throbber
GDA201.QXD 03/22/2000 02:15 Page 162
`
`162
`
`Sloppier copier DNA polymerases involved in genome repair
`Myron F Goodman* and Brigette Tippin
`
`When chromosomal replication is impeded in the presence of
`DNA damage, members of a newly discovered
`UmuC/DinB/Rev1/Rad30 superfamily of procaryotic and
`eucaryotic DNA polymerases catalyze translesion synthesis at
`blocked replication forks. Although these polymerases share
`sequence elements essentially unrelated to the standard
`replication and repair enzymes, some of them (such as the
`SOS-induced Escherichia coli pol V) catalyze ‘error-prone’
`translesion synthesis leading to large increases in mutation,
`whereas others (an example being the Xeroderma
`pigmentosum variant gene product XPV pol h ) carry out
`aberrant, yet nonmutagenic translesion synthesis. Ongoing
`studies of these low fidelity polymerases could provide new
`insights into the mechanism of somatic hypermutation, a key
`element in the immune response.
`
`Addresses
`Department of Biological Sciences and Chemistry, University of
`Southern California, University Park, Los Angeles, California 90089-
`1340, USA
`*e-mail: mgoodman@mizar.usc.edu
`
`Current Opinion in Genetics & Development 2000, 10:162–168
`
`0959-437X/00/$ — see front matter © 2000 Elsevier Science Ltd.
`All rights reserved.
`
`Abbreviations
`HE
`holoenzyme complex
`pol
`DNA polymerase
`pol V Mut
`pol V mutasome
`SBB
`DNA single-stranded binding protein
`TLS
`translesion synthesis
`XPV
`Xeroderma pigmentosum variant gene product
`
`Introduction
`The first DNA polymerase (pol), Escherichia coli pol I, was
`discovered in 1957 by Arthur Kornberg (for review, see
`[1]). Twelve years later John Cairns [2] isolated a strain of
`E. coli containing a mutant pol I enzyme leading to the dis-
`coveries of pol II and pol III (for review, see [1]). Pol III is
`responsible primarily for replicating the bacterial genome,
`while pol I plays a major role in UV damage repair and in
`Okazaki fragment processing (for review, see [1]). The
`enigmatic pol II was recently shown to be involved in the
`reactivation of replication complexes stalled at DNA tem-
`plate lesions [3]. Thirty years have now passed since the
`discovery of the pol I mutant. Remarkably, the past
`18 months have witnessed the discovery of a variety of new
`procaryotic and eucaryotic DNA polymerases, including
`two more in E. coli.
`
`This review discusses these new DNA-damage tolerant
`polymerases with special emphasis placed on the role of
`the error-prone UmuD¢ 2C complex (E. coli pol V) in the
`well-documented SOS mutagenic response in E. coli. We
`provide an overview of the relationships between the novel
`
`UmuC/DinB/Rev1/Rad30 superfamily of DNA polymerases
`spanning procaryotic and eucaryotic organisms.
`
`A brief synopsis of the E. coli SOS-regulon
`E. coli’s SOS response involves the action of at least 25
`genes regulated at the transcriptional level by LexA
`repressor protein. Following damage to DNA, the LexA
`repressor undergoes proteolysis, mediated by RecA pro-
`tein acting as a coprotease, turning on SOS gene
`expression (Figure 1). Many of the proteins induced early
`in the SOS response are involved in nucleotide excision
`repair and recombination repair pathways [4]. These path-
`ways are ‘error-free’ meaning that they do not cause
`mutations above spontaneous background levels; however,
`it has been known since 1977 that a large, ~100-fold,
`increase in mutations accompanies SOS induction. This
`involves the action of RecA protein and UmuD¢ and
`UmuC proteins [5–7], where ‘Umu’ refers to UV mutage-
`nesis. RecA protein
`is
`required
`to process
`the
`mutagenically inactive UmuD to UmuD¢ — a shorter,
`mutagenically active form of the protein — in a reaction
`analogous to the cleavage of LexA protein [8–10]. Follow-
`ing cleavage of UmuD to UmuD¢ , UmuC and two
`molecules of UmuD¢ associate to form a UmuD¢ 2C com-
`plex, which, in the presence of activated RecA protein
`filament (RecA*), catalyzes ‘error-prone’ translesion syn-
`thesis (TLS) causing mutations at DNA damage sites [11].
