You are on page 1of 9

JOURNAL OF BACTERIOLOGY, June 2000, p. 2993–3001 Vol. 182, No.

11
0021-9193/00/$04.00⫹0
Copyright © 2000, American Society for Microbiology. All Rights Reserved.

MINIREVIEW

A Biochemical Mechanism for Nonrandom


Mutations and Evolution
BARBARA E. WRIGHT*
Division of Biological Sciences, The University of Montana,
Missoula, Montana

As this minireview is concerned with the importance of the nario begins with the starvation of a self-replicating unit for its
environment in directing evolution, it is appropriate to remem- precursor, metabolite A, utilized by enzyme 1 encoded by gene
ber that Lamarck was the first to clearly articulate a consistent 1. When metabolite A is depleted, a mutation in a copy of gene
theory of gradual evolution from the simplest of species to the 1 gives rise to gene 2 and allows enzyme 2 to use metabolite B
most complex, culminating in the origin of mankind (71). He by converting it to metabolite A. Then metabolite B is de-
published his remarkable and courageous theory in 1809, the pleted, obtained from metabolite C, and so on, as an increas-
year of Darwin’s birth. Unfortunately, Lamarck’s major con- ingly complex biochemical pathway evolves. In fact, there are

Downloaded from jb.asm.org by on November 23, 2009


tributions have been overshadowed by his views on the inher- examples in which a similar series of events can actually be
itance of acquired characters. In fact, Darwin shared some of observed in the laboratory, for example, involving enzymes
these same views, and even Weismann (106), the father of that are “borrowed” from existing pathways, via regulatory
neo-Darwinism, decided late in his career that directed variation mutations, to establish new pathways (75).
must be invoked to understand some phenomena, as random The starvation conditions that may initiate a series of events
variation and selection alone are not a sufficient explanation such as those described above target the most relevant genes
(71). This minireview will describe mechanisms of mutation for increased rates of transcription, which in turn increase rates
that are not random and can accelerate the process of evolu- of mutation (111). Transcriptional activation can result from
tion in specific directions. The existence of such mechanisms the addition of a substrate or from the removal of a repressor
has been predicted by mathematicians (6) who argue that, if or an end product inhibitor. The latter mechanism, called
every mutation were really random and had to be tested derepression, occurs in response to starvation for an essential
against the environment for selection or rejection, there would substrate or for an end product that represses its own synthesis
not have been enough time to evolve the extremely complex by feedback inhibition. Since evolution usually occurs in re-
biochemical networks and regulatory mechanisms found in sponse to stress (41), transcriptional activation via derepres-
organisms today. Dobzhansky (21) expressed similar views by sion is the main focus of this minireview.
stating “The most serious objection to the modern theory of
evolution is that since mutations occur by ‘chance’ and are
undirected, it is difficult to see how mutation and selection can EVOLUTION OF BIOCHEMICAL PATHWAYS
add up to the formation of such beautifully balanced organs as, A number of events initiated by carbon source starvation can
for example, the human eye.” facilitate the evolution of a new catabolic pathway. Under
The most primitive kinds of cells, called progenotes by these circumstances, cells with gene duplication and higher
Woese (108), were undoubtedly very simple biochemically with enzyme levels have a selective advantage (87, 95). In some
only a few central anabolic and catabolic pathways. Wächter- systems, duplicated segments are specifically subject to higher
häuser (103) theorizes that the earliest metabolic pathway was mutation rates (93), providing ideal and expendable material
a reductive citric acid cycle by which carbon fixation occurred for mutations representing minor modifications of existing
(64). At that point in time, some four billion years ago, how did genes (58). These new genes can encode modified enzymes
the additional, more complex metabolic pathways found in catalyzing reactions closely related and/or complementary to
even the simplest prokaryotes evolve? For that matter, how are those in existence (56). An additional consequence of starva-
they evolving today? As pointed out by Oparin (79), it is in- tion is the removal of feedback controls, resulting in the dere-
conceivable that a self-reproducing unit as complicated as a pression of genes previously inhibited by the now absent me-
nucleoprotein could suddenly arise by chance; a period of tabolite. Increased rates of mutation in these derepressed
evolution through the natural selection of organic substances genes increase the probability of creating a new gene-enzyme
of ever-increasing degrees of complexity must intervene. system. A number of examples exist in which derepression of a
Horowitz (40) suggests a plausible scheme by which biosyn- gene has enabled an enzyme to use a new substrate. For ex-
thetic pathways can evolve from the successive depletion and ample, altros-galactoside can be used by ␤-galactosidase after
interconversion of related metabolites in a primitive environ- it is derepressed (53); other examples are ␤-glycerolphosphate
ment, as the rich supply of organic molecules is consumed by a via alkaline phosphatase (100), putrescine via diamine-␣-keto-
burgeoning population of heterotrophs. Thus, a possible sce- glutarate transaminase (44), and D-mannitol via D-arabitol de-
hydrogenase (55).
An excellent example of the evolution of biochemical path-
* Mailing address: Division of Biological Sciences, The University of ways involves the modification of two genes to serve the new
Montana, Missoula, MT 59812. Phone: (406) 243-6676. Fax: (406) demands imposed by carbon source starvation (56, 112). Rib-
243-4184. E-mail: bewright@selway.umt.edu. itol dehydrogenase, which is induced by ribitol in wild-type

2993
2994 MINIREVIEW J. BACTERIOL.

TABLE 1. Evolution of a catabolic pathwaya ities by coordinating these activities with the presence or ab-
sence of nutrients in the environment. High mutation rates in
Generation Km (mM) [14C]xylitol uptake
Strain derepressed genes prepare cells to respond rapidly to new
time (h) (cpm)
Xylitol Ribitol challenges should the stress become more severe. As will be-
b come apparent, genetic derepression may be the only mecha-
X 4.1 0 3 104
X1c 4.1 290 3 104 nism by which particular environmental conditions of stress
X2 1.7 120 2 185 target specific regions of the genome for higher mutation rates
X3 0.9 130 2 685 (hypermutation). Although this direct avenue for increasing
a
variability is probably not available to multicellular organisms
Data taken from reference 112. in which germ cells and somatic cells are separated, the dere-
b
Wild-type strain with inducible ribitol dehydrogenase.
c
Mutant strain with constitutive ribitol dehydrogenase. pression of biosynthetic pathways is essential to increased lon-
gevity in mammals subjected to caloric restriction (54), and
amino acid limitation in rats can also induce gene expression
(9).
Enterobacter arogenes (strain X in Table 1), is unable to use
xylitol. Starvation for ribitol in the presence of xylitol results in
a mutation to strain X1, in which ribitol dehydrogenase is MECHANISMS OF MUTATIONS VERSUS
constitutive and able to use xylitol, which is a poor substrate for MECHANISMS OF EVOLUTION
the enzyme but not its inducer. By repeated growth cycling on In a scientific context, the word spontaneous is meaningless.
xylitol, derivative mutants X2 and X3 are obtained with lower Every event is preceded by, and dependent upon, innumerable
Km for xylitol. The enhanced uptake of labeled xylitol in the known and unknown prior events and circumstances. There-
final mutant, X3, is due to the acquisition of a constitutively fore, the work background will be used when referring to the
expressed active transport system for xylitol, originating from many environmental conditions, cellular events, and repair

