You are on page 1of 18

Molecular Microbiology (2014) 91(1), 39–56 ■ doi:10.1111/mmi.

12440
First published online 15 November 2013

Replication of the Escherichia coli chromosome in


RNase HI-deficient cells: multiple initiation regions
and fork dynamics

Nkabuije Z. Maduike,1 Ashley K. Tehranchi,2† Introduction


Jue D. Wang2 and Kenneth N. Kreuzer1*
1
Department of Biochemistry, Duke University Medical The normal cycle of DNA replication in Escherichia coli is
Center, Durham, North Carolina, USA. a highly coordinated process that begins with the DnaA-
2
Department of Bacteriology, University of Wisconsin mediated opening of the DNA duplex at the chromosomal
Madison, Madison, Wisconsin, USA. origin site oriC (for a review, see Mott and Berger, 2007).
Following the assembly of replication machinery at the
site, replication proceeds bidirectionally around the circu-
Summary lar chromosome until the two replication forks meet in the
terminus region, generally within the fork trap between
DNA replication in Escherichia coli is normally initi-
TerA and TerC (Duggin and Bell, 2009).
ated at a single origin, oriC, dependent on initiation
As the key bacterial initiation protein, DnaA is generally
protein DnaA. However, replication can be initiated
essential for cell viability. However, E. coli cells can utilize
elsewhere on the chromosome at multiple ectopic oriK
an alternative mode of DNA replication, constitutive stable
sites. Genetic evidence indicates that initiation from
DNA replication (cSDR), that is independent of DnaA or
oriK depends on RNA-DNA hybrids (R-loops), which
concomitant protein synthesis (Kogoma, 1997). cSDR can
are normally removed by enzymes such as RNase HI to
drive chromosomal DNA replication in knockout mutants of
prevent oriK from misfiring during normal growth.
the rnhA and recG genes, which encode RNase HI and
Initiation from oriK sites occurs in RNase HI-deficient
RecG respectively (Ogawa et al., 1984; Hong et al., 1995).
mutants, and possibly in wild-type cells under certain
These two proteins are involved in the removal of RNA
unusual conditions. Despite previous work, the loca-
from R-loops (RNA-DNA hybrids in otherwise duplex DNA)
tions of oriK and their impact on genome stability
that form on the chromosome. While cSDR is normally
remain unclear. We combined 2D gel electrophoresis
repressed under physiological conditions, it may be acti-
and whole genome approaches to map genome-wide
vated and play an important role even in wild-type cells
oriK locations. The DNA copy number profiles of
under unusual conditions such as entry into stationary
various RNase HI-deficient strains contained multiple
phase or replication after DNA damage (Hong et al., 1996;
peaks, often in consistent locations, identifying candi-
Camps and Loeb, 2005; also see below).
date oriK sites. Removal of RNase HI protein also leads
Based on these and other observations, the late Tokio
to global alterations of replication fork migration
Kogoma and colleagues proposed that replication initiation
patterns, often opposite to normal replication direc-
via cSDR involves specific chromosomal sites called oriK,
tions, and presumably eukaryote-like replication fork
where R-loops form by invasion of duplex DNA with a
merging. Our results have implications for genome
homologous RNA transcript (von Meyenburg et al., 1987;
stability, offering a new understanding of how RNase
Kogoma, 1997). In this model, the invading RNA strand
HI deficiency results in R-loop-mediated transcription-
acts as a primer for replication initiation and the displaced
replication conflict, as well as inappropriate replica-
DNA strand serves as an assembly site for the replicative
tion stalling or blockage at Ter sites outside of the
helicase. While oriK sites have never been precisely iden-
terminus trap region and at ribosomal operons.
tified, Kogoma’s group attempted to map them using an
imprecise chromosomal marker frequency approach that
utilized 21 hybridization probes around the chromosome
(de Massy et al., 1984a; see Masters and Broda, 1971, for
Accepted 23 October, 2013. *For correspondence. E-mail kenneth the first use of copy number analysis in analysing E. coli
.kreuzer@duke.edu; Tel. 919 684 6466; Fax 919 684 6525. †Present
address: Department of Biology, Stanford University, Palo Alto, Cali- replication). Based on the ratio of copy number of these
fornia, USA. probes in exponentially growing versus resting rnhA− ΔoriC

© 2013 John Wiley & Sons Ltd


40 N. Z. Maduike, A. K. Tehranchi, J. D. Wang and K. N. Kreuzer ■

cells, they argued for the existence of four or five oriK mutation and amplification. This study not only highlights
sites in the chromosome, mapped to broad (150–200 kb) the importance of R-loops in genome instability but
regions (de Massy et al., 1984a). Each oriK site was also provides evidence for R-loop-dependent replication
argued to be quite weak in replication activity relative to in wild-type (non-growing) cells. Recent work has also
that of oriC. Interestingly, the terminus region of the chro- begun to reveal a connection between genomic instabil-
mosome showed the highest copy number in rnhA− ΔoriC ity and the formation of RNA-DNA hybrids in cancer cells
cells, which led the authors to conclude that at least two (Potenski and Klein, 2011). Our study of the cSDR
oriK sites exist in this region. Kogoma’s group made great system in E. coli is aimed at improving our understand-
progress towards understanding the protein requirements ing of the mechanisms underlying the propagation of
for cSDR, which include DnaB, DnaC, DnaG, PriA, DNA R-loop structures in the cell, and their effects on chro-
polymerases I and III, and RecA (but no other recombina- mosomal integrity.
tion protein) (see Kogoma, 1997 for review). However, the Our goals in these studies were to better understand
true nature of oriK sites, proposed to be transcription units both the initiation of replication from R-loops at oriK sites
prone to R-loop formation, is still unknown. in the chromosome and the consequences of the aberrant
A better understanding of cSDR may illuminate impor- replication that is thereby induced. The prior studies on
tant aspects of chromosomal replication, its connection to cSDR and Hot sites focused our attention on the terminus
chromosome segregation and the cell cycle, and genome region in our search for oriK sites. First, two of the major
stability. Assuming that multiple weak oriK sites exist oriK sites identified by Kogoma were apparently located
around the chromosome, rnhA− cells, even when they have within this region, although their positions were only
a functional oriC, should have replication forks running in crudely defined. Second, Tus-dependent Hot fragments
the wrong direction relative to normal replication, an unco- were argued to result from the activity of nearby oriK
ordinated replication cycle, and opposing replication forks sites, and Tus-independent Hot fragments in the terminus
meeting in unusual places around the chromosome. region could conceivably be dependent on oriK sites
Notably, E. coli strains with a reduced capacity to remove within (or, again, close to) the Hot fragment (Horiuchi
R-loops are unhealthy, as evidenced by the SOS- et al., 1994). Third, focusing on the terminus region pro-
constitutive phenotypes of rnhA− and recG− mutants, as vides an additional tool that can be used to assess oriK
well as the inviability of rnhA recG double mutants (Hong activity: the accumulation of blocked forks at Ter sites
et al., 1995). Intriguing hotspots for homologous recombi- during replication termination.
nation (Hot sites), mostly in the chromosomal terminus By employing the technique of two-dimensional agarose
region, are also activated in rnhA− E. coli cells (Nishitani gel electrophoresis (Friedman and Brewer, 1995), we were
et al., 1993; Horiuchi et al., 1994). A subset of these Hot able to visualize and quantify stalled replication forks on
sites are adjacent to Ter sites flanking the terminus region, the non-permissive side of Ter sites in RNase HI-deficient
and the induced recombination was shown to be depend- cells. We attempted to localize oriK sites by inserting
ent on the Tus protein (Horiuchi et al., 1994). These results artificial Ter sites at specific locations within the terminus
led to a model in which aberrant replication initiates from region and observing the effects on the proportion of forks
R-loops, the resulting forks stall at the Ter sites for abnor- stalled at the natural TerB site downstream. We then went
mally long times, and breakage at the stalled forks triggers on to use next-generation sequencing (NGS) analyses of
RecA/RecBCD-dependent homologous recombination the genomes of various rnhA− strains to identify regions of
(later results suggested that a second fork from the same origin activity and to analyse the replication dynamics
direction could collide with the blocked fork to generate the when forks are initiated at sites other than oriC. The use of
broken end; Bidnenko et al., 2002). NGS with deep read coverage provides a high level of
In a more general sense, replication disturbances are accuracy in determining copy number around the E. coli
responsible for genome instability caused by R-loops chromosome – for example, pinpointing the location of oriC
(see Discussion) (French, 1992; Rudolph et al., 2009; in dnaA+ cells to within a few kb of its known location. The
2010; Boubakri et al., 2010; Srivatsan et al., 2010; Gan NGS results show a dramatic reversal in apparent replica-
et al., 2011; De Septenville et al., 2012; Wimberly et al., tion directions in an rnhA− dnaA− double mutant, with most
2013). Indeed, R-loops have been linked to DNA forks apparently initiating in the general vicinity of the
rearrangements, double-strand break (DSB) formation, terminus region and replication often terminating in the
and transcriptional elongation defects in species ranging general vicinity of oriC. We also found dramatic disconti-
from yeast to humans (Aguilera and Garcia-Muse, nuities in DNA copy number, implying replication fork stall-
2012). Wimberly et al. (2013) recently provided evidence ing or blockage, at Ter sites outside of the terminus trap
that R-loop-triggered replication in starving phase (wild- region and at ribosomal operons. Finally, the NGS profiles
type) E. coli leads to DNA breaks when the fork encoun- under several conditions strongly suggest that the E. coli
ters nicks in the template, resulting in stress-induced chromosome contains multiple oriK sites, with some