`RecA plays a direct biochemical role during SOS mutage-
`nesis that is distinct from generalized recombination and
`coproteolysis [10,12,13] and that is apparently responsible
`for targeting UmuD¢ 2C to a template-lesion site proximal
`to the tip of the RecA* filament [14,15].
`
`Mutagenically inactive complexes formed with UmuD2C
`and UmuD¢ DC are thought to act as a regulatory switch to
`turn off mutagenesis once DNA damage sites have been
`either repaired or bypassed [16]. Early reviews on SOS
`were written by Witkin [17] and Walker [4], and Friedberg
`et al. [18] provide a recent comprehensive review of the
`SOS regulatory system, written prior to the discoveries of
`error-prone E. coli pol IV and pol V.
`
`SOS translesion synthesis reconstituted
`in vitro
`A replication complex confronting a damaged DNA tem-
`plate strand may be likened to a major train wreck
`resulting in ‘derailment’ of the core polymerase and its
`accessory subunits. When faced with excessive amounts of
`DNA damage, the cell sends out an SOS signal, perhaps in
`the form of a segment of single-stranded chromosomal
`DNA bound by RecA protein. A specialized group of pro-
`teins are induced that can copy damaged template sites,
`making errors along the way. The proteins required for
`SOS-induced mutation (also called SOS error-prone repair)
`
`Columbia Ex. 2085
`Illumina, Inc. v. The Trustees
`of Columbia University in the
`City of New York
`IPR2020-00988, -01065,
`-01177, -01125, -01323
`
`

`

`GDA201.QXD 03/22/2000 02:15 Page 162
`
`162
`
`Sloppier copier DNA polymerases involved in genome repair
`Myron F Goodman* and Brigette Tippin
`
`When chromosomal replication is impeded in the presence of
`DNA damage, members of a newly discovered
`UmuC/DinB/Rev1/Rad30 superfamily of procaryotic and
`eucaryotic DNA polymerases catalyze translesion synthesis at
`blocked replication forks. Although these polymerases share
`sequence elements essentially unrelated to the standard
`replication and repair enzymes, some of them (such as the
`SOS-induced Escherichia coli pol V) catalyze ‘error-prone’
`translesion synthesis leading to large increases in mutation,
`whereas others (an example being the Xeroderma
`pigmentosum variant gene product XPV pol h ) carry out
`aberrant, yet nonmutagenic translesion synthesis. Ongoing
`studies of these low fidelity polymerases could provide new
`insights into the mechanism of somatic hypermutation, a key
`element in the immune response.
`
`Addresses
`Department of Biological Sciences and Chemistry, University of
`Southern California, University Park, Los Angeles, California 90089-
`1340, USA
`*e-mail: mgoodman@mizar.usc.edu
`
`Current Opinion in Genetics & Development 2000, 10:162–168
`
`0959-437X/00/$ — see front matter © 2000 Elsevier Science Ltd.
`All rights reserved.
`
`Abbreviations
`HE
`holoenzyme complex
`pol
`DNA polymerase
`pol V Mut
`pol V mutasome
`SBB
`DNA single-stranded binding protein
`TLS
`translesion synthesis
`XPV
`Xeroderma pigmentosum variant gene product
`
`Introduction
`The first DNA polymerase (pol), Escherichia coli pol I, was
`discovered in 1957 by Arthur Kornberg (for review, see
`[1]). Twelve years later John Cairns [2] isolated a strain of
`E. coli containing a mutant pol I enzyme leading to the dis-
`coveries of pol II and pol III (for review, see [1]). Pol III is
`responsible primarily for replicating the bacterial genome,
`while pol I plays a major role in UV damage repair and in
`Okazaki fragment processing (for review, see [1]). The
`enigmatic pol II was recently shown to be involved in the
`reactivation of replication complexes stalled at DNA tem-
`plate lesions [3]. Thirty years have now passed since the
`discovery of the pol I mutant. Remarkably, the past
`18 months have witnessed the discovery of a variety of new
`procaryotic and eucaryotic DNA polymerases, including
`two more in E. coli.
`
`This review discusses these new DNA-damage tolerant
`polymerases with special emphasis placed on the role of
`the error-prone UmuD¢ 2C complex (E. coli pol V) in the
`well-documented SOS mutagenic response in E. coli. We
`provide an overview of the relationships between the novel
`
`UmuC/DinB/Rev1/Rad30 superfamily of DNA polymerases
`spanning procaryotic and eucaryotic organisms.