Downloaded from jb.asm.org by on November 23, 2009


the modification of an inducible transport system for D-arabi- processes that affect mutation rates in nature. Although back-
tol. Thus, two preexistent gene-enzyme systems evolve to ini- ground mutations, such as deamination, alkylation, and depuri-
tiate a new catabolic pathway in response to the stress of nation, occur with low frequencies, they have characteristic,
imminent starvation. finite activation energies under physiological conditions (31,
In the evolution of a growth rate-limiting amino acid bio- 32, 48, 60), and the molecular mechanisms by which they occur
synthetic pathway, starvation, derepression, and higher muta- are well established. For example, protonation of the N-3 of
tion rates can result in a lower Km for the rate-controlling cytosine results in its deamination to uracil. This process oc-
endogenous precursor of the pathway or in the ability to use a curs 140 times more frequently in single-stranded DNA
new and more plentiful precursor for the synthesis of that (ssDNA) than in double-stranded DNA (dsDNA) (in which
amino acid. N-3 is hydrogen bonded to N-1 of guanine), in mismatched
base pairs, and in AT-rich regions that “breathe.” These back-
SPECIFICITY OF STARVATION-INDUCED ground mutations and frameshift-induced deletions and addi-
DEREPRESSION tions are DNA sequence directed in that they occur most
frequently in ssDNA and in unpaired, mispaired, and methyl-
Starvation for any essential nutrient activates systems that ated bases (15, 26, 28, 31, 32, 60). Such vulnerable bases can
protect the vulnerable cells from environmental damage (37, arise as a consequence of slippage in tandem repeats or as a
72, 91). In addition, elaborate and specific feedback mecha- result of stem-loop DNA secondary structures that arise from
nisms are deployed that counteract the particular crisis created sequences containing intrastrand inverted complements. The
by the absent nutrient (9, 13, 39, 54, 66, 77). For example, loops in these structures consist of unpaired bases particularly
inorganic phosphate (Pi) starvation derepresses the pho regu- vulnerable to mutations (other more complex structures con-
lon, including a new high-affinity Pi transport system able to taining vulnerable bases are also important as precursors of
cope with lower phosphate levels, and a hydrolase able to mutations but will not be discussed in this minireview).
obtain Pi from new sources (67). Nitrogen starvation dere- In an evolutionary context, we are not concerned with the
presses the ntr regulon, including glutamine synthetase, which above molecular mechanisms by which individual types of mu-
has a higher affinity for NH4⫹ than the constitutive glutamate tation occur but with another kind of mechanism. Evolution
dehydrogenase (65). Starvation for leucine specifically targets depends upon events that enhance mutation rates, thus in-
derepression of the genes in the leu operon (111). Regulation creasing the supply of variants from which the fittest are se-
of amino acid biosynthetic operons by attenuation is exquis- lected. Therefore, the word mechanism in the present context
itely sensitive (over ranges of 1,000-fold) to the need for, and will refer to the circumstances affecting mutation rates. That
the supply of, all the amino acids (18, 49, 97). Attenuation any DNA-destabilizing event will increase mutation rates is
regulation is an impressive example of the remarkable mech-
anisms that have evolved to ensure the conservation of pre-
cious reserves and the derepression and activation of only TABLE 2. Leader sequences of attenuation-regulated operons
those systems essential for survival under particular conditions Operon Amino acid sequence of codons in leader RNAa
of starvation. As seen in Table 2, each amino acid operon
his..................Met-Thr-Arg-Val-Gln-Phe-Lys-His-His-His-His-His-His-
encodes in its leader sequence a series of codons for the amino His-Pro-Asp-
acid product of that operon. This sets the stage for a very phe.................Met-Lys-His-Ile-Pro-Phe-Phe-Phe-Ala-Phe-Phe-Phe-Thr-
complex and specific mechanism to monitor the precise Phe-Pro-
amount of amino acid required relative to the amount avail- thr..................Met-Lys-Arg-Ile-Ser-Thr-Thr-Ile-Thr-Thr-Thr-Ile-Thr-
able (49, 107). If the Leu codons are replaced with Thr codons, Ile-Thr-Thr-Gly-
regulation of the leu operon by leucine is abolished (12). leu..................Met-Ser-His-Ile-Val-Arg-Phe-Thr-Gly-Leu-Leu-Leu-Leu-
Asn-Ala-Phe-
Presumably, feedback mechanisms existing today evolved in
the past to prevent unnecessary and wasteful metabolic activ- a
Operon-specific sequences shown in bold type.
VOL. 182, 2000 MINIREVIEW 2995

axiomatic. In growing cells, such events include polymerase tator phenotypes (35, 98), horizontal gene transfer by trans-
and proofreading errors during DNA replication, as well as duction with a viral particle, and mobile genetic elements that
recombination, transcription, and repair. DNA transcription increase mutation rates by inserting at particular regions or at
and repressor binding affect the rate of deletions in Escherichia target sequences within the genome (73, 76). Such mechanisms
coli plasmids (101). The availability of ssDNA (leading- and are undirected because, for example, a mismatch repair defi-
lagging-strand DNA templates) facilitates the slippage of tan- ciency will result in failure to repair a particular kind of lesion
dem repeats and the formation of stem-loop structures (89). In regardless of whether or not it confers a selective advantage
nongrowing cells, however, the DNA-destabilizing events of upon its host.
replication are probably not primary causes of mutations. In higher organisms, environmental conditions of stress do
Within an hour following starvation, bacterial cells undergo not have direct access to the cells involved in reproduction, and
major metabolic transitions (the stringent response [13]) in different mechanisms resulting in hypervariation have evolved.
which genes required for cell division are repressed while a For example, localized DNA rearrangements and shuffling
number of other genes (depending upon the starvation regi- produce extensive beneficial variation (82, 96), and hypervari-
men) are derepressed. During this transition from exponential able sequences provide continual changes in the composition
growth to stationary phase, events related to gene activation of venoms produced by snakes (29) or snails (78) to overcome
parallel a sharp increase in supercoiling, suggesting that tran- resistance developed by their predators or prey. These mech-
scriptional activation may drive supercoiling and the resulting anisms are also random. The threat of predators does not
DNA secondary structures that are precursors of mutations result in hypermutation; there is no evidence that the circum-
(discussed below). Among the known DNA-destabilizing stances selecting such hypermutable genes bear any metabolic
events, only transcription can be selectively activated (7), ei- relationship to the mechanisms by which they originally arose.
ther by induction or derepression. Derepression of the leu A gene may be hypermutable because it contains a hot spot
operon in E. coli is specifically correlated with an increased due to a particular DNA sequence, and if a high mutation rate
rate of leuB mRNA turnover (62; J. M. Reimers, A. Longacre, is advantageous to its host, that gene will be selected during

Downloaded from jb.asm.org by on November 23, 2009


and B. E. Wright, Conf. DNA Repair Mutag., abstr. B29, p. 80, evolution. However, its hypermutability per se is undirected,
1999) and an increased reversion rate of the leuB mutant gene; since it is unrelated to those selective conditions and to the
this mutation is located at the site of a predicted stem-loop function of the gene. These random mechanisms resulting in
structure (111). hypermutation are in essence serendipitous relationships; in
contrast, hypermutation resulting from derepression is local-
RANDOM VERSUS NONRANDOM ized as a direct consequence of a specific response to environ-
HYPERMUTATION mental challenge.