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
Replication in RNase HI-deficient Escherichia coli 41

specific candidate locations evident from small peaks in tion fragment produces a ‘Y arc’ on 2D gels, and stalling of
the profiles. forks at a specific location such as a Ter site results in a
discrete spot on the Y arc, which can subsequently be
quantified (relative to the linear restriction fragment; see
Results Fig. 2A). This approach was previously used to visualize
Blocked forks at Ter sites accumulate at greater levels and quantify Tus-Ter-blocked forks in wild-type E. coli, and
in rnhA− strains as expected, the highest fraction of blocked forks was
found at the innermost TerA and TerC sites (0.19% and
Previous work implicated the terminus region of the E. coli
0.85%, compared with the linear restriction fragment
chromosome as a likely site of oriK elements (see Intro-
respectively; see Duggin and Bell, 2009).
duction). We therefore began our analysis of DNA replica-
As a result of the added oriK replication origin activity in
tion in rnhA knockout E. coli by measuring the blocked
rnhA− cells, a greater fraction of blocked replication forks
forks at Ter sites. The termination events of replication in
are expected to be found at Ter sites in these cells than in
E. coli generally take place in an area known as the repli-
the wild-type, presumably dependent on the location of
cation trap, a 267 kb region delimited by the innermost Ter
oriK sites around the chromosome. For all experiments in
sites, TerA and TerC. The ten known Ter sites, shown in
this study, we grew cells in LB broth because rnhA− mutants
Fig. 1, are sequences to which the termination protein Tus
express higher levels of SOS in LB than in minimal media,
binds in a polar manner, blocking the progression of incom-
consistent with a higher level of cSDR activity (Kogoma
ing replication forks from one direction but not the other.
et al., 1993; also see Usongo et al., 2013). We compared
In the trap region, clockwise-moving forks stall first at
replication fork intermediates at the TerB and TerC sites
TerC, while counter-clockwise forks do so at TerA. In order
of a wild-type and an rnhA− derivative of E. coli strain
to visualize these stalled forks, we made use of two-
MG1655. These two sites were chosen because previous
dimensional (2D) gel electrophoresis, a technique that
studies had suggested that the TerB-TerC region contains
allows digested DNA fragments to be separated by shape
oriK site(s) (see Introduction), which could be particularly
as well as size, thus making it possible to distinguish
evident from an accumulation of blocked forks at TerB (the
between different kinds of replication intermediates. The
upstream TerC should block most forks originating from
collection of Y-shaped replication intermediates resulting
outside this region; see Fig. 1).
from the movement of a replication fork through a restric-
Growing cells were harvested and lysed in low melting
point agarose plugs to extract their large genomic DNAs
without shearing, and restriction digestions were per-
formed in the plugs prior to 2D gel electrophoresis and
Southern blotting with an appropriate probe. Each blot
clearly revealed an accumulation of replication intermedi-
ates as a discrete spot at the appropriate location on the Y
arc, indicative of stalled forks at the Ter site (Fig. 2B). The
fraction of blocked forks was quantified (in repeated experi-
ments) by dividing the signal of the blocked forks by the
sum total of the blocked forks and linear spot (Fig. 2C).
Quantification of blocked forks in the wild-type gave
values modestly higher than those of Duggin and Bell
(2009), perhaps due to differences in growth conditions
(they also used MG1655 as the wild-type, arguing against
a strain difference). We found that the fraction of forks
blocked at the TerB and TerC sites was dramatically higher
in the rnhA− mutant (11-fold and sevenfold respectively;
Fig. 2C). Fully 15% of the TerC restriction fragment signal
consisted of blocked forks, indicating that a clockwise fork
often arrives at TerC well before a counter-clockwise fork –
Fig. 1. Map of Ter sites and rrn operons in the E. coli that is, a strong asymmetry in replication. A sizable fraction
chromosome. Ter sites C, B, F, G, and J are oriented to block
clockwise replication forks from oriC, whereas A, D, E, I, and H (5.7%) of blocked forks was also detected in the TerB
block counter-clockwise forks. The replication fork trap is the restriction fragment; these forks must have originated from
267 kb region between TerC and TerA. The positions of the seven oriK site(s) within the TerB-TerC region and/or forks that
rrn operons are shown (genome coordinate reflects 5′ end of rrs
gene sequence in each). The oriC site is located at position originated upstream of TerC but somehow leaked through
3923.9 kb. All listed genome positions are in kb. it.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
42 N. Z. Maduike, A. K. Tehranchi, J. D. Wang and K. N. Kreuzer ■

Fig. 2. Blocked forks at TerB and TerC.


A. The Y-arc represents the series of Y-shaped replication intermediates that are formed by the movement of a single replication fork through
a segment of digested DNA. The Linear DNA spot (also known as the monomer or 1n spot) is the collection of non-replicating restriction
fragments from the digest. The EcoRI fragments containing TerC and TerB are also diagrammed.
B. Genomic DNA was extracted from wild-type MG1655 and its rnhA− derivative (AQ12251) and digested with EcoRI. Restriction fragments
were run on a 2D gel and probed for the TerB and TerC regions by Southern blotting.
C. The concentrations of blocked replication forks at TerB and TerC in wild-type and rnhA− strains were quantified from 2D gel blots of the
regions (such as the ones shown in Panel B). The concentrations are represented as percentages of the total DNA per blot, and the error
bars represent standard deviations.

In an attempt to distinguish between these two alterna- The insertion of a Ter site at yneK, flxA or ynfL reduced
tives, we inserted additional co-oriented Ter sites within the percentage of blocked forks at TerB by about threefold,
non-essential genes between TerB and TerC in the rnhA− to values around 2%, with somewhat different averages
background. We used a recombineering technique (Baba but overlapping error bars (Fig. 3B). However, insertion of
et al., 2006) to construct four different rnhA− strains con- a Ter site at ydgH caused a more dramatic decrease, down
taining a 23 bp TerA site insertion in place of the non- nearly 10-fold (to 0.59%; Fig. 3B). We also tested double
essential genes yneK, flxA, ynfL, or ydgH (see Fig. 3A). Ter site insertions, one pair at the upstream flxA and yneK
The rationale is that an additional Ter site should sub- sites and the other pair at ydgH and ynfL, closer to TerB
stantially reduce the number of clockwise forks that leak (see Fig. 3A). The percentage of blocked forks at TerB
through TerC and subsequently arrive at TerB, but should averaged 1.4% with the pair of upstream Ter sites, but less
not reduce blocked forks at TerB that originated from oriK than 0.4% with the pair of Ter sites inserted closer to TerB
site(s) between TerB and the inserted site. We confirmed (Fig. 3B). All of these results are consistent with the pres-
that every inserted Ter site did accumulate blocked forks ence of one or more oriK sites in the 8.8 kb ydgH-ynfL
(i.e. was functional; data not shown), but we focused our interval, forks from which can only be blocked by the
quantitative analysis on the forks at the TerB site. leftmost (ydgH) Ter site insertion. However, an important

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
Replication in RNase HI-deficient Escherichia coli 43

Fig. 3. Blocked forks at TerB in strains with


artificial Ter insertions in the TerB-TerC region.
A. Artificial TerA sites were introduced by
recombineering into specific gene loci within
the TerB-C region of rnhA− strains, as
described in Experimental procedures. Two
‘double-insertion’ strains were made by
inserting an additional TerA site into pre-
existing single-insertion strains. All strains were
derivatives of rnhA− strain AQ12251.
B. The concentrations of blocked replication
forks at TerB were quantified from 2D gel blots
of the four single-insertion and two double-
insertion rnhA− strains and compared with that
of the parent strain AQ12251 (with no insert).

caveat remains, since the efficiency of the various inserted Kouzminova and Kuzminov, 2008; Sangurdekar et al.,
Ter sites may not be equal to each other. Further experi- 2010; Skovgaard et al., 2011; Kuong and Kuzminov, 2012;
ments are needed to definitively test for the presence of the Rudolph et al., 2013).
proposed oriK site(s) in this interval and the rest of the The analysis is illustrated first with wild-type MG1655
terminus region. (Fig. 4). Samples were prepared from both growing
and chloramphenicol-treated cells, sequence reads were
aligned to the reference MG1655 genome, and the
Deep sequencing analysis reveals excess replicated
numbers of sequence reads within 100 bp windows were
DNA in the TerA-TerC region of rnhA− cells
tabulated. Prior to tabulation, we eliminated all sequence
We next pursued a more global approach to analyse the bins that included repetitive DNA segments (e.g. rRNA
chromosomal replication pattern of rnhA− strains, using operons, IS elements, REP elements) which cannot be
next generation sequencing (NGS) in an approach similar aligned correctly. The tabulation generates a much more
to those used in recent studies (Muller and Nieduszynski, robust genome copy number profile than that from the
2012; Rudolph et al., 2013). Several prior global studies prior hybridization analysis for oriK sites, which measured
using either microarrays or NGS have documented the only 21 loci (de Massy et al., 1984a).
highest copy numbers for wild-type E. coli in the region The read count profile for chloramphenicol-treated
centred at oriC and the lowest copy numbers in the ter- MG1655 (oriC inactive) was, as expected, fairly uniform
minus region, as expected for bidirectional replication across the genome with no obvious trend (Fig. 4B; note
from oriC (Simmons et al., 2004; Breier et al., 2005; that all NGS data are presented using Log2 scales).

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
44 N. Z. Maduike, A. K. Tehranchi, J. D. Wang and K. N. Kreuzer ■

Fig. 4. NGS profile of MG1655.


Genomic DNA from cultures of
MG1655 without (A) and with (B)
chloramphenicol were prepared for
NGS as described in Experimental
procedures, with repetitive DNA
regions removed from the analysis.
The absolute read counts of 100 bp
bins (Log2) are shown in Panels A and
B (sum of read counts from each of
the 100 positions within the bin); the
relative level (Log2) of the read counts
for replicating cell DNA (sample) over
reference DNA (MG1655+CAM in
Panel B) is plotted in Panel C. The red
line is the LOESS regression curve
(see Experimental procedures).