`
`A brief synopsis of the E. coli SOS-regulon
`E. coli’s SOS response involves the action of at least 25
`genes regulated at the transcriptional level by LexA
`repressor protein. Following damage to DNA, the LexA
`repressor undergoes proteolysis, mediated by RecA pro-
`tein acting as a coprotease, turning on SOS gene
`expression (Figure 1). Many of the proteins induced early
`in the SOS response are involved in nucleotide excision
`repair and recombination repair pathways [4]. These path-
`ways are ‘error-free’ meaning that they do not cause
`mutations above spontaneous background levels; however,
`it has been known since 1977 that a large, ~100-fold,
`increase in mutations accompanies SOS induction. This
`involves the action of RecA protein and UmuD¢ and
`UmuC proteins [5–7], where ‘Umu’ refers to UV mutage-
`nesis. RecA protein
`is
`required
`to process
`the
`mutagenically inactive UmuD to UmuD¢ — a shorter,
`mutagenically active form of the protein — in a reaction
`analogous to the cleavage of LexA protein [8–10]. Follow-
`ing cleavage of UmuD to UmuD¢ , UmuC and two
`molecules of UmuD¢ associate to form a UmuD¢ 2C com-
`plex, which, in the presence of activated RecA protein
`filament (RecA*), catalyzes ‘error-prone’ translesion syn-
`thesis (TLS) causing mutations at DNA damage sites [11].
`RecA plays a direct biochemical role during SOS mutage-
`nesis that is distinct from generalized recombination and
`coproteolysis [10,12,13] and that is apparently responsible
`for targeting UmuD¢ 2C to a template-lesion site proximal
`to the tip of the RecA* filament [14,15].
`
`Mutagenically inactive complexes formed with UmuD2C
`and UmuD¢ DC are thought to act as a regulatory switch to
`turn off mutagenesis once DNA damage sites have been
`either repaired or bypassed [16]. Early reviews on SOS
`were written by Witkin [17] and Walker [4], and Friedberg
`et al. [18] provide a recent comprehensive review of the
`SOS regulatory system, written prior to the discoveries of
`error-prone E. coli pol IV and pol V.
`
`SOS translesion synthesis reconstituted
`in vitro
`A replication complex confronting a damaged DNA tem-
`plate strand may be likened to a major train wreck
`resulting in ‘derailment’ of the core polymerase and its
`accessory subunits. When faced with excessive amounts of
`DNA damage, the cell sends out an SOS signal, perhaps in
`the form of a segment of single-stranded chromosomal
`DNA bound by RecA protein. A specialized group of pro-
`teins are induced that can copy damaged template sites,
`making errors along the way. The proteins required for
`SOS-induced mutation (also called SOS error-prone repair)
`
`

`

`DNA polymerases involved in genome repair Goodman and Tippin 163
`
`GDA201.QXD 03/22/2000 02:15 Page 163
`
`Figure 1
`
`Model for UV induction of SOS genes in
`E. coli. (a) Binding of the LexA repressor (red
`squares) to regulatory operators upstream of
`SOS genes (black boxes) limits their
`expression under normal growth conditions.
`(b) Upon UV-induced DNA damage, RecA
`protein (green circles) becomes activated to
`RecA* by binding to regions of single-
`stranded DNA. RecA* can then act as a
`coprotease in the autocleavage of LexA
`allowing SOS genes to be turned on.
`
`have been known since the mid 1980s, thanks to extensive
`genetic data from many different laboratories [18]. These
`proteins include UmuC, UmuD¢ , RecA, and pol III
`holoenzyme complex (HE).
`
`In contrast to the extensive progress made in identifying
`the genetic elements required for SOS-induced mutation,
`attempts to identify biochemical roles for the SOS proteins
`were stymied by the insolubility of UmuC protein in
`aqueous solution. Nevertheless, Harrison Echols and co-
`workers [19] succeeded in purifying a denatured form of
`UmuC that, following renaturation, gave rise to low-level
`bypass of a site-directed abasic DNA-template lesion in
`vitro in the presence of UmuD¢ , RecA, and pol III HE.
