was observed in
running-start assays: the Pol
+
and Pol
S565G
poly-
merases were unable to detectably insert dGTP
opposite template T. The Pol
Y567A
polymerase both
inserted and extended a mismatch, while the
Pol
S565G/Y567A
polymerase was again much less
efficient in both steps (Fig. 7). The Exo
variants of
the Pol
+
, Pol
S565G
, and Pol
S565G/Y567A
polymerases
were similarly less efficient in both activities, in
contrast to the Pol
Y567A
polymerase, which inserted
and extended the T·dGMP mismatch efficiently
(data not shown).
The Pol
S565G/Y567A
Exo
polymerase supports
the replication of T4 phage in vivo
It was reported previously that the Pol
Y567A
Exo
polymerase was unable to support the in vivo growth
of either T4 43amam or T4 43
+1
. This dominant-lethal
phenotype was ascribed to the low fidelity of DNA
synthesis conducted by this enzyme, resulting in
Fig. 3. Exonuclease activities of RB69 DNA polymerases on ssDNA using a
32
P-labeled 20-mer, on normal dsDNA
using a
32
P-labeled 20-mer/35-mer, and on dsDNA containing a terminal G·T mismatch using a
32
P-labeled 20-mer/35-
mer. Reactions were incubated at 37 °C for 10 s, 20 s, 40 s, 80 s, and 300 s. Products of degradation were analyzed by
electrophoresis on a 15% polyacrylamide gel with 7 M urea, followed by autoradiography. Exonuclease assays on (a)
normal dsDNA, (b) dsDNA with a terminal G·T mismatch, and (c) ssDNA.
785Suppression of a Mutator DNA Polymerase
lethally mutated progeny phage. In the complemen-
tation assay, the Pol
S565G/Y567A
Exo
polymerase is
also unable to support the growth of T4 43amam but
can support the growth of T4 43
+
(Fig. 8), suggesting
that Pol
+
can compete sufficiently with Pol
S565G/Y567A
to produce some viable progeny whose plaques are,
however, small compared to T4 plaques growing on
BB cells harboring a plasmid expressing Pol
+
Exo
or
Pol
S565G
Exo
.
Discussion
Reversal of a powerful mutator activity by a
nearby fidelity-neutral substitution
The phage RB69 DNA polymerase mutant Pol
Y567A
displays a powerful mutator activity that can exceed
10
3
-fold at some sites.
1,16
In the crystalstructure of the
RB69 DNA polymerase ternary complex, Y567 dis-
plays an unfavorable rotamer conformation that is
maintained by hydrogen-bonding interactions with
two nearby residues.
20
These interactions are essential
for forming a nucleotide binding pocket. The aromatic
phenyl ring of Y567 plays an important role in base
discrimination, forming a hydrogen bond with the
minor-groove edge of the DNA duplex at the primer
terminus and thus helping to check WatsonCrick
base-pair geometry. Removing the γ-hydroxyl from
Y567 (Y567F) disrupts the hydrogen-bonding net-
work, allows Y567F to adopt a more favorable
rotamer conformation that distorts the nucleotide
binding pocket, reduces the affinity of incoming
dNTP, and blocks DNA replication.
20
On the other
hand, replacing the phenolic side chain with a methyl
group (Y567A) produces a vigorous polymerase that
is a powerful base substitution mutator both in vivo
and in vitro and a moderate mutator for frameshift
mutations in vitro. The Pol
Y567A
mutant polymerase is
better able to accommodate noncanonical base-pair-
ings, and the YA replacement does not interfere
with the geometry of correctly paired bases.
1,16
Thus,
Y567 plays a major role in base discrimination by
providing a steric gate to check pairing (Fig. 9).
The role of Ser565 in maintaining fidelity has come
under scrutiny only recently. Because Ser565 con-
tributes to the rear wall of the binding pocket, we
expected the S565G replacement to enlarge the size of
the pocket, thus more readily accepting base mispairs
and displaying mutator activity. However, the
Pol
S565G
polymerase displays, at most, very weak
mutator and antimutator activities both in vivo and in
vitro (Tables 15). We also anticipated that the double
mutant Pol
S565G/Y567A
polymerase would contain a
further expanded nucleotide binding pocket com-
pared with the Pol
Y567A
polymerase and would
therefore display further increased mutation rates.
Surprisingly, however, the Pol
S565G/Y567A
polymerase
displayed sharply decreased mutation rates both in
vivo and in vitro compared to the Pol
Y567A
polymerase
(Tables 15). On average, Pol
S565G/Y567A
was 50-fold
more accurate than Pol
Y567A
in rII reversion assay and
40-fold more accurate in rI forward mutation assay.
L561 is another residue that contributes to the
geometry of the Pol pocket and, when reduced in
bulk by the L561A replacement, displays modest
mutator activities in mutation tests in vivo and in
kinetic assays in vitro.
10
In contrast to the antimuta-
genic action of S565G in the Pol
S565G/Y567A
context,
we observed no antimutagenic effect of L561A in the
Pol
L561A/S565G
context (data not shown). Combining
the L561A and Y567A replacements produced a
Fig. 3 (continued)
786 Suppression of a Mutator DNA Polymerase
polymerase with exceptionally high mutation rates,
and the strong mutator phenotype was again
suppressed when S565G was added to the combi-
nation (data not shown).
Associated biochemical parameters
We searched for altered behaviors that might
explain the sharply contrasting fidelities of the
Pol
S565G/Y567A
and Pol
Y567A
polymerases. Polymer-
ase activities were modestly reduced both in vitro
and in vivo, but not in informative ways. Proofread-
ing estimated by comparing mutation rates in the
Exo
+
and Exo
versions of the mutant polymerases
(Table 4), and exonuclease activities on either
dsDNA or ssDNA (Fig. 3) were uninformative as
to the fidelity paradox. In mispair formation and
extension assays in vitro under both standing-start
and running-start conditions, the Pol
+
, Pol
S565G
, and
Pol
S565G/Y567A
polymerases were inefficient, while
the Pol
Y567A
polymerase was substantially more
efficientresults consistent with, but not explain-
ing, the mutation patterns. Pre-steady-state kinetic
data showed that the Pol
+
, Pol
S565G
, and Pol
S565G/
Y567A
polymerases display high K
d,app
values for
both G·dTMP and T·dGMP mispairs, while the
Pol
Y567A
polymerase has reduced K
d,app
values for
both mispairs.
22
Our present kinetic data for the
Pol
+
and Pol
Y567A
polymerases are in good agree-
ment with published data.
Fig. 4. RB69 DNA polymerase binding to dsDNA with normal base-pairing or a terminal G·T mismatch. The 5-labeled
double-stranded substrates (5 nM) were incubated with indicated RB69 polymerases at 5 nM, 10 nM, 25 nM, 50 nM,
100 nM, and 200 nM. After gel electrophoresis, the bands corresponding to bound or free DNA were quantified by
autoradiography. (a) Binding to normal dsDNA. (b) Binding to mismatched dsDNA.