Spikes and outliers were observed above and below the The overall shape of the adjusted MG1655 genome
mainstream, often in the same location in the two datasets profile was similar to those from the above-cited past
(plus and minus chloramphenicol). These spikes and a studies, with a dominant peak near oriC and the lowest
component of the data scatter might be caused by read copy numbers in the terminus region (Fig. 4C). The peak of
count artefacts (PCR or sequencing variations based on the LOESS curve is at 3916.2 kb, less than 8 kb from the
local sequence characteristics). We therefore calculated actual location of oriC (3923.9 kb), and the lowest value
the ratio of the read counts in the growing cell sample to was at position 1602.5 kb, very close to TerC (1607.2 kb).
those in the chloramphenicol-treated sample for each bin As discussed above, TerC is the Ter site with the highest
(and also divided by the overall ratio of all aligned read frequency of blocked forks in the wild-type, consistent with
counts in the two samples to standardize between most replication termination events occurring in the vicinity
curves). This resulted in a significant smoothing of the of TerC. The profile displayed a heightened slope for about
data, with few spikes remaining (Fig. 4C). We also gen- 300 kb on each side of oriC. This result suggests that
erated a LOESS curve (locally weighted polynomial replication forks travel slower over this interval, or perhaps
regression; red line) which engages the 1000 adjacent that there was some unusual perturbation of replication
data points in each direction to approximate the overall initiation under our conditions; further experiments are
copy number pattern. needed to clarify this point. All the subsequent experiments

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
Replication in RNase HI-deficient Escherichia coli 45

Fig. 5. NGS profiles of rnhA− strains.


Genomic DNA from cultures of rnhA−
cells (strain NZ8; Panels A and B) or
rnhA− rzpR::Ter cells (derivative of NZ8;
Panel C) was prepared for NGS as
described in Experimental procedures.
Log2 of the ratio of read counts (per
100 bp bin) for sample DNA to
reference DNA (chloramphenicol-
treated MG1655) are plotted against
genome position. Panels B and C
present expanded views of the
1200–1800 kb region (the full profile for
the rnhA− rzpR::Ter strain is in Fig. S2).
Locations of the cryptic prophage origin
oriJ and the artificial Ter insertion at
the rzpR locus are also indicated in
Panel C.

below will utilize the same manipulations to generate cor- (counter-clockwise) forks are blocked for prolonged
rected datasets (always relative to the chloramphenicol- periods at TerA, while in other chromosomes, rightward
treated MG1655 data in Fig. 4B). (clockwise) forks are blocked for prolonged periods at
Turning to the analysis of rnhA− E. coli, we found that the TerC, then each of these two classes of chromosomes
oriC peak was poorly defined and that the copy numbers would have two copies of replication fork trap DNA, two
were more evenly distributed across most of the chromo- copies of DNA on the oriK-proximal side of the trap, but only
some (Fig. 5A). Indeed, the main peak is no longer even one copy of the DNA distal to the fork trap region (2:2:1 and
centred at oriC, but rather covers a broad region from about 1:2:2 for leftwards: trap: rightwards). Summing up all chro-
3900 kb to 4300 kb (Fig. 5A; see Discussion). The data mosomes in an equal mix of chromosomes would give a
also revealed the presence of a secondary peak in the peak in the TerA-TerC interval (3:4:3) even if no single
replication fork trap/terminus region, which appears to be chromosomes have replicated only that region. For both
bounded by the TerA site on the left and TerC on the right MG1655 and the rnhA− single mutant, similar (though less
(with a more subtle discontinuity at TerB; see blow-up of well-defined) patterns were also seen using a microarray
this region in Fig. 5B). This strain contains a functional oriC approach to measuring genome copy number profiles
but is also active for oriK sites, which have previously been (Fig. S1).
linked to the terminus region (see above). The peak in the The replication origin of a cryptic prophage within the
fork trap region is consistent with the presence of one or TerA-TerC region, called oriJ, has previously been shown
more oriK sites in this region. Another possible interpreta- to induce autonomous plasmid replication in cells lacking
tion, not mutually exclusive, is that oriK sites exist in both the remainder of the prophage (Diaz and Pritchard, 1978;
halves of the chromosome but fire in an uncoordinated Diaz et al., 1979). We wondered whether oriJ might be
manner. Assume that in some chromosomes, leftward responsible for replication from the terminus region in

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
46 N. Z. Maduike, A. K. Tehranchi, J. D. Wang and K. N. Kreuzer ■

Fig. 6. NGS profile of rnhA− dnaA−


strain. Genomic DNA from a culture of
rnhA− dnaA− cells (strain AQ12257)
was prepared for NGS as described in
Experimental procedures. Log2 of the
ratio of read counts (per 100 bp bin)
for sample DNA to reference DNA
(chloramphenicol-treated MG1655) are
plotted against genome position (Panel
A). Panel B presents an expanded
view of the 1000–1800 kb region.

rnhA− cells, contributing to the strong peak in the terminus and TerG (1682–2375 kb). No discontinuity was evident
region. To test this possibility, we inserted an ectopic Ter at TerF (2315.7 kb), about 60 kb upstream of TerG (with
site in the rzpR locus at position 1421.5 kb, upstream of respect to clockwise-moving forks; Fig. S3). TerF is known
both oriJ and TerA, and in the same orientation as TerA to be a weak Ter site (Sharma and Hill, 1992; Duggin and
(see Fig. 5C). If oriJ is generating much of the excess DNA Bell, 2009), perhaps explaining why forks accumulate at
just to the right of TerA, the peak should be maintained, TerG rather than TerF. A similar, though less well-defined,
while if oriK sites to the right of rzpR are responsible, we profile was revealed by microarray analysis of the rnhA−
should see the overall terminus region peak contract. dnaA− double mutant (Fig. S1C).
Indeed, the inserted Ter site at rzpR now became the An independent way to generate cells that are replicat-
border of the high-copy terminus region, and no disconti- ing only from oriK sites is to treat the rnhA− single mutant
nuity was evident at TerA (Fig. 5C; Fig. S2). We conclude with chloramphenicol, which inhibits replication from oriC
that the peak in the terminus region is not caused by a but not oriK sites (Kogoma, 1978; von Meyenburg et al.,
strong oriK in the TerA-rzpR interval, which includes oriJ, 1987). The profile of DNA from this condition looked very
and this experiment also provides direct confirmation that similar to the profile of the rnhA− dnaA− double mutant
the Ter sites are necessary for the striking discontinuities in above (compare Figs 7A and 6A; also see Fig. S4 for the
the peak shape. profile of the rzpR::Ter insertion strain treated with chlo-
Next, analysis of DNA from an rnhA− dnaA− double ramphenicol). The peaks in the terminus region were very
mutant revealed a striking profile (Fig. 6A). The terminus similar, and both profiles showed the lowest region to the
region provided the peak, this time bordered by disconti- left of oriC (roughly from 3500 kb to 4000 kb). A closer
nuities at TerA and TerB (Figs 6B and S1; see Discus- look at the regions at each edge of this trough is revealing
sion). A novel discontinuity was also evident at TerE, well (Fig. 7B; see also Fig. S5). The rrnD operon (3426.8 kb) is
to the left of the normal terminus region (Fig. 6B). TerE is at or very near a copy number discontinuity on the left,
oriented to trap leftward-moving forks, suggesting that and the rrnE operon (4206.2 kb) is at a discontinuity on
forks trapped at this site were initiated at oriK site(s) the right (Fig. 7B; Fig. S5). There also appear to be less
between TerE and TerD (1081–1279 kb). Also, the region dramatic discontinuities at or near rrnB and rrnA. Several
with the lowest read counts was not far from oriC, with a past studies have shown that replication forks have par-
notable increase in read counts in the region to the right of ticular difficulty travelling through oppositely oriented,
oriC (see below). A striking discontinuity in read counts actively transcribing rrn operons (see Discussion). All of
was also evident very near or at the TerG site (2375 kb; these rrn operons are oriented in the same direction as
Fig. S3). TerG is oriented to trap rightward-moving forks, replication in the wild-type E. coli, but would be oriented in
suggesting that many forks trapped at this site were initi- the opposite direction for replication forks that emanate
ated at oriK site(s) to the left of TerG, likely between TerB from an oriK element in or near the terminus region.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
Replication in RNase HI-deficient Escherichia coli 47

Fig. 7. NGS profile of rnhA− cells


treated with chloramphenicol. A culture
of rnhA− cells (strain NZ8) was treated
for 90 min with chloramphenicol
(150 μg ml−1), and genomic DNA was
prepared for NGS as described in
Experimental procedures. Log2 of the
ratio of read counts (per 100 bp bin)
for sample DNA to reference DNA
(chloramphenicol-treated MG1655) are
plotted against genome position (Panel
A). Panel B presents an expanded
view of the 3200–4400 kb region, with
locations of the rrn operons also
shown. Note that bins containing rrn
operons were removed due to the
repetitive nature of rrn sequences,
resulting in the paucity of data points
at those locations.