`There remained considerable difficulties, however, obtain-
`ing reproducible yields and TLS activity using the
`denatured-renatured UmuC protein [20]. These difficulties
`were alleviated following purification of a soluble, native
`UmuD¢ 2C complex [21]. This complex actively catalyzed
`TLS [22••], as did a maltose-binding protein–UmuC
`(MBP-UmuC) fusion protein [23]. Both systems required
`RecA protein to catalyze TLS, with one surprising differ-
`ence: the native UmuD¢ 2C complex did not require the
`presence of pol III core to carry out TLS suggesting that it
`might contain an intrinsic DNA polymerase activity [22••].
`UmuD¢ 2C is a novel error-prone DNA
`polymerase, E. coli pol V
`To resolve the discrepancy between the two studies,
`UmuD¢ 2C was purified from a pol III temperature-sensi-
`tive strain containing a pol II deletion [24••] and was found
`
`to copy undamaged DNA at nonpermissive temperatures
`but required RecA to carry out TLS. A purified mutant
`complex, UmuD¢ 2C104 (Asp101fi Asn), failed to catalyze
`TLS. These data demonstrated that UmuD¢ 2C contained
`an intrinsic error-prone DNA polymerase activity, E. coli
`pol V. It was subsequently confirmed
`that
`the
`MBP–UmuC fusion protein also contained polymerase
`activity in the absence of the pol III core [25].
`
`Biochemical basis of SOS mutagenesis
`Having an in vitro assay available enables the following
`four basic questions to be addressed. What are the roles of
`each of the proteins required to catalyze TLS and most
`importantly what is the biochemical mechanism of pol V in
`relation to RecA protein? What are the efficiencies for
`bypassing diverse types of template DNA damage? How
`does the specificity of nucleotide incorporation measured
`in vitro compare with in vivo mutation spectra for different
`DNA lesions? What can be said about UmuD¢ 2C-catalyzed
`mutations at undamaged template sites?
`
`The proteins involved in lesion bypass are pol V
`(UmuD¢ 2C), RecA, b processivity clamp, g clamp-loading
`complex, and DNA single-stranded binding protein (SSB)
`[22••,24••]. Although pol V alone can form W–C base pairs
`with relatively low efficiency opposite undamaged tem-
`plate sites, it cannot catalyze incorporation opposite the
`commonly occurring abasic, cis-syn T–T dimer, or 6-4 T–T
`photoproduct lesions (M Tang, MF Goodman, unpub-
`lished data). While addition of RecA, b processivity clamp,
`g clamp-loading complex, or SSB stimulates pol V activity,
`
`

`

`GDA201.QXD 03/22/2000 02:15 Page 164
`
`164 Chromosomes and expression mechanisms
`
`Figure 2
`
`Pol V (UmuD¢ 2C) error-prone lesion bypass. Immediately following
`DNA damage and induction of the SOS response, E. coli attempt to
`repair their genome by various error-free mechanisms. (a) If any
`damage escapes these pathways and the replicative pol III HE
`complex encounters a DNA lesion, the pol III core is effectively blocked
`from further DNA synthesis and (b) dissociates from the DNA leading
`to uncoupling of the replication fork. Activated RecA* forms a filament
`on the damaged template and (c) ~40 minutes post-induction of SOS,
`the mutagenically active (UmuD¢ 2C) pol V is formed. The assembly of
`(UmuD¢ 2C) pol V on the 3¢ -OH vacated by pol III core at the site of the
`lesion is believed to be targeted by RecA*. (d) Pol V Mut, consisting of
`UmuD¢ 2C, RecA, b sliding clamp, g clamp loading complex, and SSB
`(not shown), subsequently catalyzes error-prone TLS past a 6–4 T–T
`photoproduct incorporating G preferentially at the 3¢ T leading to Tfi C
`transitions, consistent with genetic data. (e) Synthesis by pol V is
`distributive in the presence of RecA leading to its dissociation
`following the incorporation of only a few nucleotides beyond the lesion.
`Pol III core can then re-assemble on the primer terminus and resume
`replication of the remaining chromosome.
`
`TLS requires the presence of all of the above proteins
`[22••,24••]. We will refer to the UmuD¢ 2C, RecA, b pro-
`cessivity clamp, g
`clamp-loading complex, SSB protein
`
`combination by the term pol V Mut (mutasome), as origi-
`nally suggested by Harrison Echols [11]. Remarkably, RecA
`stimulates pol V activity by 14,000 fold, reflecting its essen-
`tial contribution to a functional pol V Mut [26••]. TLS by
`pol V occurs in the absence of b processivity clamp and g
`clamp-loading complex when non-hydrolyzable ATPg S
`replaces ATP in the reaction, suggesting that RecA filament
`disassembly eradicates pol V’s ability to copy past template
`damage sites. As pol V Mut and the pol III core compete for
`the same 3¢ -primer termini, pol-V-catalyzed TLS is actually
`inhibited in the presence of pol III core [24••]. Neverthe-
`less, pol III HE has a vital role to play in taking over from
`the distributive (i.e. rapidly dissociating) pol V, once the
`lesion is bypassed, to carry out processive replication on the
`next undamaged stretch of template. A model depicting
`pol V Mut catalyzed TLS is shown in Figure 2.