Table 6. DNA binding affinities of Exo
+
RB69 DNA
polymerase mutants
Polymerase
K
d(DNA)
(nM)
[dsDNA]
K
d(DNA)
(nM) [dsDNA, T·G
mismatch]
Pol
+
6.9± 2.1 38.2 ± 4.3
Pol
Y567A
13.2± 1.4 25.7 ± 3.2
Pol
S565G
38.9± 3.4 132.1 ± 5.7
Pol
S565G/
Y567A
29.9± 1.9 79.9 ± 4.8
787Suppression of a Mutator DNA Polymerase
An article (M. Wang, et al., unpublished results)
explores the kinetic parameters of mispair formation,
extension, and proofreading using templates that
mimic the sequences of either a mutational hot spot or
a region of apparently low mutation rates in the rI
mutation reporter used here. Although not addres-
sing the question on the antimutagenic properties of
the S565G replacement, the assays do largely repro-
duce in vitro the relative mutation rates observed by
us in vivo as a function of local sequence, the first such
demonstration known to us. Because spontaneous
mutation rates vary greatly from site to site (for
instance, by about 10
3
-fold in phage T4
23
), rII
reversion rates are expected to vary considerably
among sites, and mutation spectra typically show a
wide range of site-specific rates of forward mutation.
Thus, it is often desirable to use a diversity of
substrates for kinetics studies of polymerase accuracy.
Based on measurements of DNA binding using
gel retardation assays with correct or mispaired
termini, the Pol
S565G
and Pol
S565G/Y567A
poly-
merases displayed weaker binding compared to
the Pol
+
or Pol
Y567A
polymerases (Fig. 4 and Table 6).
Weaker DNA binding for both Pol
S565G
and
Pol
S565G/Y567A
polymerases compared to Pol
+
and
Pol
Y567A
polymerases may reflect less specific
polymeraseDNA interactions whose in vivo con-
sequences could be manifested in a less efficient
DNA synthesis, but direct extrapolation to replica-
tion in vivo is difficult and requires caution. Using a
Fig. 5. T·dGTP mismatch forma-
tion by RB69 DNA polymerases. (a)
A
32
P-labeled primer/template was
present at 50 nM in all reactions.
The sequence of a portion of the
substrate is shown on the left with
an incoming dGTP (dG). For each
polymerase, 0.05 nM enzyme and
1 mM dGTP were incubated for
1 min, 3 min, and 5 min. A band at
the +1 position indicates insertion
of dGMP, while bands below the
primer are degradation products.
(b) Percentages of mismatch inser-
tion and primer degradation are
plotted against reaction time for
each of the polymerases.
788 Suppression of a Mutator DNA Polymerase
different assay and different interaction conditions,
no such difference were observed (M. Wang, et al.,
unpublished results).
Candidate mechanisms for the antimutagenic
activity of S565G
One possible explanation for the increased fidelity
of the Pol
S565G/Y567A
polymerase compared to
Pol
Y567A
is the increased partitioning of the mis-
matched primer-terminus to the Exo site. Diverse
changes in the Pol site can perturb the tight
coordination between the Pol cycle and the Exo
cycle,
6,24
and changing Y567 to alanine not only
enlarges the nucleotide binding pocket but probably
affects Pol/Exo partitioning in favor of the Pol site,
1,16
increasing mismatch extension and thus decreasing
proofreading. Conversely, proofreading may become
more proficient with any inhibition in polymeriza-
tion, as might be caused by mutations in the Pol site
that reduce the ability to form polymerizing com-
plexes and thus shift the balance in favor of Exo
complexes.
14,15
(A side effect is that increased
proofreading lowers discrimination between mis-
matched primer termini and matched primer-termini
and increases the degradation of newly synthesized
DNA.
25,26
) Because both Pol
Y567A
and Pol
S565G/Y567A
enzymes bind less tightly todsDNA substrates (Fig. 4
and Table 6), they are expected to form exonuclease
complexes more readily and, as a consequence, to
display increased exonuclease activities. However,
we observed no such increase in exonuclease activity.
One explanation for this failure would be increased
polymerase dissociation from the primer/template
DNA (whether or not mismatched), followed by
rebinding to DNA to form either an Exo complex in
Fig. 6. T-dGMP mismatch exten-
sion by RB69 DNA polymerases. (a)
A
32
P-labeled primer/template
with a terminal T-dGMP mismatch
was present at 50 nM in all reac-
tions. The sequence of a portion of
the substrate is shown on the left.
For each polymerase, 0.05 nM en-
zyme and 1 mM correct dGTP were
incubated for 1 min, 3 min, and
5 min. (b) Percentages of mismatch
extension and primer degradation
are plotted versus reaction time for
each of the polymerases.
789Suppression of a Mutator DNA Polymerase
the presence of a mismatch or a Pol complex and
continued replication. T4 DNA polymerase molecules
do exchange during DNA replication,
27
and RB69
wouldbeexpectedtodothesame.Frequent
dissociation will increase the opportunity to excise a
mismatch by either the original polymerase molecule
or another polymerase molecule, and thus may be
responsible for the increased fidelity of replication by
Pol
S565G/Y567A
compared to Pol
Y567A
.Thisexplana-
tion for the increased fidelity of Pol
S565G/Y567A
compared to Pol
Y567A
is supported by our observation
of decreased DNA binding by the Pol
Y567A
and
Pol
S565G/Y567A
enzymes, but is challenged by the
failure to observe such a difference (M. Wang, et al.,
unpublished results). However, the conditions used
Fig. 7. Running-start reactions for T·dG mismatch formation and extension for the indicated RB69 DNA Exo
+
polymerases. A
32
P-labeled primer/template was present at 50 nM in all reactions. A partial sequence of the substrate is
shown on the left. The reaction conditions were the same as for the standing-start reactions, and the reaction time was
5 min. dGTP was added at the indicated concentrations (μM). RB69 Pol
Y567A
not only inserts dGTP opposite template T
(band +3) but extends the mismatch, creating the next correct G·C base pair (band +4). Neither the Pol
+
enzymes nor the
Pol
S565G
enzymes can create the mismatch, but terminate synthesis at band + 2. The Pol
S565G/Y567A
double mutant has a
sharply reduced ability to insert, extend, and create a second mismatch compared to Pol
Y567A
.
Fig. 8. Complementation assay. Phage growth images
are spot tests in which 20,000 (left) or 100 (right) viable
particles of either T4 43am or 43
+
were spotted onto lawns
of cells carrying a plasmid producing the indicated Exo
RB69 gp43.