As discussed above, background mutations are sequence TWO MECHANISMS BY WHICH TRANSCRIPTION CAN
directed and not random in the sense that they occur in bases INCREASE MUTATION RATES
made vulnerable by virtue of their particular location within
specific DNA sequences, such as tandem repeats, or the un- Transcription exposes ssDNA. The most common base sub-
paired and mispaired bases of stem-loop structures. Dobzhan- stitution events in the spectra of background mutations in
sky’s statement (22) enlarges upon this point: “The structure of E. coli and mammalian cells are G 䡠 C-to-A 䡠 T transitions. Fix
a gene is a distillate of its history, and the mutations that may and Glickman (28) observe that 77% of these mutations orig-
occur in a gene are determined by the succession of environ- inate on the nontranscribed strand in E. coli mutants unable to
ments in which that gene and its ancestors existed since the repair deaminated cytosines. This suggests that the unpro-
beginnings of life. The environment prevailing at the time tected single strand in the transcription “bubble” is significant-
mutation takes place is only a component of the environmental ly more vulnerable to mutations than the transcribed strand,
complex that determines the mutation.” The definitions of which is protected as a DNA-RNA hybrid (Fig. 1A). The fre-
directed and random that are appropriate in the above context quency of UV-induced lesions in the lacI gene is also higher in
are neither relevant nor useful, however, when discussing the nontranscribed strand than in the transcribed strand (46).
mechanisms of evolution. By the neo-Darwinian definition, a In fact, cytosines deaminate to uracils in ssDNA at more than
mutation is random if it is unrelated to the metabolic function 100 times the rate in dsDNA (31, 32, 60). The relative muta-
of the gene and if it occurs at a rate that is undirected by bility of the nontranscribed strand is also seen in a plasmid
specific selective conditions of the environment. For example, system in which a fourfold increase in the frequency of tran-
mutagenic DNA-destabilizing events associated with cell divi- sitions occurs selectively in the nontranscribed strand when
sion are random, as they are dependent upon growth rate and transcription is induced (4). Transcription may therefore be a
selective conditions of the environment only insofar as those prerequisite for many C-to-T transition mutations, since other
conditions affect the rate of cell division. However, the focus of mechanisms resulting in the transient generation of single-
this minireview concerns the consequences of environmental stranded sequences, such as replication or breathing (102) do
stress on evolution. What are the DNA-destabilizing processes not lead to asymmetry in the two strands. Apparently, the ob-
operative in stressed, nongrowing organisms forced to mutate served strand bias cannot be explained by transcription-cou-
before they can continue to multiply? Mechanisms must have pled repair (36), since base mismatches are poor substrates for
evolved in starving cells to stimulate metabolic changes and this kind of repair, and the same strand bias is observed when
mutations that facilitate adaptation to new circumstances. the host is deficient in repairing U 䡠 G and T 䡠 G mismatches
With the above neo-Darwinian definition of random in (4). Thus, transcription may be implicated as a major cause of
mind, an impressive array of circumstances that enhance back- background transition mutations in nature.
ground mutation rates in response to environmental stress may Transcriptional activation as a mechanism for increasing
be examined with respect to whether or not they are random mutation rates was first proposed in 1971, by Brock (8) and
(undirected). Examples of conditions that result in undirected, Herman and Dworkin (38). Their work demonstrates that
genomewide hypermutation include those caused by UV irra- recA-independent lac reversion rates of frameshift and point
diation, reactive oxygen species, mismatch repair-deficient mu- mutations are higher when transcription is induced by isopro-
2996 MINIREVIEW J. BACTERIOL.

activation triggered by gene derepression, not induction. In


this system, mutations arise during the transition between
growth and stationary phase and they are recA independent,
similar to the lac reversions mentioned above. This distin-
guishes them from prolonged stress-induced adaptive muta-
tions (11) and from DNA damage-induced SOS mutagenesis
(104), both of which require recA (and will not be discussed in
this minireview). It is noteworthy that the experiments de-
scribed above on the effects of artificially induced transcription
on mutation rates in growing cells are all examples of specifi-
cally directed mutations. However, none of the researchers
come to that conclusion or challenge the assumptions and
implications inherent in the experiments of Luria and Del-
bruck (63), which reinforce neo-Darwinism. This situation may
be due to the dominance of current dogma and to the assump-
tion that mechanisms operative during growth cannot also be
critical during evolution under conditions of environmental
stress. In fact, the limited evidence now available suggests that
only growing cells, or cells in transition between growth and
stationary phase, have the metabolic potential required for
specific, transcription-induced mutations in response to envi-
ronmental challenge. Thus, IPTG induction enhances lac re-
version rates in growing cells (38) but not in cells subjected to

Downloaded from jb.asm.org by on November 23, 2009


prolonged stress (17). Transcriptional activation is the mech-
anism for enhancing mutation rates both in the artificially
induced systems and in stringent response mutations (90, 111;
J. M. Reimers, A. Longacre, and B. E. Wright, Conf. DNA
Repair Mutag., 1999). However, only the latter are relevant to
evolution, since they occur naturally as a result of starvation-
induced derepression. Mutations that most benefit organisms
and accelerate evolution may occur as an immediate response
to imminent starvation, when cells still have the metabolic
resources to respond specifically to the particular conditions of
stress at hand.
Transcription drives localized supercoiling. Chromosomal
DNA from bacterial cells is negatively supercoiled. The level of
global negative supercoiling in E. coli cells is maintained within
a physiologically acceptable range by two opposing enzyme
activities: DNA gyrase, which introduces negative supercoils,
and topoisomerase I, which relaxes them. Investigations with
plasmids grown in E. coli (59, 81) demonstrate the presence of
stem-loop structures in naturally occurring supercoiled circular
DNA molecules (Fig. 1B). Analyses with single strand-specific
nuclease show that DNA molecules with high superhelical
densities are selectively cleaved, in contrast to their linearized
counterparts with which they are in dynamic equilibrium in
vivo. The sequence surrounding the area of cleavage reveals
inverted complementary sequences that hydrogen bond to be-
come the stem separated by noncomplementary bases that
become the single-stranded loop and the substrate for nuclease
cleavage. Such complex structures form preferentially in easily
denatured AT-rich stretches of DNA and occur about 10,000
times more frequently than expected by chance (30, 59), sug-
gesting their selection during evolution. Data indicate that
stem-loop-based recombination may have evolved in the early
“RNA world” (94) and that the potential to generate stem-
FIG. 1. (A) Exposure of the nontranscribed strand during transcription; (B)
effect of transcription on supercoiling; (C) a typical stem-loop structure contain-
loops was later conserved, for example, in hypervariable snake
ing unpaired and mispaired bases; (D) mutation 1, a C-to-T transition in the loop. venom genes under strong selection to keep one step ahead of
both predators and prey (29).
A number of variables, such as temperature, anaerobiosis,
pyl-␤-D-thiogalactopyranoside (IPTG), and that the effect is osmolarity, and nutritional shifts, affect DNA supercoiling (1,
specific. More recently, specifically induced, transcription-en- 19, 47, 85, 86). Some environmental perturbations affect plas-
hanced mutations have also been shown for a lys frameshift mid systems and chromosomes in a similar manner, while some
mutation in Saccharomyces cerevisiae (16, 74). Starvation-in- apparently do not (25). Transcription both responds to and
duced stringent response mutations in E. coli (62, 109–111) promotes changes in supercoiling. The optimal level of super-
and Bacillus subtilis (90) occur as a result of transcriptional coiling for gene expression varies for different genes, and su-
VOL. 182, 2000 MINIREVIEW 2997