`
`Several important results have begun to emerge from in
`vitro assays using pol V Mut. A comparison of TLS using
`pols III, IV, and V revealed that only pol V Mut was able to
`catalyze efficient bypass of abasic, cis-syn T–T and 6-4
`T–T lesions [26••]. Another key issue concerns the speci-
`ficity of incorporation opposite each lesion. For example,
`in accordance with the in vivo mutation spectra, pol V
`favors incorporation of G opposite the 3¢ -T of the 6-4 pho-
`toproduct resulting in Tfi C transition mutations, whereas
`pols III and IV favor incorporation of A almost exclusively
`[26••]. The TLS and incorporation specificity data taken
`together for these three lesions suggest that pol V Mut is
`responsible for generating most, if not all, SOS mutations
`targeted at DNA damage sites.
`
`Pol V Mut also exhibits remarkably low fidelity when
`copying undamaged DNA, with error rates of about 10–3
`for most transition and transversion base mispairs [26••].
`This observation is consistent with the requirement for
`UmuD¢ and C in order to observe mutations in the absence
`of DNA damage in RecA730 cells with constitutive induc-
`tion of SOS [27]. The recently discovered pol IV (encoded
`by dinB) is also induced as part of the SOS regulon but its
`only known phenotype is in causing an increase in simple
`frameshift mutations on undamaged lambda phage DNA
`[28]. A deletion of the gene encoding pol IV (D dinB) has no
`measurable effect on either targeted on untargeted chro-
`mosomal mutations; however, an increase in F¢ episomal
`frameshift mutations accompanying the overproduction of
`pol IV [29] suggests that pol IV might also act on chromo-
`somal DNA. A recent in vitro study shows that pol IV is
`able to extend mismatched primer 3¢ -ends with unusually
`high efficiency [30••], a property also exhibited by pol V
`Mut [22••].
`
`Like pol V, pol IV can utilize the b processivity clamp and
`g clamp-loading complex resulting in a 3000 fold increase
`in pol IV activity [26••]. E. coli pol IV (DinB) and pol V
`(UmuD¢ 2C) share common sequence elements with two
`yeast polymerases, Rev1 and Rad30, and with their animal
`cell counterparts. These ‘parent’ enzymes make up a
`
`

`

`GDA201.QXD 03/22/2000 02:15 Page 165
`
`DNA polymerases involved in genome repair Goodman and Tippin 165
`
`superfamily of aberrant DNA polymerases that has little in
`common with the well-known polymerase families A
`(e.g. E. coli pol I and Bacteriophage T7 pol), B (e.g. E. coli
`pol II and eucaryotic pols a
`, d , e ), C (e.g. E. coli pol III
`subunit) and X (e.g. eucaryotic pol b ) involved in DNA
`replication and repair [31].
`
`Sequence domains of the UmuC/DinB/Rev1/
`Rad30 superfamily of DNA polymerases
`The UmuC/DinB/Rev1/Rad30 superfamily members
`(Figure 3) possess five highly conserved regions, domains
`I–V, presumed to be involved in binding and catalysis for
`template directed nucleotide incorporation. Site-directed
`mutations eliminating polymerase activity have been
`found in the most conserved residues in domain I
`(Asp8fi His) and domain II (Arg49fi Phe) in E. coli DinB,
`as well as mutations in domain III, including the double
`mutant Asp155fi Ala Glu156fi Ala of Rad30 from Saccha-
`romyces cerevisiae, and Asp103fi Asn in DinB and UmuC
`Asp101fi Asn from E. coli [24••,30••,32••]. It remains to be
`determined whether the substrate-binding or catalysis
`
`steps are affected. Two helix-hairpin-helix DNA-binding
`motifs are found in domains IV and V. A carboxy-termi-
`nal deletion of S. cerevisiae Rad30 resulting in truncation
`of domain V is also devoid of detectable polymerase
`activity [32••].