Replication fork pausing or blockage at the rrn operons As expected, introduction of the tus mutation into both
therefore provides a likely explanation for the discontinui- rnhA− strains led to a flattening of the terminus region peak
ties highlighted in Figs 7B and S5. We also note that the and the disappearance of the discontinuities at Ter sites
profile in this region is not strictly defined by discontinui- (Fig. 8A and B). The overall profile of the rnhA− tus− strain
ties at rrn operons – this is particularly noticeable around was somewhat flatter than the rnhA− single mutant
rrnD, where gene copy number is changing in a progres- (compare Figs 8A and 5A). This general flattening could
sive way on both flanks. Nearby oriK sites and/or fork result from the global effect mentioned in the paragraph
stalling at sites surrounding rrnD presumably contribute to above.
this aspect of the profiles. The overall shape of the rnhA− dnaA− tus− profile was
similar to that of the double rnhA− dnaA− mutant, absent
the discontinuities and peaks in the regions of Ter sites
Replication profiles of rnhA− strains in the absence of
(compare Figs 8B and 6A). However, the change in profile
Tus-mediated fork blockage
caused by the tus mutation in the terminus region was
Several of the read count discontinuities above occur informative – instead of a single strong peak in the
very near or at Ter sites, and the introduction of a Ter TerA-TerB region, the two regions flanking TerA-TerB had
site at rzpR created a new discontinuity as well. These noticeable peaks (LOESS maxima at 1099.8, 1278.3
results already show that fork blockage by Tus-Ter com- and 1766.5 kb). This result implies that the dramatic peak
plexes plays a major role in the shapes of the profiles. In covering the TerA-TerB region in rnhA− dnaA− tus+ cells is
addition to the local effects on read counts caused by NOT caused primarily by one or more strong oriK site(s)
fork blockage at a particular Ter site, Tus-mediated fork within that region. The region still might contain one or
blockage would likely have more global effects on the more oriK(s), but they are not strong enough to cause this
profile – for example, by extending the expanse of the peak. Reflecting back on the differential hybridization
chromosome replicated by a fork from a given oriK when analysis of de Massy et al. (1984a), these data also
its partner fork is blocked at a Ter site. This effect could negate the evidence for strong oriK site(s) in the TerA-TerB
be complex considering that multiple oriK elements are interval.
likely spread around the chromosome, presumably dif- So what does the data indicate about the location of oriK
fering in strength from one another. To approach this sites in the chromosome? Several features of the two tus−
issue, we constructed rnhA− and rnhA− dnaA− derivatives profiles perhaps provide clues. First, the highest LOESS
with a deletion of the tus gene and analysed their value in the rnhA− tus− profile occurs not at oriC but at
genomic profiles. position 4273 kb, more than 300 kb clockwise from oriC

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
48 N. Z. Maduike, A. K. Tehranchi, J. D. Wang and K. N. Kreuzer ■

Fig. 8. NGS profiles of rnhA− tus−


strains. Genomic DNA from cultures of
rnhA− tus− cells (derivative of strain
AQ12251; Panel A) and rnhA− dnaA−
tus− cells (derivative of strain
AQ12257; Panel B) were prepared for
NGS as described in Experimental
procedures. Log2 of the ratio of
read counts (per 100 bp bin) for
sample DNA to reference DNA
(chloramphenicol-treated MG1655) are
plotted against genome position.

(Fig. 8A). This peak could reflect the presence of oriK profile centred on oriC (with significantly less scatter than
site(s) in this region and/or replication fork blockage at rrn in the microarray data of Kuong & Kuzminov, 2012)
and other operons (see Discussion). Second, in the rnhA− (Fig. 9A). The LOESS-averaged intensity at the oriC
dnaA− tus− triple mutant, where oriC is non-functional, the peak was about fourfold higher than the intensity near the
peak region of the entire genome is broadly in the region bottom of the peak (about 500 kb in either direction) and
of 700 to 1600 kb, with the maximum LOESS value at 8.6-fold higher than the lowest intensity near position
1278 kb (Fig. 8B). This peak region is not very sharply 1591 kb. The highest LOESS value was at position
defined, consistent with the presence of two or more oriK 3925.8 kb, less than 2 kb from the known position of oriC,
sites. As in the tus+ strain, the genome pattern in this attesting to the high precision of the data. The median
mutant is basically opposite the normal, with the oriC distance travelled by forks from oriC should be roughly the
region being the lowest copy number. Third, several dis- distance where the copy number drops in half, which
crete ‘bumps’ on the LOESS curves coincide in the two corresponds to about 250 kb in either direction (261 kb
different tus− profiles; the most obvious is near 1760 kb, clockwise and 245 kb counter-clockwise).
with the exact peak values being within 10 kb of each other A prominent oriC peak is also evident in the profile of
in the two profiles (Fig. 8A and B; also see Discussion). the rnhA− strain treated with HU, as would be expected
(Fig. 9B). Again, the highest LOESS value in this peak
(position 3922.4 kb) was within 2 kb of the known posi-
Hydroxyurea treatment amplifies origin peaks in the
tion of oriC. Also similar to the MG1655 profile, the oriC
genome profile
peak in the rnhA− curve drops by half at about 250 kb in
Hydroxyurea (HU) inhibits the enzyme ribonucleotide either direction from oriC (250 kb clockwise and 241 kb
reductase, leading to depletion of deoxyribonucleotides counter-clockwise).
(dNTPs) in the cell and inhibition of replication. Kuzminov The terminus region of the rnhA− strain treated with HU
and colleagues recently showed a dramatic sharpening of (Fig. 9B) shows a modest peak which is essentially elimi-
the oriC peak with prolonged HU treatment, as measured nated in the rnhA− tus− strain treated with HU (Fig. S6).
by microarray analysis of genome copy number (Kuong The shapes of these two profiles imply that oriK sites
and Kuzminov, 2012). Apparently, oriC initiation can con- distant from oriC are active. First, the small peak in the
tinue in the presence of HU but the forks so initiated terminus region of the rnhA− single mutant seems to
become stalled or move very slowly as they attempt to reflect some oriK activity in the region to the right of TerA,
traverse the chromosome. This seemed like an excellent since the copy number is higher on that side. Second, the
tool to try to define the location of oriK sites more precisely. entire oriC-distal region of this profile is not as depressed
We began by treating wild-type MG1655 with HU for 4 h as in the MG1655 profile, with a ratio between the copy
and analysing the genome profile. In agreement with the number at oriC compared with the lowest point (near
Kuzminov study, we found a very dramatic and sharp position 1228 kb) being only about 4.8-fold (compared

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
Replication in RNase HI-deficient Escherichia coli 49

Fig. 9. NGS profiles of hydroxyurea-


treated cells. Cultures of wild-type cells
(strain MG1655; Panel A), rnhA− cells
(strain AQ12251; Panel B), and rnhA−
dnaA− cells (strain AQ12257; Panel C)
were treated with hydroxyurea
(80 mM) for 4 h, and genomic DNA
was then prepared for NGS as
described in Experimental procedures.
Log2 of the ratio of read counts (per
100 bp bin) for sample DNA to
reference DNA
(chloramphenicol-treated MG1655) are
plotted against genome position. The
scales of the y-axis in this figure are
very different from the previous figures
due to the more extreme variation in
copy number across the genome. Also
note that the DNA from this
experiment was initially sequenced
with 100 base reads, but we truncated
these reads to 50 bases prior to data
analysis so that this data would be
directly comparable to all the other
NGS data in this study. In the
HU-treated profiles, we noted with
interest a set of four 100 bp bins with
very high copy number (this figure and
Fig. S6). These bins are centred at
about position 1019.1 kb at the 5′ end
of the ompA gene. We believe that this
copy number peak is some kind of
PCR artefact, because the excess
reads are all from the same DNA
strand and because no peak was
observed in an analysis of the same
chromosomal DNA sample by a
PacBio direct sequencing protocol that
does not involve PCR (data not
shown). We do not understand why
this peak is seen only in the
HU-treated samples.

with 8.6-fold in MG1655; see above). Third, since the these peaks reflect the activity of specific oriK sites, but
terminus peak is dependent on the presence of Tus, the stronger evidence is needed to confirm this inference. It is
peak presumably consists of Tus-blocked forks that were also likely that both Tus-Ter complexes and rrn operons
triggered by origins that are relatively nearby (since forks have some effects on the overall profile. Based on both
travel only about 250 kb on average in the presence of the rnhA− and rnhA− dnaA− profiles, we infer that multiple
HU). These results support the conclusion that multiple oriK elements are spread around the chromosome and
weak oriK sites exist in the general region of the terminus each initiates replication at a much lower frequency than
(e.g. within a few hundred kb of the replication fork trap). does oriC.
The copy number profile of the rnhA− dnaA− strain
treated with HU argues that oriK activity is occurring
throughout the genome, with the greatest activity in the Discussion
region between about 1200 and 2400 kb (which have
Global replication profiles in rnhA− E. coli
the highest average copy numbers from throughout the
genome; Fig. 9C). Rather than a few discrete peaks In this study, we analysed the replication profiles of wild-
shaped like the oriC peak (only smaller), the profile is type and various rnhA− mutants of E. coli using a precise
quite complex, with multiple peaks and slope changes. NGS approach, backed up by 2D gel and microarray
Small but noticeable peaks do occur with high LOESS approaches. Our studies revealed a dramatic increase in
values at about 1483 kb; 1908 kb; 2592 kb; 2832 kb; the levels of blocked forks at Ter sites in the terminus
3234 kb; and 4397 kb. It seems likely that some or all of region when the rnhA gene is disrupted (Fig. 2), implying