sion in response to starvation for any essential nutrient. The


two “alarmones,” cAMP and ppGpp, activate transcription in
derepressed genes in a number of systems. Both are required
for the synthesis of enzymes that catabolize alternative carbon
sources, such as ␤-galactosidase (83, 84, 91). The alarmone
ppGpp activates the synthesis of ␴S (33), which in turn governs
the expression of a number of stationary-phase genes involved
in the starvation-mediated resistance to osmotic, oxidative, and
heat damage (37, 72). Under circumstances in which a group of
related genes become activated, such as those dependent upon
␴S, the topological changes in DNA could provide a mechanism
by which transcriptional activation in one gene may influence
adjacent genes (47, 85). Changes in stem-loop formation and
superhelicity similar to that caused by glucose starvation (Fig.
2) are also observed immediately following amino acid starva-
tion or the inhibition of protein synthesis (19). Within 30 min
following treatment with chloramphenicol or valine (which cre-
ates an isoleucine deficit and ppGpp accumulation in E. coli),
FIG. 2. Depletion of limiting levels of glucose by E. coli immediately triggers
several metabolic events. Depletion of a limiting amino acid has similar conse-
stem-loop formation is evident. Isoleucine starvation results in
quences. Curve A represents growth (optical density). Curve B represents the the formation of stem-loops in a much smaller number of
concentration of cAMP, ppGpp, ␴S, ␤-galactosidase, 30 other proteins, and DNA molecules than are affected by chloramphenicol inhibi-
supercoils. tion, consistent with the selective derepression of relatively few
genes in the absence of a single amino acid.

Downloaded from jb.asm.org by on November 23, 2009


In response to starvation for any essential metabolite, the
percoiling-induced conformational changes may be required immediate problem is addressed specifically (e.g., derepression
for structural changes in regulatory complexes and for recog- of a higher-affinity transport system for that metabolite), cou-
nition by RNA polymerase (RNAP) (85). Transcription has a pled with a general increase in stress resistance. Starvation
profound effect on supercoiling, because RNAP distorts and results in derepression, and transcription drives localized su-
destabilizes dsDNA. As indicated in the twin-domain model percoiling; the formation of stem-loop structures at regions of
(Fig. 1B) of Liu and Wang (61), negative supercoiling is gen- high superhelicity results in localized hypermutation (Fig. 1).
erated behind, and positive supercoiling in front of, the ad- Although energetic considerations do not favor the creation of
vancing RNAP transcription complex. Many investigations complex structures in metabolically inactive dsDNA, transcrip-
provide evidence demonstrating that transcription drives su- tion clearly accelerates supercoiling and transitions to second-
percoiling in vivo (1, 19, 20, 27, 86) and that the wave gener- ary DNA structures (1, 19, 20, 27, 47, 59, 69, 70, 81, 86).
ated can be as long as 800 bp (47). Negative supercoiling
induces and stabilizes a transition from the right-handed B- SECONDARY DNA STRUCTURES: ARE THEY
form to the left-handed Z-form of DNA (42); a chemical assay PRECURSORS TO MUTATIONS?
detecting these distortions reveals that transcription-induced
supercoiling is highly localized (47, 86). During the induction Almost 40 years ago, Benzer (5) demonstrated that the mu-
of transcription, supercoils are found inside each transcribed tability of specific sites in the genome varied by orders of
region, as well as upstream and downstream of each individual magnitude, suggesting that these differences in mutation rate
RNAP complex. Transcription from a strong promoter leads to might reflect particular characteristics of the DNA sequence
greater negative supercoiling than transcription from a weak associated with hot spots. In fact, mutable sites are frequently
one (27). A major role of DNA topoisomerase I is now con- the consequence of their location within DNA secondary struc-
sidered to be the relaxation of local negative supercoiling dur- tures. Simple stem-loops arise from ssDNA sequences contain-
ing transcription, thus preventing unacceptably high levels of ing two segments that are inverted complements, usually about
supercoiling and associated R-loops that form when nascent 10 to 15 bases long, separated by 5 to 10 noncomplementary
RNA moves behind the advancing RNAP to bond with its bases that become the loop at the end of the stem formed by
original template DNA (69, 70, 105). The bulk of plasmid hydrogen bonding of the two complementary segments (Fig.
DNA does not exhibit stem-loop conformations during loga- 1C). These structures are called hairpins if the loop is very
rithmic growth. However, supercoiling may play a particularly small and cruciforms if they form opposite one another in each
important role in stressed cells, in which a disruption can occur DNA strand. Perfect complementarity (a palindrome) is rare;
between transcription and translation, thus promoting both the more common quasipalindromes or stem-loops contain
R-loop formation and supercoiling (68, 69). The inhibition of bases that are left unpaired or mispaired and therefore vulner-
protein synthesis by chloramphenicol, which uncouples tran- able to deamination (mutations 1 and 2 in Fig. 1C), deletion
scription and translation, induces stem-loop formation in the (mutations 3 and 4), replacement (mutations 5 and 6), or
overwhelming majority of DNA molecules (19). complementation by the insertion of a new base to the struc-
When E. coli is grown with limiting levels of glucose (Fig. 2), ture (mutation 7). The stem-loop is in dynamic equilibrium
a burst in supercoiling occurs precisely at the moment of glu- with linearized DNA, and changes such as those indicated in
cose depletion, as the cells cease logarithmic growth and enter Fig. 1C only become fixed as mutations in the course of further
stationary phase (1). This is also the moment at which a sharp metabolic events such as repair or replication (Fig. 1D). For
increase occurs in the concentration of cyclic AMP (cAMP) example, the entire structure depicted in Fig. 1C will be ex-
(10), ppGpp (23, 51), ␴S (50), and about 30 new proteins, cluded and deleted by virtue of new DNA synthesis across the
including ␤-galactosidase (34, 37). The increase in negative base of its stem. However, if this structure returns to its linear
supercoiling under these circumstances can be due to the in- form prior to new DNA synthesis, the potential changes indi-
crease in transcription known to occur as a result of derepres- cated above can be immortalized due to synthesis templated by
2998 MINIREVIEW J. BACTERIOL.

the altered sequence. An example of this process is indicated in


Fig. 1D, in which a C in the loop is deaminated to uracil, which
codes for A, which then codes for T during DNA synthesis,
resulting in a C-to-T transition.
What is the probability that these structures actually exist in
vivo and constitute precursors of background mutations in
nature? One kind of evidence for their existence is the striking
correlation of deletion end points with tandem repeats and the
ends of potential secondary structures. The leuB, argH, and
pyrD mutations in E. coli all occur at the end of predicted
stem-loops (111). The lacI gene has frequently been used as a
model system for investigating these correlations by comparing
its nucleotide sequence (26) to those of various mutant strains.
Among the sequenced deletion mutations in lacI, 7 occur in
tandem repeats, 6 consist of deleted stem-loop structures, and
4 of these 13 mutations share both characteristics. Although
three remaining mutations fail to coincide exactly with pre-
dicted vulnerable sites, they may be explained by different
DNA secondary structural intermediates including interstrand
misalignments (26, 88, 92). Todd and Glickman (99) analyzed
102 amber and 71 ochre mutants in the lacI gene, correlating
mutation hot spots with the locations of predicted unpaired
sites, many of which are located in stem-loop structures. Mu-

Downloaded from jb.asm.org by on November 23, 2009


tational hot spots are highly localized, and 50% of the non-
sense mutations arose in a segment comprising only 6% of the
DNA sequence analyzed. Each hot spot is found to be located FIG. 3. An algorithm for evolution.
at an unpaired site within potential secondary structures.
Clearly, these correlations depict causal relationships.
The mechanism of frameshift mutagenesis has also been that the sequence-dependent secondary structures created and
examined in vitro during DNA polymerization (80). In this stabilized by supercoiling are precursors to mutations.
system, sequence misalignments result from intrastrand com-
plementary pairings between two segments within a single CONCLUSIONS
newly synthesized strand, as well as from interstrand pairings
between a segment in the new strand and a complementary Many scientists may share Dobzhansky’s intuitive conviction
sequence in the template strand. When these misaligned seg- that the marvelous intricacies of living organisms could not
ments are used as templates for DNA synthesis, mutant se- have arisen by the selection of truly random mutations. This
quences are produced. Polymerase pausing, or the local rate of minireview suggests that sensitive, directed feedback mecha-
DNA polymerization, was also measured and correlated with nisms initiated by different kinds of stress might facilitate and
the misalignments, since pausing is sequence specific as well accelerate the adaptation of organisms to new environments.
(43). Pausing serves to increase the time of exposure of muta- The specificity in the series of events summarized by Fig. 3
ble bases and is known to promote mutagenesis (2, 3). The resides entirely in the first step, which is meant to suggest a
correlation between pausing, positions of frameshift misalign- pattern of derepression elicited by a corresponding pattern of
ments, and subsequent deletions can account for 97% of the adverse conditions. Microorganisms in nature must be con-
mutations observed. Moreover, the most common mutations fronted simultaneously by a complex set of problems, for ex-
coincide precisely with the strongest pause sites, and the ter- ample, the threat of oxidative or osmotic damage together with
mini of pause sites correlate with the sequence at which the suboptimum concentrations of many essential nutrients. Tran-
deletions begin. scriptional activation of genes derepressed to various degrees
Thus, in vivo and in vitro investigations strongly implicate would expose the nontranscribed strands to mutations and
the existence of DNA secondary structures as mutagenic sub- stimulate localized supercoiling. Vulnerable bases in the com-
strates and/or as structural precursors to mutagenic substrates plex DNA structures resulting from supercoiled DNA will also
that give rise to mutations that are immortalized during new contribute to localized hypermutation in the genes activated to
DNA synthesis or repair (Fig. 1D). As discussed earlier, mu- cope with the stresses that initiate the above series of events.
tations occur preferentially in ssDNA and in unpaired and A multitude of random mechanisms result in hypermutation
mispaired bases (28, 31, 32, 48, 60). Other kinds of evidence under conditions of environmental stress and clearly contrib-
also support a role for secondary structures as precursors of ute to the variability essential to evolution. However, since
mutations. Many insertion mutations (Fig. 1C) can best be most mutations are deleterious, random mechanisms that in-
explained if the other strand of a predicted transient stem were crease mutation rates also result in genomewide DNA damage.
used as a template for DNA synthesis prior to replication. The Among microorganisms, from phage to fungi, the overall mu-
fact that mutations are grouped closely together much more tation rate per genome is remarkably constant (within 2.5-
frequently than could occur by chance implicates a single ini- fold), presumably reflecting an obligatory, delicate balance be-
tiating event (structure). As the researchers of the above in- tween the need for variation and the need to avoid general
vestigations point out, other variables undoubtedly contribute genetic damage (24, 45, 57). Thus, mutator strains are not
to and modify the correlation observed between DNA se- selected in nature but remain at 1 to 2% of the population (35,
quences, misaligned structures, polymerase pausing, and mu- 52); under certain adverse conditions, they flourish for short
tations. Nevertheless, these investigations are but a small frac- periods but are then selected against, apparently because of
tion of an enormous literature providing compelling evidence widespread deleterious effects intrinsic to genomewide hyper-
VOL. 182, 2000 MINIREVIEW 2999

mutation. In contrast, hypermutation that is the consequence 14. Chatterji, D., N. Fujita, and A. Ishihama. 1998. The mediator for stringent
of starvation-induced derepression and transcriptional activa- control, ppGpp, binds to the beta-subunit of Escherichia coli RNA poly-
merase. Genes Cells 3:279–287.
tion represents a very rapid and specific response to each 15. Coulondre, C., J. H. Miller, P. J. Farabaugh, and W. Gilbert. 1978. Mo-
adverse circumstance. The extent to which normal background lecular basis of base substitution hotspots in Escherichia coli. Nature 274:
mutations in nature are due to derepression mechanisms is 775–780.
difficult to estimate, but the location of most C-to-T transitions 16. Datta, A., and S. Jinks-Robertson. 1995. Association of increased sponta-
neous mutation rates with high levels of transcription in yeast. Science
on the nontranscribed strand suggest that it may be significant. 268:1616–1619.
Regardless, a mechanism that limits an increase in mutation 17. Davis, B. D. 1989. Transcriptional bias: a non-Lamarkian mechanism for
rates to genes that must mutate in order to overcome prevail- substrate-induced mutations. Proc. Natl. Acad. Sci. USA 96:5005–5009.
ing conditions of stress would surely be beneficial and there- 18. Davis, M. G., and J. M. Calvo. 1977. Relationship between messenger
ribonucleic acid and enzyme levels specified by the leucine operon of
fore selected during evolution. Escherichia coli K-12. J. Bacteriol. 131:997–1007.
The environment gave rise to life and continues to direct 19. Dayn, A., S. Malkhosyan, D. Duzhy, V. Lyamichev, Y. Panchenko, and S.
evolution. Environmental conditions are constantly controlling Mirkin. 1991. Formation of (dA-dT)n cruciforms in Escherichia coli cells
and fine-tuning the transcriptional machinery of the cell. Feed- under different environmental conditions. J. Bacteriol. 173:2658–2664.
20. Dayn, A., S. Malkhosyan, and S. M. Mirkin. 1992. Transcriptionally driven
back mechanisms represent the natural interactive link be- cruciform formation in vivo. Nucleic Acids Res. 20:5991–5997.
tween an organism and its environment. An obvious selective 21. Dobzhansky, T. 1950. The genetic basis of evolution. Sci. Am. 182:32–41.
advantage exists for a relationship in which particular environ- 22. Dobzhansky, T. 1950. Heredity, environment, and evolution. Science 111:
mental changes are metabolically linked through transcription 161–166.
23. Donini, P., V. Santonastaso, J. Roche, and A. J. Cozzone. 1978. The rela-
to genetic changes that help an organism cope with new de- tionship between guanosine tetraphosphate, polysomes and RNA synthesis
mands of the environment. In nature, nutritional stress and in amino acid starved Escherichia coli. Mol. Biol. Rep. 4:15–29.
associated genetic derepression must be rampant. If mutation 24. Drake, J. W. 1991. A constant rate of spontaneous mutation in DNA-based
rates can be altered by the many variables controlling specific, microbes. Proc. Natl. Acad. Sci. USA 88:7160–7164.
25. Drlica, K., R. J. Franco, and T. R. Steck. 1988. Rifampin and rpoB muta-
stress-induced transcription, one might reasonably argue that

Downloaded from jb.asm.org by on November 23, 2009


tions can alter DNA supercoiling in Escherichia coli. J. Bacteriol. 170:4983–
many mutations are to some extent directed as a result of the 4985.
unique metabolism of every organism responding to the chal- 26. Farabaugh, P. J., U. Schmeisser, M. Hofer, and J. H. Miller. 1978. Genetic
lenges of its environment. Thus, mutations are brought within studies of the lac repressor. VII. On the molecular nature of spontaneous
hotspots in the lacI gene of Escherichia coli. J. Mol. Biol. 126:847–863.
the realm of metabolic events to become the final, irreversible 27. Figueroa, N., and L. Bossi. 1988. Transcription induces gyration of the
act of metabolism in the constant struggle to adapt or die. DNA template in Escherichia coli. Proc. Natl. Acad. Sci. USA 85:9416–
9420.
ACKNOWLEDGMENTS 28. Fix, D. F., and B. W. Glickman. 1987. Asymmetric cytosine deamination
revealed by spontaneous mutational specificity in an UngG strain of Esch-
I am indebted to my colleagues in the Division of Biological Sciences erichia coli. Mol. Gen. Genet. 209:78–82.
and to Lynn Ripley and Karl Drlica for suggestions and valuable 29. Forsdyke, D. R. 1995. Conservation of stem-loop potential in introns of
criticisms of this manuscript. snake venom phospholipase A2 genes: an application of FORS-D analysis.
This work was supported in part by the Eppley Foundation, PHS Mol. Biol. Evol. 12:1157–1165.
30. Forsdyke, D. R. 1995. Reciprocal relationship between stem-loop potential
grant GM/OD54279, a MONTS grant from the NSF EPSCoR pro- and substitution density in retroviral quasispecies under positive Darwinian
gram, and by the University of Montana. selection. J. Mol. Evol. 41:1022–1037.
31. Frederico, L. A., T. A. Kunkel, and B. R. Shaw. 1990. A sensitive genetic
REFERENCES assay for the detection of cytosine deamination: determination of rate
1. Balke, V. L., and J. D. Gralla. 1987. Changes in the linking number of constants and the activation energy. Biochemistry 29:2532–2537.
supercoiled DNA accompany growth transitions in Escherichia coli. J. Bac- 32. Frederico, L. A., T. A. Kunkel, and B. A. Shaw. 1993. Cytosine deamination
teriol. 169:4499–4506. in mismatched base pairs. Biochemistry 32:6523–6530.
2. Bebenek, K., C. M. Joyce, M. P. Fitzgerald, and T. A. Kunkel. 1990. The 33. Gentry, D. R., V. J. Hernandez, L. H. Nguyen, D. B. Jensen, and M. Cashel.
fidelity of DNA synthesis catalyzed by derivatives of Escherichia coli DNA 1993. Synthesis of the stationary-phase sigma factor ␴S is positively regu-
polymerase I. J. Biol. Chem. 265:13878–13887. lated by ppGpp. J. Bacteriol. 175:7982–7989.
3. Bebenek, K., J. Abbotts, S. H. Wilson, and T. A. Kunkel. 1993. Error-prone 34. Groat, R. G., and A. Matin. 1986. Synthesis of unique proteins at the onset
polymerization by HIV-1 reverse transcriptase. J. Biol. Chem. 268:10324– of carbon starvation in Escherichia coli. J. Ind. Microbiol. 1:69–73.
10334. 35. Gross, M. D., and E. C. Siegel. 1981. Incidence of mutator strains and
4. Beletskii, A., and A. S. Bhagwat. 1996. Transcription-induced mutations: coliforms in nature. Mutat. Res. 91:107–110.
increase in C-to-T mutations in the non-transcribed strand during transcrip- 36. Hanawalt, P., and I. Mellon. 1993. Stranded in an active gene. Curr. Biol.
tion in Escherichia coli. Proc. Natl. Acad. Sci. USA 93:13919–13924. 3:67–69.
5. Benzer, S. 1961. On the topography of the genetic fine structure. Proc. Natl. 37. Hengge-Aronis, R. 1996. Regulation of gene expression during entry into
Acad. Sci. USA 47:403–415. stationary phase, p. 1497–1512. In F. C. Neidhardt, R. Curtis III, J. L.
6. Bernhard, R. 1967. Heresy in the halls of biology: mathematicians question Ingraham, E. C. C. Lin, K. B. Low, B. Magasanik, W. S. Reznikoff, M.
Darwinism. Sci. Res. (New York) 2:59–66. Riley, M. Schaechter, and H. Umbarger (ed.), Escherichia coli and Salmo-
7. Bick, M. D., C. S. Lee, and C. A. Thomas. 1972. Local destabilization of nella: cellular and molecular biology, 2nd ed., vol. 1. American Society for
DNA during transcription. J. Mol. Biol. 71:1–9. Microbiology, Washington, D.C.
8. Brock, R. D. 1971. Differential mutation of the ␤-galactosidase gene of 38. Herman, R. K., and N. B. Dworkin. 1971. Effect of gene induction on the
Escherichia coli. Mutat. Res. 11:181–186. rate of mutagenesis by ICR-191 in Escherichia coli. J. Bacteriol. 106:543–550.
9. Bruhat, A., C. Jousse, and P. Fafournoux. 1999. Amino acid limitation 39. Hinnebusch, A. G. 1988. Mechanisms of gene regulation in the general
regulates gene expression. Proc. Nutr. Soc. 58:625–632. control of amino acid biosynthesis in Saccharomyces cerevisiae. Microbiol.
10. Buettner, M. J., E. Spitz, and H. V. Rickenberg. 1973. Cyclic adenosine 3⬘, Rev. 52:248–273.
5⬘-monophosphate in Escherichia coli. J. Bacteriol. 14:1068–1073. 40. Horowitz, N. H. 1945. On the evolution of biochemical syntheses. Proc.
11. Cairns, J., and P. L. Foster. 1991. Adaptive reversion of a frameshift Natl. Acad. Sci. USA 31:153–157.
mutation in Escherichia coli. Genetics 128:695–701. 41. Jablonski, D., J. Sepkoski, Jr., D. J. Bottjer, and P. M. Sheehan. 1983.
12. Carter, P. W., D. L. Weiss, H. L. Weith, and J. M. Calvo. 1985. Mutations Origin of species in stressed environments. Science 222:1112–1124.
that convert the four leucine codons of the Salmonella typhimurium leu 42. Jaworski, A., N. P. Higgins, R. D. Wells, and W. Zacharias. 1991. Topo-
leader to four threonine codons. J. Bacteriol. 162:943–949. isomerase mutants and physiological conditions control supercoiling and
13. Cashel, M., D. R. Gentry, V. J. Hernandez, and D. Vinella. 1996. The Z-DNA formation in vivo. J. Biol. Chem. 266:2576–2581.
stringent response, p. 1458–1496. In F. C. Neidhardt, R. Curtis III, J. L. 43. Kaguni, L. S., and D. A. Clayton. 1982. Template-directed pausing in in
Ingraham, E. C. C. Lin, K. B. Low, B. Magasanik, W. S. Reznikoff, M. vitro DNA synthesis by DNA polymerase ␣ from Drosophila melanogaster
Riley, M. Schaechter, and H. Umbarger. (ed.), Escherichia coli and Salmo- embryos. Proc. Natl. Acad. Sci. USA 79:983–987.
nella: cellular and molecular biology, 2nd ed., vol. 1. American Society for 44. Kim, K.-H. 1963. Isolation and properties of a putrescine-degrading mutant
Microbiology, Washington, D.C. of Escherichia coli. J. Bacteriol. 86:320–323.
3000 MINIREVIEW J. BACTERIOL.

45. Kimura, R. 1960. Optimum mutation rates and degree of dominance as 75. Mortlock, R. P. 1984. Microorganisms as model systems for studying evo-
determined by the principle of minimum genetic load. J. Genet. 57:21–34. lution. Plenum Press, New York, N.Y.
46. Koehler, D. R., S. S. Awadallah, and B. W. Glickman. 1991. Sites of 76. Naas, T., M. Blot, W. M. Fitch, and W. Arber. 1995. Dynamics of IS-related
preferential induction of cyclobutane pyrimidine dimers in the nontran- genetic rearrangements in resting Escherichia coli K-12. Mol. Biol. Evol.
scribed strand of lacI correspond with the sites of UV-induced mutation in 12:198–207.
Escherichia coli. J. Biol. Chem. 266:11766–11773. 77. Ninfa, A. J. 1996. Regulation of gene transcription by external stimuli,
47. Krasilnikov, A. S., A. Podtelezhnikov, A. Vologodskii, and S. M. Mirkin. p. 1246–1262. In F. C. Neidhardt, R. Curtis III, J. L. Ingraham, E. C. C. Lin,
1999. Large-scale effects of transcriptional DNA supercoiling in vivo. J. K. B. Low, B. Magasanik, W. S. Reznikoff, M. Riley, M. Schaechter, and H.
Mol. Biol. 292:1149–1160. Umbarger (ed.), Escherichia coli and Salmonella: cellular and molecular
48. Kunkel, T. A. 1984. Mutational specificity of depurination. Proc. Natl. biology, 2nd ed., vol. 1. American Society for Microbiology, Washington, D.C.
Acad. Sci. USA 81:1494–1498. 78. Olivera, B. M. 1997. Conus venom peptides, receptor and ion channel
49. Landick, R., C. L. Turnbough, Jr., and C. Yanofsky. 1996. Transcription targets and drug design: 50 million years of neuropharmacology. Mol. Biol.
attenuation, p. 1263–1286. In F. C. Neidhardt, R. Curtis III, J. L. Ingraham, Cell 8:2101–2109.
E. C. C. Lin, K. B. Low, B. Magasanik, W. S. Reznikoff, M. Riley, M. 79. Oparin, A. I. 1938. The origin of life. Macmillan, New York, N.Y. [Trans-
Schaechter, and H. Umbarger (ed.), Escherichia coli and Salmonella: cel- lated by S. Margulis.]
lular and molecular biology, 2nd ed., vol. 1. American Society for Micro- 80. Pananicolaou, C., and L. S. Ripley. 1991. An in vitro approach to identifying
biology, Washington, D.C. specificity determinants of mutagenesis mediated by DNA misalignments.
50. Lange, R., and R. Hengge-Aronis. 1994. The cellular concentration of the J. Mol. Biol. 221:805–821.
␴S subunit of RNA polymerase in Escherichia coli is controlled at the levels 81. Panayotatos, N., and R. D. Wells. 1981. Cruciform structures in supercoiled
of transcription, translation, and protein stability. Genes Dev. 8:1600–1612. DNA. Nature 289:446–470.
51. Lazzarini, R. A., M. Cashel, and J. Gallant. 1971. On the regulation of 82. Prescott, D. M. 1992. The unusual organization and processing of genomic
guanosine tetraphosphate levels in stringent and relaxed strains of Esche- DNA in hypotrichous ciliates. Trends Genet. 8:439–445.
richia coli. J. Biol. Chem. 246:4381–4385. 83. Primakoff, P. 1981. In vivo role of the relA⫹ gene in regulation of the lac
52. Le Clerk, J. E., B. Li, W. L. Payne, and T. A. Cebula. 1996. High mutation operon. J. Bacteriol. 145:410–416.
frequencies among Escherichia coli and Salmonella pathogens. Science 274: 84. Primakoff, P., and S. W. Artz. 1979. Positive control of lac operon expres-
1208–1211. sion in vitro by guanosine 5N-diphosphate 3N-diphosphate. Proc. Natl.
53. Lederberg, J. 1951. Genetic studies with bacteria, p. 263–289. In L. C. Dunn Acad. Sci. USA 76:1726–1730.
(ed.), Genetics in the 20th century. Macmillan, New York, N.Y. 85. Pruss, G. J., and K. Drlica. 1989. DNA supercoiling and prokaryotic tran-

Downloaded from jb.asm.org by on November 23, 2009


54. Lee, C.-K., R. G. Klopp, R. Weindruch, and T. A. Prolla. 1999. Gene scription. Cell 56:521–523.
expression profile of aging and its retardation by caloric restriction. Science 86. Rahmouni, A. R., and R. D. Wells. 1992. Direct evidence for the effect of
285:1390–1393. transcription on local DNA supercoiling in vivo. J. Mol. Biol. 223:131–144.
55. Lerner, S. A., and E. C. C. Lin. 1962. A method for isolating constitutive 87. Rigby, P. W. J., B. D. Burleigh, Jr., and B. S. Hartley. 1974. Gene dupli-
mutants for carbohydrate-catabolizing enzymes. Fed. Proc. 21:268. cation in experimental enzyme evolution. Nature 251:200–204.
56. Lerner, S. A., T. T. Wu, and E. C. C. Lin. 1964. Evolution of a catabolic 88. Ripley, L. S., and B. W. Glickman. 1983. Unique self-complementarity of
pathway in bacteria. Science 146:1313–1315. palindromic sequences provides DNA structural intermediates for muta-
57. Levins, R. 1967. Theory of fitness in a heterogenious environment. VI. The tion. Cold Spring Harbor Symp. Quant. Biol. 47:851–861.
adaptive significance of mutation. Genetics 56:163–178. 89. Rosche, W. A., L. S. Ripley, and R. Sinden. 1998. Primer-template misalign-
58. Lewis, E. B. 1951. Pseudoallelism and gene evolution. Cold Spring Harbor ments during leading strand DNA synthesis account for the most frequent
Symp. Quant. Biol. 16:159–174. spontaneous mutations in a quasipalindromic region in Escherichia coli. J.
59. Lilley, D. M. J. 1980. The inverted repeat as a recognizable structural Mol. Biol. 284:633–646.
feature in supercoiled molecules. Proc. Natl. Acad. Sci. USA 77:6468–6472. 90. Rudner, R. A. Murray, and N. Huda. 1999. Is there a link between mutation
60. Lindahl, T., and B. Nyberg. 1974. Heat-induced deamination of cytosine rates and the stringent response in Bacillus subtilus? Ann. N. Y. Acad. Sci.
residues in deoxyribonucleic acid. Biochemistry 13:3405–3410. 870:418–422.
61. Liu, L. F., and J. C. Wang. 1987. Supercoiling and the DNA template 91. Saier, M. H., Jr., T. M. Ramseier, and J. Reizer. 1996. Regulation of carbon
during transcription. Proc. Natl. Acad. Sci. 84:7024–7027. utilization, p. 1325–1343. In F. C. Neidhardt, R. Curtis III, J. L. Ingraham,
62. Longacre, A. L., J. M. Reimers, and B. E. Wright. 1999. Specificity of E. C. C. Lin, K. B. Low, B. Magasanik, W. S. Reznikoff, M. Riley, M.
transcription-enhanced mutations. Ann. N. Y. Acad. Sci. 870:383–385. Schaechter, and H. Umbarger (ed.), Escherichia coli and Salmonella: cel-
63. Luria, S., and M. Delbruck. 1943. Mutations of bacteria from virus sensi- lular and molecular biology, 2nd ed., vol. 1. American Society for Micro-
tivity to virus resistance. Genetics 28:491–504. biology, Washington, D.C.
64. Maden, B. E. H. 1995. No soup for starters? Autotrophy and the origins of 92. Schaaper, R. M., B. N. Danforth, and B. L. Glickman. 1986. Mechanisms of
metabolism. Trends Biochem. Sci. 20:337–341. spontaneous mutagenesis: an analysis of the spectrum of spontaneous mu-
65. Magasanik, B. 1996. Regulation of nitrogen utilization, p. 1344–1356. In tation in the Escherichia coli lacI gene. J. Mol. Biol. 189:273–284.
F. C. Neidhardt, R. Curtis III, J. L. Ingraham, E. C. C. Lin, K. B. Low, B. 93. Selker, E. U. 1990. Premeiotic instability of repeated sequences in Neuro-
Magasanik, W. S. Reznikoff, M. Riley, M. Schaechter, and H. Umbarger spora crassa. Annu. Rev. Genet. 24:579–613.
(ed.), Escherichia coli and Salmonella: cellular and molecular biology, 2nd 94. Sobell, H. M. 1972. Molecular mechanism for genetic recombination. Proc.
ed., vol. 1. American Society for Microbiology, Washington, D.C. Natl. Acad. Sci. USA 69:2483–2487.
66. Mager, W. H., and A. J. J. De Kruijff. 1995. Stress-induced transcriptional 95. Sonti, R. V., and J. R. Roth. 1989. Role of gene duplication in the adapta-
activation. Microbiol. Rev. 59:506–531. tion of Salmonella typhimurium to growth on limiting carbon sources. Ge-
67. Makino, K., M. Amemura, S.-K. Kim, A. Nakata, and H. Shinagara. 1994. netics 123:19–28.
Mechanism of transcriptional activation of the phosphate regulon in Esch- 96. Stemmer, W. P. 1994. Rapid evolution of a protein in vitro by DNA
erichia coli, p. 5–12. In A. Torriani-Gorini, E. Yagil, and S. Silver (ed.), shuffling. Nature 370:389–391.
Phosphate in microorganisms: cellular and molecular biology. ASM Press, 97. Stephens, J. C., S. W. Artz, and B. Ames. 1975. Guanosine 5N-diphosphate
Washington, D.C. 3N-diphosphate (ppGpp): positive effector for histidine operon transcrip-
68. Masse, E., P. Phoenix, and M. Drolet. 1997. DNA topoisomerases regulate tion and general signal for amino acid deficiency. Proc. Natl. Acad. Sci.
R-loop formation during transcription of the rrnB operon in Escherichia USA 72:4389–4393.
coli. J. Biol. Chem. 272:12816–12823. 98. Taddei, F., M. Radman, J. Maynard-Smith, B. Toupance, P. H. Gouyon,
69. Masse, E., and M. Drolet. 1999. Relaxation of transcription-induced neg- and B. Godelle. 1997. Role of mutator alleles in adaptive evolution. Nature
ative supercoiling is an essential function of Escherichia coli DNA topo- 387:700–702.
isomerase I. J. Biol. Chem. 274:16654–16658. 99. Todd, P. A., and B. W. Glickman. 1982. Mutational specificity of UV light
70. Masse, E., and M. Drolet. 1999. Escherichia coli DNA topoisomerase I in Escherichia coli: indications for a role of secondary structure. Proc. Natl.
inhibits R-loop formation by relaxing transcription-induced negative super- Acad. Sci. USA 79:4123–4127.
coiling. J. Biol. Chem. 274:16659–16664. 100. Torriani, A., and F. Rothman. 1961. Mutants of Escherichia coli constitutive
71. Mayr, E. 1982. The growth of biological thought: Diversity, evolution, and for alkaline phosphatase. J. Bacteriol. 81:835–840.
inheritance. Belknap Press of Harvard University Press, Cambridge, Mass. 101. Vilette, D., M. Uzest, D. Ehrlich, and B. Michel. 1992. DNA transcription
72. McCann, M. P., J. P. Kidwell, and A. Matin. 1991. The putative ␴ factor has and repressor binding affect deletion formation in Escherichia coli plas-
a central role in development of starvation-mediated general resistance in mids. EMBO J. 11:3629–3634.
Escherichia coli. J. Bacteriol. 173:4188–4194. 102. von Hipple, P. H., and K.-Y. Wong. 1971. Dynamic aspects of native DNA
73. McClintock, B. 1950. The origin and behavior of mutable loci in maize. structures: kinetics of the formaldehyde reaction with calf thymus DNA.
Proc. Natl. Acad. Sci. USA 36:344–355. J. Mol. Biol. 61:587–613.
74. Morey, N. J., C. N. Greene, and S. Jinks-Robertson. 2000. Genetic analysis 103. Wächterhäuser, G. 1990. Evolution of the first metabolic cycles. Proc. Natl.
of transcription-associated mutation in Saccharomyces cerevisiae. Genetics Acad. Sci. USA 87:200–204.
154:109–120. 104. Walker, G. C. 1995. SOS-regulated proteins in translesion DNA synthesis
VOL. 182, 2000 MINIREVIEW 3001

and mutagenesis. Trends Biochem. Sci. 20:416–420. 110. Wright, B. E., and M. F. Minnick. 1997. Reversion rates in a leuB auxo-
105. Wang, J. C. 1985. DNA topoisomerases. Annu. Rev. Biochem. 54:665–697. troph of Escherichia coli K-12 correlate with ppGpp levels during expo-
106. Weismann, A. 1893. Über die Vererbung. Gustav Fischer, Jena, Germany. nential growth. Microbiology 143:847–854.
107. Wessler, S. R., and J. M. Calvo. 1981. Control of the leu operon expression 111. Wright, B. E., A. Longacre, and J. M. Reimers. 1999. Hypermutation in
in Escherichia coli by a transcription attenuation mechanism. J. Mol. Biol. derepressed operons of Escherichia coli K12. Proc. Natl. Acad. Sci. USA
149:579–597. 96:5089–5094.
108. Woese, C. R. 1987. Bacterial evolution. Microbiol. Rev. 51:221–271. 112. Wu, T. T., E. C. C. Lin, and S. Tanaka. 1968. Mutants of Aerobacter
109. Wright, B. E. 1996. The effect of the stringent response on mutations in aerogenes capable of utilizing xylitol as a novel carbon. J. Bacteriol. 96:447–
Escherichia coli K-12. Mol. Microbiol. 19:213–219. 456.

Downloaded from jb.asm.org by on November 23, 2009

You might also like