`
`The four major subgroups of this family can be
`distinguished by their unique domains. The Rev1 subfam-
`ily — so far found only in eucaryotes — has three distinct
`domains. The most amino-terminal is the BRCT (BRca1
`C-terminal) domain, found in many other eucaryotic
`enzymes such as XRCC1 (X-ray cross complementing 1)
`and the breast cancer gene BRCA-1, and is believed to
`mediate protein–protein interactions necessary to form
`coordinated complexes involved in both cell cycle check-
`points and DNA repair [33]. Rev1 proteins require
`association with Rev3–Rev7 (pol z ) in order to perform
`error-prone TLS by incorporating C opposite abasic sites
`[34]. Perhaps the unique carboxy-terminal end of the Rev1
`proteins has been acquired and conserved to mediate this
`specific interaction.
`
`Figure 3
`
`Alignment of some members of the UmuC/DinB/Rev1/Rad30
`superfamily. A schematic representation of the conserved and unique
`domains present in the UmuC/DinB/Rev1/Rad30 superfamily is
`shown. The highly conserved domains I–V containing probable
`catalytic residues that have been mutated in several studies and helix-
`hairpin-helix DNA-binding motifs are denoted above by Roman
`numerals. E. coli UmuC is the least conserved family member followed
`by the newly discovered human Rad30B, which shares the small extra
`region of homology (light blue) found in both the DinB and Rad30
`subgroups. UmuC and human Rad30B both have unique carboxy-
`terminal ends (thin black lines). The DinB subgroup shows remarkable
`
`conservation of three short motifs (shown in purple), which are
`present from E. coli to humans. The C2H2 and C2HC zinc binding
`motifs (shown as green and yellow diamonds respectively) are
`presumed to be involved in DNA binding and perhaps in selective
`targeting. The BRCT domain is shown (pink oval) at the amino-
`terminal end of the Rev1 subgroup. Conserved regions of unknown
`function are found in the amino (pink ovals) and carboxyl termini
`(peach squares) of human and C. elegans DinB. Additional motifs
`conserved within subgroups are indicated by arrows. Amino acid
`lengths are indicated in parenthesis. Ce (C. elegans), Ec (E. coli),
`h (human), Sc (S. cerevisiae).
`
`a
`

`

`GDA201.QXD 03/22/2000 02:15 Page 166
`
`166 Chromosomes and expression mechanisms
`
`Both the DinB and Rad30 subfamilies have acquired extra
`DNA-binding domains represented by zinc fingers or clus-
`ters, although each has done so in its own way. The Rad30
`members in yeast and humans have added the C2H2 zinc
`finger DNA binding motif while the DinB group in the
`higher eucaryotes (Caenorhabditis elegans and humans) have
`selected the C2HC type of zinc cluster ([35•]; B Tippin,
`MF Goodman, unpublished data). These acquisitions
`might increase DinB’s stability on DNA or to lend speci-
`ficity to the type of template and lesion to which it can
`bind. DinB in C. elegans and humans has also acquired two
`other unique amino- and carboxy-terminal motifs, the lat-
`ter containing putative nuclear localization sequences.
`Human Rad30B is an interesting case in that it shares a
`small extra region of homology between domains II and III
`with the DinB proteins but shows slightly more identity at
`the amino-acid level to yeast Rad30, and thus cannot be
`classified definitively into either subfamily [36••]. Like
`UmuC, human Rad30B has an extensive carboxy-terminal
`region that cannot be characterized by any known motifs,
`while maintaining the minimal domain conservation seen
`across the entire family.
`
`Similar folks, different strokes: a functional
`family of aberrant polymerases
`UmuD¢ 2C (pol V) and Rad30 (pol h ) homologs appear to
`have a similar biological role in replicating past template-
`damage sites, but with profoundly different biological
`consequences. E. coli strains with deletions of pol V are
`nonmutable by UV radiation and are incapable of perform-
`ing error-prone TLS, whereas yeast with deletion of the
`gene encoding pol h exhibit a slight increase in UV muta-
`bility [37] and are no longer able to catalyze error-free TLS
`[38]. Therefore, diametrically opposite TLS phenotypes
`are generated: error-prone with E. coli pol V Mut and error-
`free with yeast pol h
`. The human Rad30 homolog has been
`identified as the Xeroderma pigmentosum variant gene
`product, XPV pol h
`[39••], which by analogy with yeast
`pol h
`[32••] bypasses T–T cyclobutane dimers by correct-
`ly incorporating two A’s. Skin cancers occurring in the
`absence of pol h are caused, presumably, by errant copying
`of T–T dimers by the human homolog of yeast pol z
`(Rev3–Rev7) or perhaps using Rad30B.