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
50 N. Z. Maduike, A. K. Tehranchi, J. D. Wang and K. N. Kreuzer ■

poorly coordinated replication in the rnhA− mutant (i.e. the based in vivo system showed fork blockage upon encoun-
two oppositely oriented replication forks arrive at these ter with an oppositely oriented transcription complex
Ter sites at very different times). NGS and microarray (Mirkin and Mirkin, 2005). In addition, Pomerantz and
analyses confirmed the unusual replication behaviour of O’Donnell (2010) showed that E. coli replisomes have
rnhA− and particularly rnhA− dnaA− mutants. Wild-type great difficulty replicating through oppositely oriented
cells show little or no excess of DNA in the replication fork transcribing RNA polymerases in vitro. The ability of the
trap interval (TerA-TerC), consistent with well-coordinated bacterial replisome to traverse protein-bound DNA in vivo
replication (Duggin and Bell, 2009; also see Figs 4C and and in vitro was shown to be dependent on the helicases
S1A). However, both of the rnhA− strains show a promi- Rep and UvrD (Guy et al., 2009). Furthermore, Boubakri
nent peak in this region (Figs 5, 6, S1B and C), which et al. (2010) showed that the helicases DinG, Rep and
could be caused by oriK sites in the trap interval and/or UvrD play overlapping roles in allowing replication fork
uncoordinated replication (with some chromosomes progression through oppositely oriented, transcribing rrn
having TerA-trapped clockwise forks and others having operons in E. coli. Finally, replication fork progression in
TerC-trapped counter-clockwise forks). The sharp peak Bacillus subtilis was slowed by about 30% in wild-type
delineated by TerA and TerC, however, was essentially cells when a segment of DNA was replicated in the
abolished in tus− derivatives of the two rnhA− strains unnatural direction, and this slowing was shown to be
(Fig. 8). In the rnhA− dnaA− tus− triple mutant, peaks dependent on oppositely oriented transcription (Wang
appeared in the regions flanking TerA-TerB rather than et al., 2007; also see Srivatsan et al., 2010).
within that region. This argues that blockage of forks from We found the most obvious effect of rrn operons on
outside the TerA-TerB interval plays a major role in the replication profiles in the rnhA− strain when oriC function
formation of the terminus peak in Tus-containing cells. was inhibited by chloramphenicol (Fig. 7B; Fig. S4). The
Additional results from the NGS analysis clearly reveal profiles showed some abrupt but fairly subtle discontinui-
fork blockage by Tus-Ter complexes in rnhA− strains. The ties right at rrnA and rrnE, suggesting direct blockage of the
discontinuity at TerA in the rnhA− single mutant is abol- replisome by the transcribing RNA polymerase. In each of
ished and replaced by a similar discontinuity at an artificial these cases, copy number decreased in the direction
Ter site at the upstream rzpR locus (compare Fig. 5B and expected for oppositely oriented replication forks (since
C). In addition, new and very striking discontinuities are forks are likely travelling in both directions in the rnhA−
evident at TerE, TerB, and TerG in the rnhA− dnaA− double strains). However, there was a more dramatic decrease in
mutant (Fig. 6; Fig. S3). copy number surrounding the rrn operons, apparently in
The tus− strains provide an informative view by elimi- both directions. For example, gradual declines are evident
nating the effects of Ter sites. The rnhA− dnaA− tus− strain for about 100 kb before and 50 kb after rrnD (considering a
shows a peak of copy number at 1278 kb, in the region rightward-moving fork in Fig. 7B). Additional experiments
nearly opposite oriC, and the lowest copy number at are needed to test whether these gradual declines are
3883 kb, only about 40 kb from oriC (Fig. 8B; note that all caused by rrn operon transcription – for example, by a
comparisons of copy number in this Discussion are based mechanism involving transcription-driven supercoiling.
on the LOESS regression values). This result implies that
much of the DNA replication in this strain is opposite in
Where does replication initiate in the chromosomes of
direction to that of wild-type strains. The oriC-competent
rnhA− E. coli?
rnhA− tus− mutant shows the flattest genomic profile of
any of the tested strains, with only a 1.29-fold difference One goal of this study was to localize the origin sequences
between the most and least represented sequences responsible for constitutive stable DNA replication (cSDR)
(Fig. 8A). This is not surprising, since the rnhA− dnaA− tus− in E. coli, the so-called oriK sites. We first focused attention
triple mutant shows nearly the reverse pattern from wild- on the terminus region of the chromosome for reasons
type but the rnhA− tus− double mutant should have an outlined in the Introduction. 2D gel analyses of the rnhA−
active oriC. mutant strain showed elevated levels of blocked replication
The NGS analysis presented here also contributes to forks at both TerB and TerC (Fig. 2). We were particularly
our understanding of the replication-transcription conflict. struck by the level of blocked forks at TerB, since the TerC
The seven ribosomal RNA operons (rrn) of E. coli are site, 75 kb upstream, would be expected to block most
co-oriented with the direction of DNA replication to mini- incoming clockwise forks that originated outside this region
mize head-on collisions between the fork and active tran- (see map in Fig. 1). This suggested that the TerB-TerC
scription complexes in rrn operons (for reviews, see region itself might contain one or more oriK sites. Results
Brewer, 1988; Rudolph et al., 2007; Merrikh et al., 2012). with strains containing one or two additional co-oriented
French (1992) first showed that replication rate is reduced Ter sites inserted within this region were consistent with a
during a head-on collision, and studies with a plasmid- possible oriK site in the 8.8 kb ydgH-ynfL interval, although

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
Replication in RNase HI-deficient Escherichia coli 51

Table 1. Some detectable peaks in rnhA− cells (LOESS curve maxima).

Strain, condition Identified peaks

rnhA 1473.6 4255.8


rnhA dnaA 789.5 1479.3 1867.8 3227.0 4536.6
rnhA, + CAM 847.2 1522.6 1840.2 3234.8 4388.0 4541.2
rnhA tus 1757.6 2667.3 4273.1
rnhA dnaA tus 1766.5
rnhA rzpR::Ter 1508.9 4327.2
rnhA rzpR::Ter, + CAM 846.0 1524.2 1835.5 2598.5 2849.3 3227.7 4341.0
rnhA, + HU 1486.1 1902.2 2621.9 2905.7
rnhA dnaA, + HU 1483.0 1907.5 2591.7 2831.8 3233.7 4396.8
rnhA tus, + HU 1757.5 1893.6 2594.0 2904.1

All numbers in genome coordinates (kb). Bold entries are particularly strong peaks. This list is not exhaustive, but highlights peaks seen in multiple
profiles. The oriC peaks, when present, are not included in this table but are mentioned directly in the text.

this inference depends on the assumption that all inserted 350 kb clockwise from oriC (Table 1, positions near
artificial Ter sites are about equally efficient. We attempted 4300 kb). This result raises the possibility that an oriK site,
to confirm the existence of an oriK site in this region by stronger than oriC, is localized in this region. A different
flanking this region on both sides with oppositely facing Ter interpretation is that this region contains multiple ele-
sites to create a replication ‘bubble trap’. While each of the ments that can block forks travelling in either direction,
two new Ter sites successfully blocked forks as shown by with a resulting accumulation of clockwise forks that origi-
the accumulation of Y-form DNA in 2D gels, we did not nated from oriC, and counter-clockwise forks that origi-
detect any bubble molecules in restriction fragments that nated from oriK sites distal to this region. If the replicated
contained the two Ter sites and intervening DNA (Maduike, regions in the two sets of molecules overlap, a peak could
2012). This result argues against an oriK in this small be generated (much like the peak in the terminus region in
interval, unless some unknown problem prevents forma- Tus-proficient rnhA− cells). This region contains multiple
tion or detection of the expected bubbles. rrn operons, which can impede replication from the
Our global analysis using NGS and microarrays counter-clockwise direction (see above). Furthermore,
changed our perspective on oriK sites in the terminus Gan et al. (2011) recently showed that an R-loop that is
region. The prominent peak in the TerA-TerC replication co-directional with a replication fork in a ColE1-based
trap region of rnhA− strains (Figs 5, 6, 7A, and Figs S1, S2 plasmid can stall replication. Thus, in the rnhA− mutant,
and S4) was essentially abolished in the rnhA− tus− double fork blockage in either direction (in different chromo-
mutant and very much broadened (with local peak posi- somes) could be sufficient to explain the unusual peak.
tions outside of TerA-TerC) in the rnhA− dnaA− tus− triple What have we learned about the location of oriK sites
mutant (Fig. 8). These results argue that the peak within around the chromosome? Their positions should be most
TerA-TerC does not primarily result from DNA molecules evident in DNA from cells in which the powerful oriC is not
that have undergone bidirectional replication from an oriK functional – i.e. rnhA− dnaA− double mutants and rnhA−
within the TerA-TerC region, but rather an accumulation of mutants treated with chloramphenicol. First, the disconti-
some molecules with clockwise forks blocked at TerC and nuity at the TerE site suggested the presence of oriK
other molecules with counter-clockwise forks blocked at site(s) between TerE and TerD (1081–1279 kb; see
TerA. These results likewise argue against the prior iden- above). Second, the discontinuity at TerG suggested the
tification of two oriK sites in the terminus region (de Massy presence of oriK site(s) between TerB and TerG (1682–
et al., 1984a) and the putative oriX site in this same region 2375 kb; see above). Third, the rnhA− dnaA− double
(de Massy et al., 1984b); both studies relied on relative mutant showed a peak in the terminus region bordered by
copy number measurements but with many fewer data TerA and TerB, while the peak in the rnhA− single mutant
points, and both sets of measurements would have been was bordered by TerA and TerC (compare Figs 6B and
similarly impacted by Tus-mediated fork blockage. We 5B). The discontinuity at TerC in the single mutant argues
also obtained evidence that the cryptic phage replication that TerC is efficient at collecting clockwise forks from
origin oriJ does not function as an oriK element in rnhA− oriC, which suggests that the prominent discontinuity at
cells (Fig. 5; also see Maduike, 2012, for negative results TerB in the double mutant results from oriK site(s) in the
with an oriJ bubble trap experiment). TerC-TerB region (1607–1682 kb). 2D gel approaches
A surprising finding from this study is that the overall also provided evidence, albeit equivocal, for oriK site(s) in
copy number of the three rnhA− strains peaked about this interval (Figs 2 and 3). Fourth, the overall peak in the

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
52 N. Z. Maduike, A. K. Tehranchi, J. D. Wang and K. N. Kreuzer ■

NGS genome profile of the rnhA− dnaA− tus− cells was provided supporting evidence that many of these conver-
broadly in the region of about 700–1600 kb (Fig. 8B), sions were dependent on R-loop formation. Such a high
suggesting that oriK site(s) exist in this region. Fifth, the level of R-loop formation in wild-type E. coli seems incon-
profiles of the rnhA− tus− and rnhA− dnaA− tus− cells, which gruous with the normal physiology of wild-type cells
have no fork blockage at Ter sites, show several discrete compared with rnhA− mutants (see below), and this dis-
bumps (maxima in the LOESS curves) that we hypoth- crepancy remains to be explained. If confirmed, these
esize represent oriK sites, the most obvious near position candidate R-loop locations provide another clue to the
1760 kb (Fig. 8). Sixth, modest bumps are also seen in locations of oriK sites (although it is not clear that all
the profiles of rnhA− dnaA− cells and rnhA− cells (with or R-loops would be functional as oriK sites).
without the rzpR::Ter insertion) treated with chlorampheni-
col, including several that show up in all three of these
Pathology resulting from removal of RNase HI
profiles (positions near 850 kb, 1840 kb and 3230 kb;
TerA-TerC region peak not included since it is likely Mutations that inactivate E. coli RNase HI cause serious
created by Tus-mediated fork blockage). Table 1 presents pathology, including an SOS-constitutive phenotype that
a summary of locations that repeatedly appeared as presumably reflects spontaneous DNA damage, and syn-
LOESS curve maxima in the various profiles; these are thetic lethality with recG mutations (see Introduction).
good candidate regions for oriK sites at diverse locations These deleterious effects presumably result from aberrant
around the E. coli chromosome. R-loop formation, but the exact pathways involved in
Following up on the recent microarray study from pathology are still uncertain. Recent studies have high-
Kuzminov’s group (Kuong and Kuzminov, 2012), we con- lighted the deleterious effects of aberrant R-loop forma-
firmed that HU treatment is a powerful tool to use in the tion, including replication fork blockage, mutagenesis,
analysis of genome replication. We found that the average DNA breakage, and DNA rearrangements (Boubakri
distance travelled by replication forks from oriC in wild- et al., 2010; Gan et al., 2011; Helmrich et al., 2011;
type cells during a 4 h HU treatment was only about Stirling et al., 2012; Wimberly et al., 2013) (for a review,
250 kb, and that approximately 8.6-fold more DNA accu- see Aguilera and Garcia-Muse, 2012).
mulated at oriC than at distant regions (Fig. 9A). The Our analysis of chromosomal replication in the absence
rnhA− cells treated with HU showed a prominent peak at of RNase HI illuminates some aspects of the RNase
oriC, and their overall NGS profile was consistent with the HI-deficient physiology. First, the data directly shows that
existence of multiple oriK sites throughout the chromo- rnhA− cells have an altered (and much flatter) profile of
some (Fig. 9B). Finally, perhaps the most informative gene copy number around the chromosome compared
profile was from the rnhA− dnaA− cells treated with HU with the wild-type (compare Fig. 4C with Fig. 5A; Fig. S1).
(Fig. 9C). As with the profiles from these cells without HU Highly expressed genes are concentrated in the early
treatment, the overall peak of the NGS profile was in the replicating region of the E. coli chromosome, and their
terminus region (1483 kb). This is located in the replica- higher copy numbers in rapidly growing cells augment
tion fork trap area (TerA-TerC), and thus fork blockage at their expression (Chandler and Pritchard, 1975; Sousa
Ter sites likely contributes. The profile shows five other et al., 1997; Couturier and Rocha, 2006). The rnhA− cells
maxima (see Table 1) that are consistent with oriK site may suffer physiological consequences due to alterations
activity, two of which are near maxima that were con- in expression caused by aberrant gene ratios. Second,
sistently detected in rnhA− dnaA− cells and rnhA− cells 2D gel analysis and both microarray and NGS profiles
treated with chloramphenicol (maxima at 1907.5 kb and showed that rnhA− cells have a much higher steady-state
3233.7 kb). The LOESS curve in the HU-treated rnhA− level of replication forks blocked at Ter sites. Fork block-
dnaA− cells was very bumpy, suggesting additional weak age at Ter sites can lead to both homologous and illegiti-
oriK sites that do not form maxima and likely also fork mate recombination, both presumably resulting from DNA
blockage at multiple sites such as Ter and rrn operons. breaks (Bierne et al., 1991; 1997; Horiuchi et al., 1994). A
Very recently, Leela et al. (2013) used bisulphite reac- defined pathway of break formation at Tus-blocked forks
tivity in an attempt to map R-loops in the chromosomes of involves the arrival of a second replication fork from the
E. coli MG1655 and a nusG derivative. They extracted same direction (Sharma and Hill, 1995; Bidnenko et al.,
chromosomal DNA from the cells and treated with bisul- 2002). Third, our NGS analysis showed evidence for fork
phite, which deaminates C residues in single-stranded stalling or blockage at rrn operons and in their general
regions, converting C to T (or G to A on the opposite vicinity (see above). Aguilera and Garcia-Muse (2012)
strand after amplification). Surprisingly, they found that discuss in detail the possible pathways of DNA breakage
approximately 7% of sequence reads in the wild-type and genomic instability resulting from fork blockage
showed C to T or G to A conversions from bisulphite during transcription of ribosomal RNA genes. Finally,
treatment, indicative of single-stranded character, and the existence of oriK sites throughout the chromosome

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
Replication in RNase HI-deficient Escherichia coli 53

implies that oppositely oriented forks will meet at many To prepare the tetracycline-resistant Mini-λ plasmid for the
locations, presumably in a fashion that is uncoordinated recombineering experiments, DH10B E. coli cells containing
with the cell division cycle. These aberrant replication the plasmid were grown up to exponential phase in 125 ml of
LB at 30°C, and then incubated at 42°C to induce the excision
termination events may contribute to the pathology of
of the plasmid from the chromosome (see Court et al., 2003).
rnhA− cells (also see Wimberly et al., 2013 for analysis Plasmid DNA was then extracted from the cells using the
of R-loop consequences in non-growing E. coli). With Plasmid Midi kit from Qiagen.
these considerations, it is perhaps surprising that RNase
HI-deficient cells, and particularly derivatives that cannot
utilize oriC, are even viable. Recombineering
While this manuscript was under review, another group Appropriate primers targeting the gene of interest on the
published a study on the aberrant replication that occurs chromosome were first used to generate a PCR product
in recG− E. coli cells using genetic and NGS copy number containing the artificial TerA site plus the kanamycin resist-
approaches (Rudolph et al., 2013). RecG-deficient cells ance gene from the plasmid pKD13 Plus. Primers typically
also undergo cSDR, although the mechanism and loca- consisted of a 50 nt 5′ end corresponding to a DNA
tions of cSDR could differ between recG− and rnhA− cells, sequence flanking the gene of interest in the chromosome
(see Supporting information for Baba et al., 2006), followed
and RecG inactivation is not sufficient to suppress the
by a 20 nt sequence (either 5′-TGTAGGCTGGAGCTG
absence of oriC function (Kogoma, 1997). In general, the CTTCG-3′ for ‘UP’ or 5′-ATTCCGGGGATCCGTCGACC-3′
genome profiles of various recG− mutant cells was remi- for ‘DN’) flanking the segment of the pKD13 Plus plasmid
niscent of the rnhA− profiles reported here, with a dramatic containing the kanamycin resistance gene and the artificial
peak in the terminus region and a reversed pattern (ter- TerA site.
minus to oriC) when oriC function was abolished (Rudolph Cells containing Mini-λ plasmid were grown up to an OD599
et al., 2013). However, Rudolph et al. (2013) concluded of 0.5 with shaking at 32°C. The cell volume was then split
into two, and one half was incubated at 42°C for 15 min to
that much of the replication in recG− cells occurs by an
induce the Mini-λ plasmid, while the other was incubated
aberrant pathway in which colliding replication forks at 32°C to serve as a negative control. The two sets were
trigger new rounds of DNA synthesis, a pathway pre- chilled on ice following their incubations and were used to
vented by the normal function of RecG helicase in wild- prepare electrocompetent cells using ice-cold water or 10%
type cells. Since all of the RNase HI-deficient cells in our glycerol. The recombineering PCR product generated above
study have functional RecG, it seems unlikely that repli- was then electroporated into both electrocompetent cells,
cation fork collisions are triggering replication and con- and kanamycin-resistant transformants were selected on LB
agar plates containing kanamycin. Mock electroporations
tributing to the genome profiles reported here. Further
were also included for both sets of cells. A successful recom-
studies of the unusual mode(s) of replication in RNase HI- bineering experiment produced kanamycin-resistant colonies
and RecG-deficient cells are clearly warranted. from the induced cells but not in the uninduced or mock
controls. Successfully recombineered strains were then
screened by PCR, using primers flanking the original gene in
Experimental procedures the chromosome.
Bacterial strains

All strains used in this study were derivatives of E. coli Preparation of genomic DNA embedded in
MG1655 (F− λ− ilvG− rfb-50 rph-1). Strains AQ12251 agarose plugs
(rnhA339::cat) and AQ12257 (rnhA339::cat dnaA850::Tn10)
were kindly provided by Dr Steven Sandler (University of Escherichia coli cells were grown in LB at 37°C to an OD599
Massachusetts, Amherst); these strains also contain a dele- of 0.4. For each agarose plug to be made, 4 ml of cells
tion of the Eut/CPZ55 prophage. The six strains containing were harvested by centrifugation at 6000 g for 10 min (4°C).
artificial Ter insertions (Fig. 3A) were all derivatives of rnhA− Using the Bio-Rad CHEF Bacterial Genomic DNA Plug Kit
strain AQ12251. (#170–3592), harvested cells were resuspended in Cell
Suspension Buffer (10 mM Tris pH 7.2, 20 mM NaCl,
50 mM EDTA), equilibrated at 50°C, and mixed with a 2%
Plasmids and primers low melting point agarose solution (also equilibrated at
50°C). The cell/agarose mixture was immediately poured
Plasmid pKD13 Plus was made by using the QuikChange II into a plug mold and allowed to solidify at 4°C for 15 min.
Site-Directed Mutagenesis Kit to insert the 23 bp TerA Afterwards, each plug was retrieved and incubated in a
sequence (5′-AATTAGTATGTTGTAACTAAAGT-3′) at posi- 1 mg ml−1 lysozyme solution for 2 h at 37°C, followed by an
tion 1311 on plasmid pKD13 (Baba et al., 2006). PCR primers overnight incubation in a 1 mg ml−1 Proteinase K solution.
used were 5′-GGAACTTCGAACTAATTAGTATGTTGTAAC Plugs were then washed four times in Wash Buffer (20 mM
TAAAGTGCAGGTCGACGGATCCCCGG-3′ and 5′-CCGGG Tris pH 8.0, 50 mM EDTA) and went on to be used for two-
GATCCGTCGACCTGCACTTTAGTTACAACATACTAATTAG dimensional gel electrophoresis. Extra plugs were stored
TTCGAAGTTCC-3′. stably at 4°C for up to 3 months.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
54 N. Z. Maduike, A. K. Tehranchi, J. D. Wang and K. N. Kreuzer ■

Digestion of agarose-embedded genomic DNA Duke Institute for Genome Sciences & Policy (IGSP). The raw
sequence read data (fastq files) have been deposited in the
Agarose plugs containing genomic DNA were immersed in National Center for Biotechnology Information Sequence
1 ml TE buffer (10 mM Tris pH 8.0, 1 mM EDTA) and chilled Read Archive (Accession Number SRP026465). The
on ice for 15–30 min, inverting occasionally. The TE buffer sequencing reads (approximately 13 to 25 million per sample;
was then replaced with 300 μl of the appropriate 1× restriction 17 million average) were imported into Geneious Pro (Biomat-
endonuclease buffer, and chilled on ice again for 15–30 min. ters) and assembled to the reference chromosome MG1655
The restriction endonuclease buffer was then replaced with (GenBank Accession Number 000913.2). The assembly
fresh buffer containing 100 units of the restriction enzyme, process was set to medium sensitivity on Geneious, with the
and the plugs were incubated overnight at 37°C. following parameters: 10–15% gaps allowed per read; word
length of 18; index word length of 13; words repeated more
Two-dimensional agarose gel electrophoresis than 12 times ignored; 20% maximum mismatches per read;
and maximum ambiguity of 4. BAM files of the resulting
Following the overnight digestion, plugs were chilled on ice assembly data were exported to JMP Genomics (SAS), where
and the buffer was carefully removed. Plugs were then equili- read counts were generated for 100 bp bins across the length
brated in 1 ml of 0.5× TBE running buffer (44.5 mM Tris-HCl, of the chromosome by adding the read counts together from
44.5 mM borate, 1 mM disodium EDTA) on ice for 15–30 min. each of the 100 base pair positions (note that every 50-base
Plugs were carefully loaded into the wells of a 12 × 14 cm oligonucleotide read therefore gets counted 50 times, once
0.4% agarose gel and covered with molten 0.5% low melting for each base pair position). The deep sequencing data are
point agarose. The wells were allowed to solidify for 15 min at located in Datasets S1–S3.
4°C, and the gel was run in 0.5× TBE buffer at 15 V for 24 h. By default, the Geneious software maps sequencing reads
Following this first-dimension run, the second dimension gel from repeated regions of the chromosome randomly during
was prepared as outlined by Friedman and Brewer (1995). assembly, and this would produce inaccurate read counts for
Briefly, a slice from each lane was carefully cut out from the these regions. Bins containing repeated regions were there-
first-dimension gel at appropriate positions containing the fore removed from the raw read count data for the samples
targeted restriction fragments and placed at a 90° counter- prior to analysis. First, using the assembly data for our refer-
clockwise orientation in a 20 × 25 cm gel box. Molten 1% ence strain, chloramphenicol-treated MG1655 (MG+CAM),
agarose containing ethidium bromide (0.3 μg ml−1) was then we identified 719 bins (out of 46 397) in the chromosome that
poured over the gel slices and allowed to solidify. This result- contained repeated regions, along with 223 additional bins
ing 2D gel was run at 150 V for 15 h in 0.5× TBE buffer whose read counts were lower than 4 standard deviations
containing ethidium bromide (0.3 μg ml−1). In order to visual- from the mean of the read counts in the set. Second, addi-
ize replication intermediates, a Southern blot was performed tional bins containing annotated RIP and REP elements, rrn
on the 2D gel using radioactive DNA probes specific to the genes, and mobile elements in MG1655 greater than 200 bp
region(s) of interest. Probes were generated by PCR and in length were removed. Third, as most of the strains used
were radiolabelled using the Random Primed DNA Labeling contained an rnhA mutation, and some also contained a
kit from Roche. All blots were visualized using the Storm 820 deletion of the Eut/CPZ55 prophage (strains AQ12251,
Phosphorimager (Molecular Dynamics) and quantified with AQ12257, and their derivatives; see Figs 5C, 6, 8, 9B and C),
ImageQuant software (Molecular Dynamics). the bins for the rnhA gene and the prophage were also
included for removal. Collectively, these identified bins (1281
Deep sequencing analysis in total) were deleted from the raw read count data for each
sample strain analysed, and are listed in Dataset S5.
DNA used for deep sequencing analysis was prepared by Following the removal of the repeat bins above, the raw
harvesting 4 ml of cells grown in LB to an OD599 of 0.5 and read counts for each sample set were divided by the values
extracting genomic DNA using the Qiagen Genomic-tip 100/G from the corresponding bins in the MG+CAM reference strain
kit and Genomic DNA Buffer Set. To prepare chloramphenicol- dataset. This ratio was then multiplied by the ratio of total
treated DNA samples, cells were grown to an OD599 of 0.5 reads in the MG+CAM reference set over total aligned reads
and incubated with chloramphenicol (150 μg ml−1) for 90 min in the relevant sample, to correct for the somewhat different
before harvesting. Hydroxyurea-treated samples were pre- numbers of aligned reads in the various samples. The Log2
pared by growing cells at 30°C in M9 CAA medium to an OD599 values of the corrected sample/reference ratios were then
of 0.2, then adding hydroxyurea to a concentration of 80 mM calculated and plotted against genomic position to generate
and incubating for 4 h. Note that rnhA− dnaA− double mutants, the replication profile charts. A LOESS utility (Peltier Tech;
used in a subset of the genome profile experiments, are http://peltiertech.com/WordPress/loess-utility-for-excel) was
unable to grow on rich media (LB) plates (Torrey et al., 1984). used to generate a smoothed curve on the corrected ratios.
While our double mutant also failed to grow on LB plates, it LOESS calculates a moving weighted regression across the
consistently grew in either minimal media or LB broth when data in Microsoft Excel, and we calculated the Log2 values of
inoculated from colonies on minimal plates. For consistency the LOESS regression values for plotting in the figures. The
with all other experiments, we therefore used rnhA− dnaA− number of points for the moving regression (Npts) in each
double mutants grown in LB broth (inoculated from colonies on case was set to 2000 (corresponding to 200 kb). To compen-
minimal media agar plates). sate for the circular nature of the E. coli chromosome, the first
The genomic DNA was sequenced via Illumina HiSeq™ and last 1000 data points in each dataset were copied to the
(50 bp or 100 bp single reads) at the Sequencing Facility of the end and beginning of the database, respectively, prior to

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
Replication in RNase HI-deficient Escherichia coli 55

running the LOESS calculations. The LOESS regression was Diaz, R., Barnsley, P., and Pritchard, R.H. (1979) Location
then performed on the data as described above, and the and characterisation of a new replication origin in the E. coli
calculated values for the copied data points were removed K12 chromosome. Mol Gen Genet 175: 151–157.
before plotting the LOESS values against genomic position, Duggin, I.G., and Bell, S.D. (2009) Termination structures in
laid over the replication profile charts. the Escherichia coli chromosome replication fork trap. J
Mol Biol 387: 532–539.
Acknowledgements French, S. (1992) Consequences of replication fork move-
ment through transcription units in vivo. Science 258:
This research was supported in part by NIH grants to KNK 1362–1365.
(GM72089) and JDW (GM084003). We are very grateful to Friedman, K.L., and Brewer, B.J. (1995) Analysis of replica-
David MacAlpine and Jason Belsky, who provided invaluable tion intermediates by two-dimensional agarose gel electro-
advice on the NGS analyses. The authors declare that they phoresis. Methods Enzymol 262: 613–627.
have no conflicts of interest with regard to this study. Gan, W., Guan, Z., Liu, J., Gui, T., Shen, K., Manley, J.L., and
Li, X. (2011) R-loop-mediated genomic instability is caused
References by impairment of replication fork progression. Genes Dev
25: 2041–2056.
Aguilera, A., and Garcia-Muse, T. (2012) R loops: from tran- Guy, C.P., Atkinson, J., Gupta, M.K., Mahdi, A.A., Gwynn,
scription byproducts to threats to genome stability. Mol Cell E.J., Rudolph, C.J., et al. (2009) Rep provides a second
46: 115–124. motor at the replisome to promote duplication of protein-
Baba, T., Ara, T., Hasegawa, M., Takai, Y., Okumura, Y., bound DNA. Mol Cell 36: 654–666.
Baba, M., et al. (2006) Construction of Escherichia coli Helmrich, A., Ballarino, M., and Tora, L. (2011) Collisions
K-12 in-frame, single-gene knockout mutants: the Keio col- between replication and transcription complexes cause
lection. Mol Syst Biol 2: 2006 0008. common fragile site instability at the longest human genes.
Bidnenko, V., Ehrlich, S.D., and Michel, B. (2002) Replication Mol Cell 44: 966–977.
fork collapse at replication terminator sequences. EMBO J Hong, X., Cadwell, G.W., and Kogoma, T. (1995) Escherichia
21: 3898–3907. coli RecG and RecA proteins in R-loop formation. EMBO J
Bierne, H., Ehrlich, S.D., and Michel, B. (1991) The replica- 14: 2385–2392.
tion termination signal terB of the Escherichia coli chromo- Hong, X., Cadwell, G.W., and Kogoma, T. (1996) Activation of
some is a deletion hot spot. EMBO J 10: 2699–2705. stable DNA replication in rapidly growing Escherichia coli at
Bierne, H., Ehrlich, S.D., and Michel, B. (1997) Deletions at the time of entry to stationary phase. Mol Microbiol 21:
stalled replication forks occur by two different pathways. 953–961.
EMBO J 16: 3332–3340. Horiuchi, T., Fujimura, Y., Nishitani, H., Kobayashi, T., and
Boubakri, H., de Septenville, A.L., Viguera, E., and Michel, B. Hidaka, M. (1994) The DNA replication fork blocked at the
(2010) The helicases DinG, Rep and UvrD cooperate to Ter site may be an entrance for the RecBCD enzyme into
promote replication across transcription units in vivo. duplex DNA. J Bacteriol 176: 4656–4663.
EMBO J 29: 145–157. Kogoma, T. (1978) A novel Escherichia coli mutant capable of
Breier, A.M., Weier, H.U., and Cozzarelli, N.R. (2005) Inde- DNA replication in the absence of protein synthesis. J Mol
pendence of replisomes in Escherichia coli chromosomal Biol 121: 55–69.
replication. Proc Natl Acad Sci USA 102: 3942–3947. Kogoma, T. (1997) Stable DNA replication: interplay between
Brewer, B.J. (1988) When polymerases collide: replication DNA replication, homologous recombination, and tran-
and the transcriptional organization of the E. coli chromo- scription. Microbiol Mol Biol Rev 61: 212–238.
some. Cell 53: 679–686. Kogoma, T., Hong, X., Cadwell, G.W., Barnard, K.G., and
Camps, M., and Loeb, L.A. (2005) Critical role of R-loops in Asai, T. (1993) Requirement of homologous recombination
processing replication blocks. Front Biosci 10: 689–698. functions for viability of the Escherichia coli cell that lacks
Chandler, M.G., and Pritchard, R.H. (1975) The effect of RNase HI and exonuclease V activities. Biochimie 75:
gene concentration and relative gene dosage on gene 89–99.
output in Escherichia coli. Mol Gen Genet 138: 127–141. Kouzminova, E.A., and Kuzminov, A. (2008) Patterns of chro-
Court, D.L., Swaminathan, S., Yu, D., Wilson, H., Baker, T., mosomal fragmentation due to uracil-DNA incorporation
Bubunenko, M., et al. (2003) Mini-lambda: a tractable reveal a novel mechanism of replication-dependent
system for chromosome and BAC engineering. Gene 315: double-stranded breaks. Mol Microbiol 68: 202–215.
63–69. Kuong, K.J., and Kuzminov, A. (2012) Disintegration of
Couturier, E., and Rocha, E.P. (2006) Replication-associated nascent replication bubbles during thymine starvation trig-
gene dosage effects shape the genomes of fast-growing gers RecA- and RecBCD-dependent replication origin
bacteria but only for transcription and translation genes. destruction. J Biol Chem 287: 23958–23970.
Mol Microbiol 59: 1506–1518. Leela, J.K., Syeda, A.H., Anupama, K., and Gowrishankar, J.
De Septenville, A.L., Duigou, S., Boubakri, H., and Michel, B. (2013) Rho-dependent transcription termination is essen-
(2012) Replication fork reversal after replication- tial to prevent excessive genome-wide R-loops in Escheri-
transcription collision. PLoS Genet 8: e1002622. chia coli. Proc Natl Acad Sci USA 110: 258–263.
Diaz, R., and Pritchard, R.H. (1978) Cloning of replication Maduike, N.Z. (2012) Native Origins for Constitutive Stable
origins from the E. coli K12 chromosome. Nature 275: DNA Replication in Escherichia coli [Dissertation]. Durham,
561–564. NC: Duke University.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56
56 N. Z. Maduike, A. K. Tehranchi, J. D. Wang and K. N. Kreuzer ■

de Massy, B., Fayet, O., and Kogoma, T. (1984a) Multiple and Lloyd, R.G. (2013) Avoiding chromosome pathology
origin usage for DNA replication in sdrA(rnh) mutants of when replication forks collide. Nature 500: 608–611.
Escherichia coli K-12. Initiation in the absence of oriC. J Sangurdekar, D.P., Hamann, B.L., Smirnov, D., Srienc, F.,
Mol Biol 178: 227–236. Hanawalt, P.C., and Khodursky, A.B. (2010) Thymineless
de Massy, B., Patte, J., Louarn, J.M., and Bouche, J.P. death is associated with loss of essential genetic informa-
(1984b) oriX: a new replication origin in E. coli. Cell 36: tion from the replication origin. Mol Microbiol 75: 1455–
221–227. 1467.
Masters, M., and Broda, P. (1971) Evidence for the bidirec- Sharma, B., and Hill, T.M. (1992) TerF, the sixth identified
tional replications of the Escherichia coli chromosome. Nat replication arrest site in Escherichia coli, is located within
New Biol 232: 137–140. the rcsC gene. J Bacteriol 174: 7854–7858.
Merrikh, H., Zhang, Y., Grossman, A.D., and Wang, J.D. Sharma, B., and Hill, T.M. (1995) Insertion of inverted Ter
(2012) Replication-transcription conflicts in bacteria. Nat sites into the terminus region of the Escherichia coli chro-
Rev Microbiol 10: 449–458. mosome delays completion of DNA replication and disrupts
von Meyenburg, K., Boye, E., Skarstad, K., Koppes, L., and the cell cycle. Mol Microbiol 18: 45–61.
Kogoma, T. (1987) Mode of initiation of constitutive stable Simmons, L.A., Breier, A.M., Cozzarelli, N.R., and Kaguni,
DNA replication in RNase H-defective mutants of Escheri- J.M. (2004) Hyperinitiation of DNA replication in Escheri-
chia coli K-12. J Bacteriol 169: 2650–2658. chia coli leads to replication fork collapse and inviability.
Mirkin, E.V., and Mirkin, S.M. (2005) Mechanisms of Mol Microbiol 51: 349–358.
transcription-replication collisions in bacteria. Mol Cell Biol Skovgaard, O., Bak, M., Lobner-Olesen, A., and Tommerup,
25: 888–895. N. (2011) Genome-wide detection of chromosomal rear-
Mott, M.L., and Berger, J.M. (2007) DNA replication initiation: rangements, indels, and mutations in circular chromosomes
mechanisms and regulation in bacteria. Nat Rev Microbiol by short read sequencing. Genome Res 21: 1388–1393.
5: 343–354. Sousa, C., de Lorenzo, V., and Cebolla, A. (1997) Modulation
Muller, C.A., and Nieduszynski, C.A. (2012) Conservation of of gene expression through chromosomal positioning in
replication timing reveals global and local regulation Escherichia coli. Microbiology 143 (Part 6): 2071–2078.
of replication origin activity. Genome Res 22: 1953– Srivatsan, A., Tehranchi, A., MacAlpine, D.M., and Wang,
1962. J.D. (2010) Co-orientation of replication and transcription
Nishitani, H., Hidaka, M., and Horiuchi, T. (1993) Specific preserves genome integrity. PLoS Genet 6: e1000810.
chromosomal sites enhancing homologous recombination Stirling, P.C., Chan, Y.A., Minaker, S.W., Aristizabal, M.J.,
in Escherichia coli mutants defective in RNase H. Mol Gen Barrett, I., Sipahimalani, P., et al. (2012) R-loop-mediated
Genet 240: 307–314. genome instability in mRNA cleavage and polyadenylation
Ogawa, T., Pickett, G.G., Kogoma, T., and Kornberg, A. mutants. Genes Dev 26: 163–175.
(1984) RNase H confers specificity in the dnaA-dependent Torrey, T.A., Atlung, T., and Kogoma, T. (1984) dnaA suppres-
initiation of replication at the unique origin of the Escheri- sor (dasF) mutants of Escherichia coli are stable DNA
chia coli chromosome in vivo and in vitro. Proc Natl Acad replication (sdrA/rnh) mutants. Mol Gen Genet 196: 350–
Sci USA 81: 1040–1044. 355.
Pomerantz, R.T., and O’Donnell, M. (2010) Direct restart of a Usongo, V., Tanguay, C., Nolent, F., Bessong, J.E., and
replication fork stalled by a head-on RNA polymerase. Drolet, M. (2013) Interplay between type 1A topoisomer-
Science 327: 590–592. ases and gyrase in chromosome segregation in Escheri-
Potenski, C.J., and Klein, H.L. (2011) R we there yet? chia coli. J Bacteriol 195: 1758–1768.
R-loop hazards to finishing the journey. Mol Cell 44: 848– Wang, J.D., Berkmen, M.B., and Grossman, A.D. (2007)
850. Genome-wide coorientation of replication and transcription
Rudolph, C.J., Dhillon, P., Moore, T., and Lloyd, R.G. (2007) reduces adverse effects on replication in Bacillus subtilis.
Avoiding and resolving conflicts between DNA replication Proc Natl Acad Sci USA 104: 5608–5613.
and transcription. DNA Repair 6: 981–993. Wimberly, H., Shee, C., Thornton, P.C., Sivaramakrishnan,
Rudolph, C.J., Upton, A.L., and Lloyd, R.G. (2009) Replica- P., Rosenberg, S.M., and Hastings, P.J. (2013) R-loops and
tion fork collisions cause pathological chromosomal ampli- nicks initiate DNA breakage and genome instability in non-
fication in cells lacking RecG DNA translocase. Mol growing Escherichia coli. Nat Commun 4: 2115.
Microbiol 74: 940–955.
Rudolph, C.J., Upton, A.L., Briggs, G.S., and Lloyd, R.G. Supporting information
(2010) Is RecG a general guardian of the bacterial
genome? DNA Repair 9: 210–223. Additional supporting information may be found in the online
Rudolph, C.J., Upton, A.L., Stockum, A., Nieduszynski, C.A., version of this article at the publisher’s web-site.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 39–56

You might also like