Fig. 9. A view of the major groove edge of the nascent
dTTP·dA base-pair binding pocket illustrating the posi-
tions of Ser565 and Tyr567 in α-helix P (wide gray ribbon).
nindicates the template dA, and n1indicates the
nucleotide upstream of the templating nucleotide. The
active-site metals are represented as green spheres A and
B. The arrow indicates the upstream direction of the DNA
duplex. The narrow gray ribbon represents the duplex
DNA backbone (residues omitted for clarity). The image
was created in Chimera
21
from Protein Data Bank ID 1IG9.
790 Suppression of a Mutator DNA Polymerase
in the two measurements were different, so that
decreased binding remains a possibility in vivo.
Note also that a more frequent dissociation from
the primer/template would also tend to result in the
production of large deletions, as observed in vitro in
the absence of the gp45 processivity clamp (Table 5).
The inhibition of a large deletion mutagenesis
between direct repeats by accessory proteins in
vitro was observed previously for RB69 DNA
polymerase and yeast DNA polymerase δ.
16,28
Another explanation for the increased fidelity of the
Pol
S565G/Y567A
polymerase compared to Pol
Y567A
is
provided by the wealth of crystallographic data (M.
Wang, et al., unpublished results). One key observa-
tion is that the replacement of the bulky Y567 side
chain by alanine reduces the rigidity of the Pol pocket
in the Y567A single mutant by disrupting the
hydrogen-bonding network involving the OH groups
of Y567, Y391, and T587, so that the templating base
can be displaced downward to accommodate mis-
pairs. When the S565G replacement was introduced
into the Y567A mutant, base selectivity was increased
compared to Pol
Y567A
even though the volume of the
Pol pocket was somewhat greater in the Pol
Y567A/S565
double mutant than in Pol
Y567A
. This appeared to be
due to an increase in the hydrophobic van der Waals
interactions of G565 with the templating base, making
it more rigid.
The two candidate explanations for the antimuta-
genic impact of S565G on Y567A are not mutually
exclusive.
Materials and Methods
Construction, expression, and purification of RB69
DNA polymerase variants
Plasmids pCW19R and pCW.50R were generous gifts
from Jim Karam (Tulane University). Plasmid pCW19R
carries a wild-type RB69 gene 43 (encoding the gp43
polymerase) under the control of the T7 Φ10 promoter of
the cloning vector pSP72 (Promega). Plasmid pCW.50R
encodes a Pol
+
Exo
polymerase carrying the exonuclease-
inactivating replacements D222A and D327A. Site-direct-
ed mutagenesis to create the Pol
S565G
, Pol
Y567A
, and
Pol
S565G/Y567A
variants was carried out using the Strata-
gene QuikChange Site-Directed Mutagenesis protocol and
was confirmed by sequencing. Expression and purifica-
tion of the Pol
S565G
, Pol
Y567A
, and Pol
S565G/Y567A
gp43 s
were performed as previously described.
17
Mutation tests in vivo
Reversion tests using T4 rII mutations were performed
as described previously.
1
T4 43amam rII mutants were used
to infect E. coli BB cells carrying a plasmid expressing the
desired allele of RB69 gene 43.rII131 reverts by + 1
(A5A6), rIIUV232 reverts by 1 (A3 A2), rIIUV256
reverts by G·CA·T transitions, and rIIUV375 reverts
probably by both transitions and transversions at three
adjacent A·T sites. Forward mutation rI assays were also
performed as described previously.
1
To determine the
kinds of rI mutations introduced in vivo by various RB69
polymerase mutants, we collected mutants of independent
origin, we resuspended mutant plaques in 40 μl of water,
and we amplified the rI gene by PCR and sequenced it as
described previously.
1
When sequencing was not per-
formed, a historical correction factor of 0.64 was used to
estimate the frequencies of rI mutants among all rmutants.
Mutation tests in vitro
Mutation tests in vitro were performed with phage
M13mp2 lacZα-gapped substrates prepared as described
previously.
19
The incubation mixture (25 μl) contained
25 mM Tris-acetate (pH 7.5), 10 mM Mg-acetate, 150 mM
K-acetate, 2 mM dithiothreitol, 150 ng of M13mp2 lacZα-
gapped substrate, 1 mM of each dNTP, and 710 pmol of
wild-type or mutant polymerase. Reactions were incubated
at 37 °C for 30 min, stopped by addition of ethylene-
diaminetetraacetic acid, and analyzed by agarose gel
electrophoresis. Products of reactions in which gap
filling was complete were introduced into MC1061
cells and plated on CSH50 cells to score M13 plaques
as either wild-type (dark blue) or mutant (white to less
dark blue). A representative set of mutants from most
collections was sequenced to determine the types of
errors and to adjust the mutation frequency for rare
phenotypic mutants without a mutation in the lacZα
reporter and for mutants with more than one detectable
mutation; a historical correction factor of 0.95 was
applied to mutant frequencies when sequencing was
not performed. Mutation rates (μ) were calculated as
described previously
19
by multiplying the net mutant
frequency [adjusted for the historical background
(6.2× 10
4
) of uncopied lacZαDNA] by the proportion
of mutants in each class and dividing by 0.6 (a correction
factor for detecting errors in E. coli) and by the number of
detectable sites (opportunities) for each class of mutation.
DNA synthesis in vivo
T4 43amam was used to infect E. coli BB cells carrying a
plasmid expressing the desired allele of RB69 gene 43.At
22 min after infection, 20μCi/ml [
3
H]thymidine was
added (at a specific activity of 20 μCi/μg dT). After
15 min,
3
H labeling was terminated in an ice bath, and
trichloroacetic-acid-precipitable
3
H counts were deter-
mined. Measurements of DNA synthesis in vivo were
performed as previously described.
1
Polymerase-specific activity
Polymerase-specific activity was measured by the incor-
poration of [α-
32
P]dATP (Hartmann Analytic) into activated
calf thymus DNA (Sigma), as previously described.
8
DNA substrates
All nucleotides were purchased from the Laboratory of
DNA Sequencing and Oligonucleotide Synthesis (Institute
791Suppression of a Mutator DNA Polymerase
of Biochemistry and Biophysics, Warsaw) and purified on
a 20% polyacrylamide gel with 7 M urea. All primers were
5end labeled with [γ-
32
P]dATP (5000 Ci/mmol; Hart-
mann Analytic) using T4 DNA polynucleotide kinase
(Takara). Annealing of primer to template was performed
in 10 mM TrisHCl (pH 7.5) at a primer/template molar
ratio of 1:1.3. The primer/template was annealed at 75 °C
for 5 min and allowed to cool slowly to 25 °C.
DNA gel retardation assay
The interactions of RB69 wild-type or mutant gp43 s
with dsDNA were assayed using 5-end-labeled 20-mer
(5-TTTGATGTATTATCAATTGT-3)hybridizedtoa
1.3 molar excess of complementary 35-mer (5-
TGCCTTCGTAATCTTACAATTGATAATACATCAAA-
3). The interactions of RB69 polymerases with a mispaired
primer terminus were assayed using a 5-end-labeled 20-
mer (5-ATGTGCTGCAAGGCGATTAG-3) hybridized to
a 30-mer (5-GTTACCCAACTTAATCGCCTTGCAGCA-
CAT-3). The 10-μl reaction mixture contained 10 mM
Hepes (pH 8.0), 0.5 mM dithiothreitol, 1 μg of bovine
serum albumin, 10 mM ethylenediaminetetraacetic acid
(EDTA), 5 nM DNA substrate, and increasing concentra-
tions of gp43. After incubation for 5 min at C, products
were separated on a precooled 6% polyacrylamide
nondenaturing gel and quantitated on a Fuji Phosphor-
Imager. To determine K
d(DNA)
, we plotted the reciprocal of
the fraction of DNA shifted as a function of enzyme
concentration.
Exonuclease activity
Assays of dsDNA and ssDNA 35exonuclease
activities were performed as previously described.
17
Exo
activity on mismatched DNA was assayed under the same
conditions using 5 nM gp43 and 50 nM
32
P5-end-labeled
20-mer (5-ATGTGCTGCAAGGCGATTAG-3) annealed
to 30-mer (5-GTTACCCAACTTAATCGCCTTGCAGCA-
CAT-3).
Extension of a primer-terminus mismatch
The assay for the extension of a terminal T·dGMP
mismatch contained 50 nM
32
P5-end-labeled 20-mer (5-
ATGTGCTGCAAGGCGATTAG-3) annealed to 30-mer
(5-GTTACCCAACTTAATCGCCTTGCAGCACAT-3).
The 10-μl reaction mixture contained 25 mM Tris-acetate
(pH 7.5), 10 mM Mg-acetate, 2 mM DTT, 0.05 nM gp43,
and 1 mM dGTP. Reactions were incubated at 37 °C for
1 min, 3 min, and 5 min, and then quenched by addition of
stop-dye solution. Product bands were resolved by
electrophoresis on a 16% polyacrylamide gel with 7 M
urea, analyzed by autoradiography, and quantified by
densitometry.
Misinsertion assay
The substrate for the misinsertion of dGMP opposite
template T under standing-start conditions was
32
P5-
end-labeled 20-mer (5-ATGTGCTGCAAGGCGATTAC-
3) annealed to a complementary 27-mer (5-GTAAGATG-
TAATCGCCTTGCAGCACAT-3). The reaction mixture
contained 25 mM Tris-acetate (pH 7.5), 10 mM Mg-acetate,
2 mM DTT, 50 nM DNA substrate, 0.05 nM gp43, and
1 mM dGTP. Reactions were incubated at 37 °C for 1 min,
3 min, and 5 min, and then quenched by addition of stop-
dye solution.
Running-start assays were conduced in the same
reaction buffer. The 10-μl reaction mixture contained
50 nM
32
P5-end-labeled 20-mer (5-ATGTGCTG-
CAAGGCGATTAC-3) annealed to a complementary 27-
mer (5-GAACTCCGTAATCGCCTTGCAGCACAT-3),
0.05 nM gp43, and increasing concentrations of dGTP
(250 μM, 500 μM, and 1000 μM). Reactions were incubated
at 37 °C for 5 min and quenched by the addition of 5 μlof
stop-dye solution. Product bands were resolved by
electrophoresis on a 16% polyacrylamide gel with 7 M
urea, analyzed by autoradiography, and quantified by
densitometry.
Acknowledgements
We thank Bill Beard for providing Fig. 9 and
both Bill Beard and Mike Murray for critical
readings of the manuscript and Bill Konigsberg
for advice on the implications of his structural
models. This research was supported, in part, by
funds allocated to project number Z01ES061054 of
the Intramural Research Program of the National
Institutes of Health, National Institute of Environ-
mental Health Sciences, USA, and by grant
N301014433 from the Polish Ministry of Science
and Higher Education (to A.B.).
References
1. Bebenek, A., Dressman, H. K., Carver, G. T., Ng, S.,
Petrov, V., Yang, G. et al. (2001). Interacting fidelity
defects in the replicative DNA polymerase of bacte-
riophage RB69. J. Biol. Chem. 276, 1038710397.
2. Wang, J., Sattar, A. K., Wang, C. C., Karam, J. D.,
Konigsberg, W. H. & Steitz, T. A. (1997). Crystal
structure of a pol alpha family replication DNA
polymerase from bacteriophage RB69. Cell,89,
10871099.
3. Shamoo, Y. & Steitz, T. A. (1999). Building a replisome
from interacting pieces: sliding clamp complexed to a
peptide from DNA polymerase and a polymerase
editing complex. Cell,99, 155166.
4. Franklin, M. C., Wang, J. & Steitz, T. A. (2001).
Structure of the replicating complex of a pol alpha
family DNA polymerase. Cell,105, 657667.
5. Li, Y., Korolev, S. & Waksman, G. (1998). Crystal
structures of open and closed forms of binary and
ternary complexes of the large fragment of Thermus
aquaticus DNA polymerase I: structural basis for
nucleotide incorporation. EMBO J. 17, 75147525.
6. Kunkel, T. A. & Bebenek, K. (2000). DNA replication
fidelity. Annu. Rev. Biochem. 69, 497529.
792 Suppression of a Mutator DNA Polymerase
7. Kunkel, T. A. (2004). DNA replication fidelity. J. Biol.
Chem. 279, 1689516898.
8. Zhong, X., Pedersen, L. C. & Kunkel, T. A. (2008).
Characterization of a replicative DNA polymerase
mutant with reduced fidelity and increased transle-
sion synthesis capacity. Nucleic Acids Res. 36,
38923904.
9. Yang, G., Franklin, M., Li, J., Lin, T. C. & Konigsberg,
W. (2002). A conserved Tyr residue is required for
sugar selectivity in a Pol alpha DNA polymerase.
Biochemistry,41, 1025610261.
10. Zhang, H., Rhee, C., Bebenek, A., Drake, J. W., Wang,
J. & Konigsberg, W. (2006). The L561A substitution in
the nascent base-pair binding pocket of RB69 DNA
polymerase reduces base discrimination. Biochemistry,
45, 22112220.
11. Echols, H. & Goodman, M. F. (1991). Fidelity
mechanisms in DNA replication. Annu. Rev. Biochem.
60, 477511.
12. Goodman, M. F., Creighton, S., Bloom, L. B. &
Petruska, J. (1993). Biochemical basis of DNA
replication fidelity. Crit. Rev. Biochem. Mol. Biol. 28,
83126.
13. Fidalgo da Silva, E., Mandal, S. S. & Reha-Krantz, L. J.
(2002). Using 2-aminopurine fluorescence to measure
incorporation of incorrect nucleotides by wild type and
mutant bacteriophage T4 DNA polymerases. J. Biol.
Chem. 277, 4064040649.
14. Spacciapoli, P. & Nossal, N. G. (1994). A single
mutation in bacteriophage T4 DNA polymerase
(A737V, tsL141) decreases its processivity as a
polymerase and increases its processivity as a 35
exonuclease. J. Biol. Chem. 269, 438446.
15. Reha-Krantz, L. J. (2010). DNA polymerase proof-
reading: multiple roles maintain genome stability.
Biochim. Biophys. Acta,1804, 10491063.
16. Bebenek, A., Carver, G. T., Dressman, H. K., Kadyrov,
F. A., Haseman, J. K., Petrov, V. et al. (2002). Dissecting
the fidelity of bacteriophage RB69 DNA polymerase:
site-specific modulation of fidelity by polymerase
accessory proteins. Genetics,162, 10031018.
17. Jacewicz, A., Makiela, K., Kierzek, A., Drake, J. W. &
Bebenek, A. (2007). The roles of Tyr391 and Tyr619 in
RB69 DNA polymerase replication fidelity. J. Mol. Biol.
368,1829.
18. Schultz, G. E., Jr, Carver, G. T. & Drake, J. W. (2006). A
role for replication repair in the genesis of templated
mutations. J. Mol. Biol. 358, 963973.
19. Bebenek, K. & Kunkel, T. A. (1995). Analyzing fidelity
of DNA polymerases. Methods Enzymol. 262, 217232.
20. Yang, G., Wang, J. & Konigsberg, W. (2005). Base
selectivity is impaired by mutants that perturb
hydrogen bonding networks in the RB69 DNA
polymerase active site. Biochemistry,44, 33383346.
21. Pettersen, E. F., Goddard, T. D., Huang, C. C., Couch,
G. S., Greenblatt, D. M., Meng, E. C. & Ferrin, T. E.
(2004). UCSF Chimeraa visualization system for
exploratory research and analysis. J. Comput. Chem. 25,
16051612.
22. Zhang, H., Beckman, J., Wang, J. & Konigsberg, W.
(2009). RB69 DNA polymerase mutants with expanded
nascent base-pair-binding pockets are highly efficient
but have reduced base selectivity. Biochemistry,48,
69406950.
23. Ronen, A. & Rahat, A. (1976). Mutagen specificity and
position effects on mutation in T4rII nonsense sites.
Mutat. Res. 34,2134.
24. Muzyczka, N., Poland, R. L. & Bessman, M. J. (1972).
Studies on the biochemical basis of spontaneous
mutation: I. A comparison of the deoxyribonucleic
acid polymerases of mutator, antimutator, and wild
type strains of bacteriophage T4. J. Biol. Chem. 247,
71167122.
25. Clayton, L. K., Goodman, M. F., Branscomb, E. W. &
Galas, D. J. (1979). Error induction and correction by
mutant and wild type T4 DNA polymerases. Kinetic
error discrimination mechanisms. J. Biol. Chem. 254,
19021912.
26. Gauss, P., Doherty, D. H. & Gold, L. (1983). Bacterial
and phage mutations that reveal helix-unwinding
activities required for bacteriophage T4 DNA replica-
tion. Proc. Natl Acad. Sci. USA,80, 16691673.
27. Yang, J., Zhuang, Z., Roccasecca, R. M., Trakselis, M. A.
& Benkovic, S. J. (2004). The dynamic processivity of the
T4 DNA polymerase during replication. Proc. Natl Acad.
Sci. USA,101, 82898294.
28. Fortune, J. M., Stith, C. M., Kissling, G. E., Burgers, P. M.
& Kunkel, T. A. (2006). RPA and PCNA suppress
formation of large deletion errors by yeast DNA
polymerase delta. Nucleic Acids Res. 34, 43354341.
793Suppression of a Mutator DNA Polymerase
... Here the most important mutants are discussed in light of the mechanisms proposed in this study for translocation and active site switching, which involve Tyr567 located at the center of the back wall between Tyr416 and Gly568 in the base pair binding pocket (Supplementary Figure S4d-f). Substitution for this residue with Ala substantially increases the error rate and also leads to other puzzling features (61,62). The available structures of Y567A mutant (e.g. ...
... Therefore the overall structural effect of Y567A is a smaller, rather than larger (63), binding pocket except that the back wall is no longer intact. Compared to the wild-type, Y567A mutant has reduced affinity to bind normal DNA duplex but elevated affinity to bind DNA duplex with mismatched base pair (62). Similar results also show that this substitution increases binding affinity of dGTP by 45-fold opposite a template G compared to the wild-type (64). ...
Article
Full-text available
DNA polymerases in family B are workhorses of DNA replication that carry out the bulk of the job at a high speed with high accuracy. A polymerase in this family relies on a built-in exonuclease for proofreading. It has not been observed at the atomic resolution how the polymerase advances one nucleotide space on the DNA template strand after a correct nucleotide is incorporated, that is, a process known as translocation. It is even more puzzling how translocation is avoided after the primer strand is excised by the exonuclease and returned back to the polymerase active site once an error occurs. The structural events along the bifurcate pathways of translocation and proofreading have been unwittingly captured by hundreds of structures in Protein Data Bank. This study analyzes all available structures of a representative member in family B and reveals the orchestrated event sequence during translocation and proofreading.
... Numerous biochemical and structural studies have focused on the active site of RB69 DNA Pol and provided information about the function of several highly conserved residues [9][10][11][12][13][14]. These include immediate neighbors of the catalytic aspartate triad D411, D621 and D623, engaged directly or indirectly in binding the divalent metal ion and the 3′-end of the primer, as well as several fingers subdomain residues constituting the dNTP binding site and responsible for mismatch discrimination. ...
... The observed antimutator phenotype of D714A polymerase in vitro may result from increased exonuclease activity and/or decreased ability to form binary polymerase-DNA complexes [11,16]. Thus, we decided to test whether the D714A substitution might decrease the DNA binding affinity of the polymerase and/or change its ability to exist in the editing mode required for efficient proofreading. ...
Article
Full-text available
Non-conserved amino acids that are far removed from the active site can sometimes have an unexpected effect on enzyme catalysis. We have investigated the effects of alanine replacement of residues distant from the active site of the replicative RB69 DNA polymerase, and identified a substitution in a weakly conserved palm residue (D714A), that renders the enzyme incapable of sustaining phage replication in vivo. D714, located several angstroms away from the active site, does not contact the DNA or the incoming dNTP, and our apoenzyme and ternary crystal structures of the Pol(D714A) mutant demonstrate that D714A does not affect the overall structure of the protein. The structures reveal a conformational change of several amino acid side chains, which cascade out from the site of the substitution towards the catalytic center, substantially perturbing the geometry of the active site. Consistent with these structural observations, the mutant has a significantly reduced k pol for correct incorporation. We propose that the observed structural changes underlie the severe polymerization defect and thus D714 is a remote, non-catalytic residue that is nevertheless critical for maintaining an optimal active site conformation. This represents a striking example of an action-at-a-distance interaction.
... Pol δ Y708A of motif B affects the same surface, contacting the minor-grove side of the base. The analogous mutant in RB69 Pol (Y567A) is a strong mutator, but this phenotype is suppressed by a second substitution in the dNTP binding site, S565G (Trzemecka et al., 2010;Xia et al., 2011). Structural and biochemical analyses show that G565 in the S565G,Y567A double mutant limits the flexibility of the template base position (Xia et al., 2011) and that, under certain conditions in vitro, the double-mutant polymerase dissociates more rapidly from duplexes with mispaired primer termini (Trzemecka et al., 2010). ...
... The analogous mutant in RB69 Pol (Y567A) is a strong mutator, but this phenotype is suppressed by a second substitution in the dNTP binding site, S565G (Trzemecka et al., 2010;Xia et al., 2011). Structural and biochemical analyses show that G565 in the S565G,Y567A double mutant limits the flexibility of the template base position (Xia et al., 2011) and that, under certain conditions in vitro, the double-mutant polymerase dissociates more rapidly from duplexes with mispaired primer termini (Trzemecka et al., 2010). Thus, changes to the binding pocket may influence not only nucleotide selectivity, but also stability of the replication complex. ...
Article
Evolution balances DNA replication speed and accuracy to optimize replicative fitness and genetic stability. There is no selective pressure to improve DNA replication fidelity beyond the background mutation rate from other sources, such as DNA damage. However, DNA polymerases remain amenable to amino acid substitutions that lower intrinsic error rates. Here, we review these 'antimutagenic' changes in DNA polymerases and discuss what they reveal about mechanisms of replication fidelity. Pioneering studies with bacteriophage T4 DNA polymerase (T4 Pol) established the paradigm that antimutator amino acid substitutions reduce replication errors by increasing proofreading efficiency at the expense of polymerase processivity. The discoveries of antimutator substitutions in proofreading-deficient 'mutator' derivatives of bacterial Pols I and III and yeast Pol δ suggest there must be additional antimutagenic mechanisms. Remarkably, many of the affected amino acid positions from Pol I, Pol III, and Pol δ are similar to the original T4 Pol substitutions. The locations of antimutator substitutions within DNA polymerase structures suggest that they may increase nucleotide selectivity and/or promote dissociation of primer termini from polymerases poised for misincorporation, leading to expulsion of incorrect nucleotides. If misincorporation occurs, enhanced primer dissociation from polymerase domains may improve proofreading in cis by an intrinsic exonuclease or in trans by alternate cellular proofreading activities. Together, these studies reveal that natural selection can readily restore replication error rates to sustainable levels following an adaptive mutator phenotype.
... Despite the information learned from structural and biochemical studies of DNA polymerases, Jan concluded that fidelities of wild-type and mutant RB69 DNA polymerases determined in vitro only partially reflected their fidelities in vivo (Bebenek et al. 2002). For example, biochemical studies of engineered RB69 DNA polymerases showed that increasing the size of the nucleotide-binding pocket reduced nucleotide insertion fidelity, but that was countered by in vivo mutation studies (Trzemecka et al. 2010) that forced reconsideration of the biochemical studies (Xia et al. 2011). Hence, there is a real and continuing need for in vivo replication fidelity experiments. ...
Article
John W. Drake died 02-02-2020, a mathematical palindrome, which he would have enjoyed, given his love of “word play and logic,” as stated in his obituary and echoed by his family, friends, students, and colleagues. Many aspects of Jan’s career have been reviewed previously, including his early years as a Caltech graduate student, and when he was editor-in-chief, with the devoted assistance of his wife Pam, of this journal for 15 impactful years. During his editorship, he raised the profile of GENETICS as the flagship journal of the Genetics Society of America and inspired and contributed to the creation of the Perspectives column, coedited by Jim Crow and William Dove. At the same time, Jan was building from scratch the Laboratory of Molecular Genetics on the newly established Research Triangle Park campus of the National Institute of Environmental Health Science, which he headed for 30 years. This commentary offers a unique perspective on Jan’s legacy; we showcase Jan’s 1969 benchmark discovery of antimutagenic T4 DNA polymerases and the research by three generations (and counting) of scientists whose research stems from that groundbreaking discovery. This is followed by a brief discussion of Jan’s passion: his overriding interest in analyzing mutation rates across species. Several anecdotal stories are included to bring alive one of Jan’s favorite phrases, “to think like a geneticist.” We feature Jan’s genetical approach to mutation studies, along with the biochemistry of DNA polymerase function, our area of expertise. But in the end, we acknowledge, as Jan did, that genetics, also known as in vivo biochemistry, prevails.
... Several studies were conducted to investigate the effects of amino acid exchanges on polymerase fidelity through mismatch extension selectivity of DNA polymerases from sequence family A. In E. coli DNA polymerase I and Thermus aquaticus DNA polymerase it has been shown that within motif C, which plays a key role in primer/template recognition [42], hydrophobic amino acid substitutions result in polymerase variants with increased mismatch extension selectivity [35,43,44]. Amino acid exchanges have been analyzed for the B-family DNA polymerase RB69 and the eukaryotic DNA polymerase δ in comparison to A-family members, showing furthermore that no direct comparisons can be made between analogous amino acids even within the same sequence family: exchanges that led to antimutator variants of Pol δ, turned the corresponding RB69 mutant into a strong mutator [45]. Although B family polymerases are used in many core biotechnological applications, they are still significantly less well understood and especially structural studies of the ternary complex remain elusive [46][47][48]. ...
Article
Full-text available
Fidelity and selectivity of DNA polymerases are critical determinants for the biology of life, as well as important tools for biotechnological applications. DNA polymerases catalyze the formation of DNA strands by adding deoxynucleotides to a primer, which is complementarily bound to a template. To ensure the integrity of the genome, DNA polymerases select the correct nucleotide and further extend the nascent DNA strand. Thus, DNA polymerase fidelity is pivotal for ensuring that cells can replicate their genome with minimal error. DNA polymerases are, however, further optimized for more specific biotechnological or diagnostic applications. Here we report on the semi-rational design of mutant libraries derived by saturation mutagenesis at single sites of a 3’-5’-exonuclease deficient variant of Thermococcus kodakaraensis DNA polymerase (KOD pol) and the discovery for variants with enhanced mismatch extension selectivity by screening. Sites of potential interest for saturation mutagenesis were selected by their proximity to primer or template strands. The resulting libraries were screened via quantitative real-time PCR. We identified three variants with single amino acid exchanges—R501C, R606Q, and R606W—which exhibited increased mismatch extension selectivity. These variants were further characterized towards their potential in mismatch discrimination. Additionally, the identified enzymes were also able to differentiate between cytosine and 5-methylcytosine. Our results demonstrate the potential in characterizing and developing DNA polymerases for specific PCR based applications in DNA biotechnology and diagnostics.
Article
Given the increasing complexity of simulated molecular systems, and the fact that simulation times have now reached milliseconds to seconds, immense amounts of data (in the gigabyte to terabyte range) are produced in current molecular dynamics simulations. Manual analysis of these data is a very time-consuming task, and important events that lead from one intermediate structure to another can become occluded in the noise resulting from random thermal fluctuations. To overcome these problems and facilitate a semi-automated data analysis, we introduce in this work a measure based on C(α) torsion angles: torsion angles formed by four consecutive C(α) atoms. This measure describes changes in the backbones of large systems on a residual length scale (i.e., a small number of residues at a time). Cluster analysis of individual C(α) torsion angles and its fuzzification led to continuous time patches representing (meta)stable conformations and to the identification of events acting as transitions between these conformations. The importance of a change in torsion angle to structural integrity is assessed by comparing this change to the average fluctuations in the same torsion angle over the complete simulation. Using this novel measure in combination with other measures such as the root mean square deviation (RMSD) and time series of distance measures, we performed an in-depth analysis of a simulation of the open form of DNA polymerase I. The times at which major conformational changes occur and the most important parts of the molecule and their interrelations were pinpointed in this analysis. The simultaneous determination of the time points and localizations of major events is a significant advantage of the new bottom-up approach presented here, as compared to many other (top-down) approaches in which only the similarity of the complete structure is analyzed.
Article
Accurate DNA synthesis is vital to maintain genome stability and ensure propagation of species. Synthetic errors have far reaching consequences. Therefore, DNA synthesis is remarkably accurate. The high fidelity is mainly achieved through three steps: ① nucleotide selection, which is based on hydrogen, base pair shape, or some other elements; ② 3'→5' exonuclease proofreading, which removes mis-incorporated nucleotides in cis or trans; ③ repair process, which could correct mismatched nucleotides escaping from proofreading, such as mismatch repair, excission repair, homologous recombination repair, and translesion DNA synthesis. Because all polymerases are suitable targets for anticancer or antiviral drugs, their fidelity is involved in drug resistance and side effects. Understanding the molecular basis of synthesis fidelity is of vital importance. In this review, the fidelity mechanisms of DNA synthesis will be discussed in detail. Furthermore, their application perspective was discussed.
Article
Full-text available
Like most phages with double-stranded DNA, phage T4 exits the infected host cell by a lytic process requiring, at a minimum, an endolysin and a holin. Unlike most phages, T4 can sense superinfection (which signals the depletion of uninfected host cells) and responds by delaying lysis and achieving an order-of-magnitude increase in burst size using a mechanism called lysis inhibition (LIN). T4 r mutants, which are unable to conduct LIN, produce distinctly large, sharp-edged plaques. The discovery of r mutants was key to the foundations of molecular biology, in particular to discovering and characterizing genetic recombination in T4, to redefining the nature of the gene, and to exploring the mutation process at the nucleotide level of resolution. A number of r genes have been described in the past 7 decades with various degrees of clarity. Here we describe an extensive and perhaps saturating search for T4 r genes and relate the corresponding mutational spectra to the often imperfectly known physiologies of the proteins encoded by these genes. Focusing on r genes whose mutant phenotypes are largely independent of the host cell, the genes are rI (which seems to sense superinfection and signal the holin to delay lysis), rIII (of poorly defined function), rIV (same as sp and also of poorly defined function), and rV (same as t, the holin gene). We did not identify any mutations that might correspond to a putative rVI gene, and we did not focus on the famous rII genes because they appear to affect lysis only indirectly.
Article
The contents of the plenary lectures presented at the Plant and Animal Genome (PAG) meeting in January 2011 are summarized in order to provide some insights into the advances in plant, animal and microbe genome studies as they impact on our understanding of complex biological systems. The areas of biology covered include the dynamics of genome change, biological recognition processes and the new processes that underpin investment in science. This overview does not attempt to summarize the diversity of activities that are covered during the PAG through workshops, posters and the suppliers of cutting-edge technologies, but reviews major advances in specific research areas.
Article
We have previously observed that stepwise replacement of amino acid residues in the nascent base-pair binding pocket of RB69 DNA polymerase (RB69pol) with Ala or Gly expanded the space in this pocket, resulting in a progressive increase in misincorporation. However, in vivo results with similar RB69pol nascent base-pair binding pocket mutants showed that mutation rates, as determined by the T4 phage rI forward assay and rII reversion assay, were significantly lower for the RB69pol S565G/Y567A double mutant than for the Y567A single mutant, the opposite of what we would have predicted. To investigate the reasons for this unexpected result, we have determined the pre-steady-state kinetic parameters and crystal structures of relevant ternary complexes. We found that the S565G/Y567A mutant generally had greater base selectivity than the Y567A mutant and that the kinetic parameters for dNMP insertion, excision of the 3'-terminal nucleotide residue, and primer extension beyond a mispair differed not only between these two mutants but also between the two highly mutable sequences in the T4 rI complementary strand. Comparison of the crystal structures of these two mutants with correct and incorrect incoming dNTPs provides insight into the unexpected increase in the fidelity of the S565G/Y567A double mutant. Taken together, the kinetic and structural results provide a basis for integrating and interpreting in vivo and in vitro observations.
Article
Full-text available
The fidelity of DNA synthesis as determined by the misincorporation of the base analogue 2-aminopurine in competition with adenine has been measured as a function of deoxynucleoside triphosphate substrate concentrations using purified mutator (L56), antimutator (L141), and wild type (T4D) T4 DNA polymerases. Although the rates of both incorporation and turnover of aminopurine and adenine decrease as substrate concentrations are decreased, the ratio of turnover/polymerase activity is increased. Thus, the nuclease/polymerase ratio of each of these three DNA polymerases can be controlled. The misincorporation of aminopurine decreases with decreasing substrate concentrations such that all three enzymes approach nearly identical misincorporation frequencies at the lowest substrate concentration. The increased accuracy of DNA synthesis corresponds to conditions producing a high nuclease/polymerase ratio. The misinsertion frequency for aminopurine is independent of substrate concentrations and enzyme phenotype; therefore, the increased accuracy of DNA synthesis with decreasing substrate concentrations is shown to be a result of increased nuclease activity and not increased polymerase or nuclease specificity. The data are analyzed in terms of a kinetic model of DNA polymerase accuracy which proposes that discrimination in nucleotide insertion and removal is based on the free energy difference between matched and mismatched base pairs. A value of 1.1 kcal/mol free energy difference, delta G, between adenine: thymine and aminopurine:thymine base pairs is predicted by model analysis of the cocentration dependence of aminopurine misincorporation and removal frequencies. An independent estimate of this free energy difference based on the 6-fold higher apparent Km of T4 DNA polymerase for aminopurine compared to adenine also gives a value of 1.1 kcal/mol. It is shown that the aminopurine misinsertion frequency for an enzyme having either extremely low 3'-exonuclease activity, Escherichia coli DNA polymerase I, or no measurable exonuclease activity, calf thymus DNA polymerase alpha, is 12 to 15%, which is similar to that for the T4 polymerases and consistent with delta G approximately 1.1 kcal/mol.
Article
Full-text available
Article
Full-text available
The deoxyribonucleic acid polymerases induced by two mutator, two antimutator, and one neutral temperature-sensitive strain of T4D have been partially purified and compared to each other and to the wild type polymerase. Measurements of polymerase, polymerase-associated exonuclease, and deoxynucleoside triphosphate turnover indicate that the mutators and antimutators may be readily distinguished from wild type by the ratios of these activities. The enzymes prepared from mutators have a much lower exonuclease to polymerase ratio than wild type, and the antimutators have a higher ratio of exonuclease to polymerase and a higher rate of nucleotide turnover. Similar measurements with the neutral temperature-sensitive mutant indicate a close resemblance to wild type. On the basis of these experiments we propose that the spontaneous mutation rate is related to the relative rates of insertion and removal of nucleotides during synthesis of deoxyribonucleic acid. Mutations in the structural gene of the polymerase which perturb this exonuclease to polymerase ratio are expressed as mutator or antimutator genes.
Article
DNA polymerase proofreading is a spell-checking activity that enables DNA polymerases to remove newly made nucleotide incorporation errors from the primer terminus before further primer extension and also prevents translesion synthesis. DNA polymerase proofreading improves replication fidelity ∼ 100-fold, which is required by many organisms to prevent unacceptably high, life threatening mutation loads. DNA polymerase proofreading has been studied by geneticists and biochemists for > 35 years. A historical perspective and the basic features of DNA polymerase proofreading are described here, but the goal of this review is to present recent advances in the elucidation of the proofreading pathway and to describe roles of DNA polymerase proofreading beyond mismatch correction that are also important for maintaining genome stability.
Article
We have previously observed that stepwise replacement of amino acid residues in the nascent base-pair binding pocket of RB69 DNA polymerase (RB69pol) with Ala or Gly expanded the space in this pocket, resulting in a progressive increase in misincorporation. However, in vivo results with similar RB69pol nascent base-pair binding pocket mutants showed that mutation rates, as determined by the T4 phage rI forward assay and rII reversion assay, were significantly lower for the RB69pol S565G/Y567A double mutant than for the Y567A single mutant, the opposite of what we would have predicted. To investigate the reasons for this unexpected result, we have determined the pre-steady-state kinetic parameters and crystal structures of relevant ternary complexes. We found that the S565G/Y567A mutant generally had greater base selectivity than the Y567A mutant and that the kinetic parameters for dNMP insertion, excision of the 3'-terminal nucleotide residue, and primer extension beyond a mispair differed not only between these two mutants but also between the two highly mutable sequences in the T4 rI complementary strand. Comparison of the crystal structures of these two mutants with correct and incorrect incoming dNTPs provides insight into the unexpected increase in the fidelity of the S565G/Y567A double mutant. Taken together, the kinetic and structural results provide a basis for integrating and interpreting in vivo and in vitro observations.
Article
We have investigated the effect of systematically enlarging the nascent base-pair-binding pocket (NBP) of a replicative DNA polymerase from bacteriophage RB69 (RB69 pol) on the incorporation efficiency (k(pol)/K(d,app)) for both correct and incorrect dNMPs. Accordingly, we replaced residues L561, Y567, and S565 in the NBP with Ala, Ala, and Gly, respectively. We combined L561A and Y567A to give a double mutant and then introduced the S565G mutation to give a triple mutant. The efficiency of incorrect dNMP insertion increased markedly relative to the wild type with the single mutants and increased further as the number of substitutions in the NBP increased. The difference in incorporation efficiency for mispairs between the mutants and the wild-type RB69 pol was due mainly to k(pol). Unexpectedly, enlarging the NBP had a minimal effect on the incorporation efficiency of correct dNMPs. Our kinetic data suggest that replicative DNA pols exert base discrimination via "negative selection" against mispairs by using residues in the NBP, particularly the three residues analyzed in this study, to allow rapid incorporation of only correct base pairs. This proposal differs from how geometry and "tightness of fit" of the NBP is often invoked to account for rapid incorporation of correct base pairs, namely, that a tighter fit within the NBP leads to an increase in insertion rates [Kool, E. T. (2002) Annu. Rev. Biochem. 71, 191-219]. We related our findings to that of a model translesion DNA pol, Sulfolobus solfataricus Dpo4. We concur with the main conclusion of a previous study [Mizukami, S., et al. (2006) Biochemistry 45, 2772-2778], namely, that lesion bypass pols exhibit low incorporation efficiencies for correct dNMPs (leading to relative promiscuity) not because of a more open NBP but because of a loose fit of substrates bound in the catalytic centers. This is a property not shared by RB69 pol and its mutants.
Article
In 14 sites in the T4rII region, spontaneous and induced interconversions of ochre (UAA) and opal (UGA) alleles, as well as the reversion of the nonsense sites to r+, were studied. The mutagens employed were 2-aminopurine (2AP), 5-bromouracil (5BU), N-methyl-N'-nitro-N-nitrosoguanidine (NTG) and hydroxylamine (HA). With the test system employed, mutagen specificity (i.e., the preferential induction of A: T leads to G: C or G: C leads to A: T mutation at a given site) can be studied. Simultaneously, the response of similar base pairs at various locations in the same or in different nucleotide triplets throughout the rII region, to a given mutagen, can be compared. 2-Aminopurine can induce transitions of both A: T and G: C base pairs at high rates. This mutagen shows no preference for either direction. Furthermore, there is a correlation between the response to 2AP of an A: T pair occupying a given site, and that of a G: C pair occupying the same site. NTG and HA induce G: C leads to A: T transitions almost exclusively. However, there is a correlation between the low rates of A: T leads to G: C transition induced in each of the various sites by these mutagens and those of G: C leads to A: T. 5-Bromouracil induces transitions from G: C to A: T more readily than from A: T to G: C. With 5BU-induced mutation, there is no correlation between the rates of G: C leads to A: T transitions and those of A: T leads to G: C. In UAA sites, all three adenine:thymidine paris respond to 2AP mutagenesis in a similar pattern, In each position in the triplet, response to 2 AP is correlated with that to 5BU. In UGA sites, there are correlations among the spontaneous as well as the 2AP-, HA- and NTG-induced transition rates. 5BU-induced transition rates are usually not correlated with those induced by other mutagens or with the sponatneous ones.
Article
An Escherichia coli strain with a mutation in the optA gene restricts the growth of bacteriophage T4 strains partially defective in gene 43 (DNA polymerase) or missing gene dda (DNA-dependent ATPase). The mutations in the dda gene inactivate a DNA-dependent ATPase that has been shown to have DNA helicase activity in vitro. We show that the restriction of phage growth after infection of the optA bacterium is the result of a block in DNA replication. We infer that the block arises from a defect in DNA unwinding.