`
`Nucleotide insertion fidelity appears to be governed by a
`geometrical selection principle in which polymerase active
`clefts are specifically designed to accommodate W–C base
`pairs [40,41]. The low fidelity of the Umu-like enzymes
`might result from a strategic trade-off in which active site
`geometrical constraints have been relaxed allowing TLS to
`occur at a cost of perhaps a 100-fold reduction in
`nucleotide insertion fidelity. Indeed, E.
`coli pol V
`[22••,24••] and yeast [42] and human pol h
`(L Prakash,
`personal communication) make excessive numbers of
`errors when copying undamaged DNA templates in vitro,
`with base substitution error rates of about 10–2 to 10–3. Sev-
`eral possible explanations may account for how the low
`fidelity pol h
`can catalyze error-free TLS. Assuming that
`
`most UV-induced T–T dimers undergo repair prior to
`replication, then the remaining few can still be copied
`accurately 95 to perhaps 99.5% of the time by errant pol h
`.
`It’s also possible that pol h
`is designed specifically to copy
`T–T cyclobutane dimers accurately. Compared with pol h
`,
`which appears targeted to a specific type of DNA damage
`site, E. coli pol V Mut can copy a variety of lesions [26••],
`and although pol V Mut is clearly not meant to replicate
`undamaged DNA, it nonetheless does so occasionally
`causing a marked increase in untargeted mutagenesis [27].
`
`In contrast to pol V Mut, the absence of an effect of dinB
`(pol IV) deletions on SOS lesion-targeted mutations
`implies that E. coli pol IV is not involved in TLS. Instead,
`pol IV might play a role in relieving replication forks
`stalled at misaligned primer–template structures in
`homopolymer runs [43], or at forks impeded perhaps by
`transiently slipped mispairs that are difficult either to
`extend or to proofread [27,44]. Such a role is consistent
`with the observation that frameshift mutation rates in
`phage lambda and on F¢ episomes are dependent on pol IV
`levels [28,29].
`
`Speculation on the biochemical basis of
`somatic hypermutation
`The discovery of the UmuC/DinB/Rev1/Rad30 superfam-
`ily of aberrant DNA polymerases may provide new
`impetus in addressing a vexing question in the field of
`immunology: what is the molecular basis of B cell somatic
`hypermutation? Hypermutation in the variable region of
`immunoglobulin genes is responsible, in part, for generat-
`ing an astounding diversity of antibodies. Yet, the
`molecular mechanisms governing this process remain a
`completely open question [45]. On the basis of studies
`using mouse B cells, at least two components are required
`to generate somatic hypermutation. An arbitrary promoter
`must be located immediately upstream from any arbitrary
`target gene, and at least one downstream V-gene specific
`enhancer element is also required, although two enhancers
`are somewhat more effective [46]. Mutations in the DNA
`target sequence are mostly transitions that are located
`proximal to the end of the promoter and diminish further
`downstream. The hypermutation rate is estimated to be
`about 10–3 per base pair [47], reminiscent of pol V Mut,
`pol IV [26••] and pol h
`[42]. Somatic hypermutation might
`therefore involve the action of an errant DNA polymerase
`targeted to the V-gene by a combination of enhancer and
`promoter binding proteins in a manner loosely analogous
`to RecA targeting of pol V to the site of a lesion [22••,24••].
`
`Conclusions and future directions
`The next stage in the study of the UmuC/DinB/Rev1/
`Rad30 superfamily of polymerases will be to elucidate with
`what proteins the polymerases interact and the mecha-
`nisms of selective targeting. Lesion bypass by pol V has
`been shown to require interaction with the b processivity
`factor of the replicative E. coli pol III HE [22••,24••], while
`RecA is thought to target pol V to lesion sites [13–15]. It is
`
`

`

`GDA201.QXD 03/22/2000 02:15 Page 167
`
`DNA polymerases involved in genome repair Goodman and Tippin 167
`
`not known, however, how the other homologs are targeted
`exclusively to sites of DNA damage nor whether or not
`they interact with replication forks or DNA repair com-
`plexes. Further biochemical studies using mutants of this
`novel family of polymerases should help answer these
`questions. We also anticipate that the recent discovery of
`this superfamily of ‘sloppier copier’ polymerases will
`reveal new insights into the mechanism of somatic hyper-
`mutation.
`
`Acknowledgements
`The comprehensive, standard-setting studies of Evelyn Witkin and
`Harrison Echols in SOS regulation and mutagenesis require special
`mention. We want to express our sincere gratitude to past and present
`collaborators Hatch Echols, Roger Woodgate, Mike O’Donnell, Ramon
`Eritja, John Petruska, Kevin McEntee and to the many graduate and
`postdoctoral students for their important contributions in elucidating
`mechanisms of DNA repair and mutagenesis. This work was supported by
`National Institutes of Health Grants GM42554, GM21422 and AG00093.
`
`References and recommended reading
`Papers of particular interest, published within the annual period of review,
`have been highlighted as:
`• of special interest
`•• of outstanding interest
`Kornberg A, Baker TA: DNA Replication, edn 2. New York:
`WH Freeman and Company; 1992:113-196.
`
`1.
`
`2.
`
`3.
`
`DeLucia P, Cairns J: Isolation of sn E. coli strain with a mutation
`affecting DNA polymerase. Nature 1969, 224:1164-1166.
`
`Rangarajan S, Woodgate R, Goodman MF: A phenotype for
`enigmatic DNA polymerase II: a pivotal role for pol II in replication
`restart in UV-irradiated Escherichia coli. Proc Natl Acad Sci USA
`1999, 96:9224-9229.
`
`4. Walker GC: Inducible DNA repair systems. Annu Rev Biochem
`1985, 54:425-457.
`
`5.
`
`6.
`
`7.
`
`8.
`
`9.
`
`Kato T, Shinoura Y: Isolation and characterization of mutants of
`Escherichia coli deficient in induction of mutagenesis by
`ultraviolet light. Mol Gen Genet 1977, 156:121-131.
`
`Steinborn G: Uvm mutants of Escherichia coli K12 deficient in UV
`mutagenesis. I. Isolation of uvm mutants and their phenotypical
`characterization in DNA repair and mutagenesis. Mol Gen Genet
`1978, 165:87-93.
`
`Sommer S, Knezevic J, Bailone A, Devoret R: Induction of only one
`SOS operon, umuDC, is required for SOS mutagenesis in E. coli.
`Mol Gen Genet 1993, 239:137-144.
`
`Shinagawa H, Iwasaki T, Kato T, Nakata A: RecA protein-dependent
`cleavage of UmuD protein and SOS mutagenesis. Proc Natl Acad
`Sci USA 1988, 85:1806-1810.
`
`Burckhardt SE, Woodgate R, Scheuremann RH, Echols H: UmuD
`mutagenesis protein of Escherichia coli: overproduction,
`purification, and cleavage by RecA. Proc Natl Acad Sci USA 1988,
`85:1811-1815.
`
`10. Nohmi T, Battista JR, Dodson LA, Walker GC: RecA-mediated
`cleavage activates UmuD for mutagenesis: mechanistic
`relationship between transcriptional derepression and
`posttranslational activation. Proc Natl Acad Sci USA 1988,
`85:1816-1820.
`
`11. Echols H, Goodman MF: Mutation induced by DNA damage: a
`many protein affair. Mutation Res 1990, 236:301-311.
`
`12. Dutreix M, Moreau PL, Bailone A, Galibert F, Battista JR, Walker GC,
`Devoret R: New recA mutations that dissociate the various RecA
`protein activities in Escherichia coli provide evidence for an
`additional role for RecA protein in UV mutagenesis. J Bacteriol
`1989, 171:2413-2415.
`
`13. Sweasy JB, Witkin EM, Sinha N, Roegner-Maniscalco V: RecA
`protein of Escherichia coli has a third essential role in SOS
`mutator activity. J Bacteriol 1990, 172:3030-3036.
`
`14. Bailone A, Sommer S, Knezevic J, Dutreix M, Devoret R: A RecA
`protein mutant deficient in its interaction with the UmuDC
`complex. Biochemie 1991, 73:479-484.
`
`15. Frank EG, Hauser J, Levine AS, Woodgate R: Targeting of the
`UmuD, UmuD¢¢ , and MucA¢¢ mutagenesis proteins to DNA by RecA
`protein. Proc Natl Acad Sci USA 199

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket