DNA 复制与修复研究进展

  • Upload
    cale

  • View
    118

  • Download
    11

Embed Size (px)

DESCRIPTION

DNA 复制与修复研究进展. 2002 年 10 月 14 日. DNA 复制:一个古老和新奇的问题 DNA 复制的保守性和多样性 DNA 修复: DNA 稳定性的重要保证. DNA 复制与修复研究进展专题. DNA 复制起点的结构和序列的保守性 参与原核生物 DNA 复制的酶类和蛋白质 参与真核生物 DNA 复制的酶类和蛋白质 细菌 DNA 复制的机制和调控 病毒 DNA 复制的机制和调控 RNA 病毒的 RNA 复制机制和调控 真核生物 DNA 复制的机制和调控 原核生物 DNA 的修复机制 真核生物 DNA 的修复机制 DNA 修复与疾病. - PowerPoint PPT Presentation

Citation preview

  • DNA20021014

  • DNADNADNADNA

  • DNADNADNADNADNADNARNARNADNADNADNADNA

  • The chromosome replication cycle

    Journal of Cell Science 115, 869-872 (2002) John F. X. Diffley1,* and Karim Labib2

    1 Cancer Research UK, Lincoln's Inn Fields and Clare Hall Institute, Blanche Lane, South Mimms, EN6 3LD, UK 2 Cancer Research UK, Christie Hospital NHS Trust, Wilmslow Road, Withington, Manchester, M20 4BX, UK

  • tubulinDNA

  • DNA

  • 2

  • TLS

  • GP45

  • Step 1. Replication origins are determined, at least in part, by the binding of the six subunit origin recognition complex (ORC), which has been conserved in evolution from yeast to humans. In budding yeast, ORC binds specifically to an essential, bipartite sequence element within replication origins. ORC binds ATP and this binding is required for DNA binding. ORC can also hydrolyse ATP; however, ATP hydrolysis is not required for DNA binding and is, in fact, inhibited upon origin binding. ORC remains bound to origins throughout the cell cycle in budding yeast, even through mitosis. In multicellular eukaryotes some ORC subunits remain bound throughout G1, S and G2 phases, although there is conflicting evidence as to whether ORC remains bound to chromosomes during mitosis.

    Figure 1. Metaphase-to-anaphase transition in rat kangaroo PtK2 cells. -Tubulin, green; centrosomes, red; DNA stained with 4',6-diamidino-2-phenylindole, blue.

    Figure 2. Structure of cohesin and a possible mechanism by which it might hold sister chromatids together. (A) Smc1 (red) and Smc3 (blue) form intramolecular antiparallel coiled coils, which are organized by hinge or junction domains (triangles). Smc1/3 heterodimers are formed through heterotypic interactions between the Smc1 and Smc3 junction domains. The COOH terminus of Scc1 (green) binds to Smc1's ABC-like ATPase head, whereas its NH2 terminus binds to Smc3's head, creating a closed ring. Scc3 (yellow) binds to Scc1's COOH-terminal half and does not make any direct stable contact with the Smc1/3 heterodimer. Scc1's separase cleavage sites are marked by arrows. Cleavage at either site is sufficient to destroy cohesion. By analogy with bacterial SMC proteins, it is expected that ATP binds both the Smc1 and Smc3 heads, alters their conformation, and possibly brings them into close proximity. By altering Scc1's association with Smc heads, ATP binding and/or hydrolysis could have a role in opening and/or closing cohesin's ring. (B) Cohesin could hold sister DNA molecules together by trapping them both within the same ring. Cleavage of Scc1 by separase would open the ring, destroy coentrapment of sister DNAs, and cause dissociation of cohesin from chromatin. (C) Smc-containing complexes other than cohesin could also function via chromatid entrapment. Condensin, for example (black), could organize mitotic chromosomes by trapping supercoils. It and/or other related complexes could hold distant loci together (arrow) and thereby facilitate the function of long-range enhancers and silencers of transcription. Figure 3. The anaphase-promoting complex induces amphitelically attached chromatids to segregate to opposite poles by destroying both cyclin B and securin. The Ipl1/Aurora B kinase both eliminates syntelically attached chromatid pairs and promotes inhibition of APC/C when centromeres fail to come under tension. Chromatin, blue; microtubules, green; centromeres, open circles; cohesin, red.

    Figure 4. Chromatid individualization as cells enter mitosis involves dissociation of most cohesin (red symbols) from chromosomes, which is regulated by PLK. This process coincides with condensin's (green symbols) association with chromosomes and their compaction. Cohesin remaining on chromosomes, largely at centromeres, is then cleaved by separase at the metaphase-to-anaphase transition.

    Figure 1. Summary of S.cerevisiae DNA damage, replication, and mitotic checkpoints. The different stages of the cell cycle are indicated above the horizontal line. Below the line are listed a subset of the proteins that function in the indicated checkpoint branches. These proteins are thought to detect the "damage" that triggers each checkpoint. The primary effect of activating each checkpoint is shown below the proteins. Listed below the three checkpoint branches that function in S-phase are many of the known proteins that function in the downstream signal transduction cascade or are targeted by this cascade. The box at the right lists human homologs of yeast checkpoint genes that are mutated in human cancer susceptibility syndromes.

    Figure 2. Illustration of break-induced replication, telomere maintenance, and de novo telomere addition reactions. (A) Break-induced replication. In this example, the 3' end of a broken DNA invades the intact DNA to form a D loop. The resulting heteroduplex region is then extended and a Holliday junction is formed. By appropriate resolution of the Holliday junction, a structure is formed that is equivalent to a replication fork. (B) Telomere maintenance functions. The diagram shows the end of a chromosome containing a single-stranded TG repeat end and the known S.cerevisiae proteins that bind to telomeres and extend the TG repeat sequences. (C) De novo telomere addition reactions. The diagram shows the end of a broken DNA molecule and the proteins that have been implicated in de novo telomere addition.

    Figure 3. Multiple pathways function to maintain genome stability in S.cerevisiae. DNA replication errors activate S-phase checkpoint sensors, whereas telomere damage likely activates DNA damage checkpoint sensors. These sensors then activate downstream responses, including DNA repair, that are required for suppression of genome instability and possibly cell cycle arrest or delay. Multiple pathways work to correct the damage; the primary nonmutagenic repair pathway appears to be break-induced replication (BIR), although double-strand break (DSB) repair and nonhomologous end joining (NHEJ) may play minor roles. The major mutagenic pathways are the translocation pathway(s), the de novo telomere addition pathway suppressed by PIF1, and aberrant recombination reactions that are suppressed by DNA mismatch repair (MMR) and other functions. The number of translocation pathways and whether there are pathways that specifically suppress translocations remain to be determined. Figure 1. Replication region of bacteriophage genome and assembly of the replication complex. Genes, promoters and terminators present in the replication region are indicated. Transcripts are represented by arrows. The scheme is not drawn to scale. See text for details.

    Figure 2. Two pathways of plasmid replication. Large circles represent plasmid DNA molecules. Small filled (orange) circles indicate heritable replication complexes. See text for details.

    Figure 3. Effects of various factors on formation and stability of the heritable replication complex. The complex is symbolized by large orange circle. Positive regulators are marked in green, and negative regulators are marked in blue. Stimulation processes are shown as arrows and inhibitory actions are shown by blunt-ended lines. See text for details.

    Figure 1. Model for TLS by specialized polymerases. High-fidelity semiconservative replication is arrested at sites of base damage or structural distortions in the replication fork (solid triangle). Multiple specialized polymerases are able to support TLS across offending template lesions. During TLS, the correct (error-free) (N) or the incorrect (error-prone) (M) nucleotide is inserted, depending on the nucleotide preference of the engaged specialized polymerase. In either case, persistence of the damaged template base precludes the resumption of normal DNA replication until the primer strand is extended for some distance (bold red), often by a second specialized polymerase, such as prokaryotic pol V and eukaryotic pol . The replicative machinery then reengages to continue high-fidelity synthesis. Once the site of base damage is cleared, it is subject to normal DNA repair. If TLS was error-free, the DNA is restored to the normal sequence (N:N). However, if TLS resulted in nucleotide misincorporation, a mutation will be fixed (M:M). The product of TLS is in red. Fig. 1. A, schematic showing the primers and template DNAs prepared for this study. The two methods of thymine oxidation are indicated as well as the position and sequence context of the lesion. B, literature values for the putative stereochemical make-up of each thymine glycol sample. Note that in the TgOsO4 template, the 5R stereoisomers are in greater relative abundance.

    Fig. 2. GST-pol bypasses thymine glycol in template TgBr2. Radiolabeled primer-templates (5nM) were incubated with all four deoxynucleotide triphosphates (100M total) and the indicated quantities of each polymerase and the resulting primer extension products resolved by DPAGE. GST-pol largely bypasses thymine glycol as does Klenow fragment (exo ) of E.coli polymerase I.In contrast, the calf thymus pol is arrested. Lanes 1, 3, 6, and 8, control experiments with either control or TgBr2 template but no enzyme added. Lane 2, undamaged template with 5nM GST-pol ; lanes 4and 5, TgBr2 template with 5and 10nM GST-pol ; lane 7undamaged template with 1nM Klenow (exo ); lane 9, TgBr2 template with 1nM Klenow (exo ); lanes 10-12, undamaged template with 0.1unit of calf thymus pol and increasing concentrations of PCNA; lanes 13and 14, TgBr2 with 0.4and 0.8units of calf thymus pol .

    Fig. 3. A, GST-pol exhibits identical bypass of an alternatively prepared thymine glycol template, TgOsO4, which is constitutionally identical to TgBr2 (Fig. 2) experiments, but stereochemically contains a greater proportion of 5R thymine glycol stereoisomers. Radiolabeled primer-templates (5nM) were incubated with all four deoxynucleotide triphosphates (100M total) at the indicated quantities of each polymerase, and the resulting primer extension products were resolved by DPAGE. Lanes 1and 4, control experiments with no enzyme added. Lanes 2and 3, undamaged template with 1and 5nM GST-pol . Lanes 5and 6, TgOsO4 template with 1and 5nM GST-pol . B, the TLS pattern for Klenow fragment of E.coli Pol I (exo ) with TgOsO4, is identical to that shown for TgBr2 in Fig. 2, and T4 DNA polymerase shows characteristic replicative arrest at Tg in TgOsO4. Experiments were performed analogously to those in A.Lanes 1, 3, 5, and 7, control experiments with no enzyme added. Lane 2, undamaged template with 1nM Klenow (exo ). Lane 4, TgOsO4 template with 1nM Klenow (exo ). Lane 6, undamaged template with 5nM T4 DNA polymerase. Lane 8, TgOsO4 template with 5nM T4 DNA polymerase. C, standing start experiments (primer P5-ox-ss) indicate nearly identical bypass of the two Tg templates by GST-pol . Experiments were performed analogously to those in A.Lanes 1and 4, control experiments with no enzyme added; lane 2, TgBr2 template with 1nM GST-pol ; lane 3, TgBr2 template with 5nM GST-pol ; lane 5, TgOsO4 template with 1nM GST-pol ; lane 6TgOsO4 template with 5nM GST-pol ; lane 7, undamaged template with 5nM GST-pol added.

    Fig. 4. GST-pol preferentially incorporates adenine across from thymine glycol. Radiolabeled primer-templates (5nM) were incubated with the indicated deoxynucleotide triphosphate(s) (100M total [dNTP]) and GST-pol (5nM) and the resulting primer extension products were resolved by DPAGE. Lanes 1-6 contained undamaged template DNA; lanes 7-12 contained TgBr2 template. Lanes 1and 7, control experiments with no enzyme added. Lanes 2and 8, only dATP; lanes 3and 9, only dCTP; lanes 4and 10, only dGTP; lanes 5and 11, only TTP. Lanes 6and 12contained a mixture of all four dNTPs.

    Fig. 5. Representative results for steady state kinetics analysis for incorporation of the four dNTPs opposite thymine glycol by GST-pol . The nucleotide incorporated is shown immediately above each gel, and the micromolar dNTP concentration incubated in each reaction is shown immediately below the corresponding lane of each gel. Experiments were performed using the undamaged template (A), the TgBr2 template (B), or the TgOsO4 template (C). Raw data were analyzed as described under "Experimental Procedures," and the resulting steady state kinetic parameters are reported in Table I. Note that most of the panels shown represent initial conditions with a broad dNTP range. dCTP and TTP gels in B show representative results from more narrowly focused working ranges used in subsequent experiments.

    Fig. 6. Graphical representation of select kinetic parameters from Table I. A, kcat/Km values for nucleotide insertion opposite thymine glycol in the TgOsO4 template (black bars), the TgBr2 template (blue bars), and the undamaged template (yellow bars). The left graph shows directly plotted kcat/Km values obtained for incorporation of the bases A, C, G, or T, emphasizing the difference in incorporation efficiency of A between the control and Tg templates. The right graph is a y axis blown up version of the same graph to emphasize that A is incorporated in great preference to the other bases in all three of the templates and that GST-pol may exhibit stereochemical preference during insertion of A opposite of thymine glycols. B, comparison of the finc values obtained for each template, measuring the degree of preference for incorporation of the correct base. finc is defined as (kcat/Km)incorrect/(kcat/Km)correct. The left graph is a direct plot of the finc values for nucleotide insertion opposite thymine glycol in the TgOsO4 template (black bars), the TgBr2 template (blue bars), and the undamaged template (yellow bars), emphasizing the high degree of preference for A incorporation in each template. The graph on the right is a y axis blown up version of the same graph to emphasize that the misincorporation frequency of G increases in template TgOsO4, suggesting that thymine glycol stereochemistry influences the ability of GST-pol to discriminate between the purines during incorporation opposite the lesion.

    Fig. 7. Schematic drawing of the most important conformers expected from the most abundant stereoisomers of thymine glycol. The 5R,6S form (left drawings, putatively in greater abundance in TgOsO4) may give rise to G-T wobble base-pairing in the GST-pol active site, resulting in higher levels of misincorporation of G.The 5S,6R isomer (right drawings) may adopt a much different conformation. dR, deoxyribose. As a result of different half-chair conformations, the stereoisomers may each present the hydrogen bonding surface of the base at different angles, giving rise to the observed differences for misincorporation of G between the TgBr2 and TgOsO4 templates.

    Figure 1. Model for the sequential assembly of the various components of nucleotide excision repair. Global genome repair involves the initial binding of the XPC-hHR23B and XPE binding proteins followed by downloading of the TFIIH transcription and helicase complex that remodels the damaged site. Transcription coupled repair involves initial response to damage by stalling of the RNA polymerase II apparatus, and coupling by the CSA and CSB proteins. Subsequent steps proceed in common, consisting of loading the XPG nuclease, the XPA-RPA DNA binding proteins, and the ERCC1-XPF nuclease. After nuclease cleavage around the dimer site, the excision complex departs and the site is resynthesized by PCNA-polymerase delta and ligase I. Figure has been redrawn (37) with addition of the transcription repair pathway.

    Figure 2. Causative mutations reported in the XPD gene after elimination of non-functional mutations. The DNA/DNA helicase boxes are shown by clear boxes with the amino acid involved indicated by numbers above the gene. DNA/RNA helicase boxes shown in lighter shade below the gene with roman numerals for numbering. The mutations are indicated by the number and the type of amino acid change, with those causing XPD or XPD/CS above, and those causing XPD/TTD below. The interaction domain between XPD and the p44 component of the TFIIH complex is shown. Reproduced from (117).

    Figure 3. The mechanism of tumor formation in basal cell nevus syndrome involving the hedgehog (Hh), patched (PTC) and smoothened (SMO) signal transduction pathway. Top left: normal signal transduction pathway. PTC inhibits SMO and no nuclear transcription signal is induced. Top right: normal pathway after sonic Hh binds to PTC, removes binding to SMO, and results in signal transduction to nuclear transcription factor Gli that induces cPTC over-expression. Bottom: three mechanisms of derangement of the signal transduction pathways in tumors involving either PTC, SHh or SMO. Redrawn from initial diagrams provided by M. Aszterbaum MD, PhD.

    Figure1.Schematic illustration of mechanisms and barriers in liposome-mediated gene delivery to hepatocytes. Potential barriers are highlighted with dark background. Lipoplexes formed from specific ligand-labeled cationic liposomes and plasmid DNA are administered through either peripheral vein or portal vein. In both administration routes, the lipoplexes may interact with plasma proteins, which results in large aggregates. The aggregates may be large enough to be trapped in the lung circulation and only small part of injected lipoplexes will reach the liver when the lipoplexes are administrated intravenously. Lipoplexes are endocytosed by hepatocytes when they have been recognized by the cells. Endocytosed lipoplex components including plasmid DNA may undergo degradation by catalytic enzymes in lysosomes or endosomes, and only a small portion of DNA is able to enter the nucleus through the nuclear pore complex. Episomal DNA in the nucleus will be transcribed, and transgene expression can be detected in cell lysates. ASGP-R = asialoglycoprotein receptor.

    Figure 1. Structure of genetically dissected replication origins in eukaryotes. Consensus ORC binding sites are indicated in red. Additional sequences important for origin activity are shown in brown. Transcription units and regulatory sequences are shown in green. Sites of ORC binding, where known, are indicated. Sites of initiation of replication are indicated with a bidirectional arrow passing through a bubble.

    Figure 2. Once-per-cell-cycle genome duplication is independent of the positions or density of replication origins. (A) Duplication of DNA exactly once-per-cell division is achieved with two mutually exclusive periods of the cell cycle during which either pre-RCs can be assembled but replication cannot initiate or replication can initiate but pre-RCs cannot be assembled. Regardless of where pre-RCs assemble, DNA is completely replicated and cannot be re-replicated until after cell division. (B) The assembly of extraneous pre-RCs can ensure that the entire segment of DNA is replicated in a timely fashion, without the need for specific DNA sequences to optimize origin spacing. Any pre-RCs that do not initiate are destroyed by passage of the replication fork and so cannot re-initiate on the replicated strands.

    Figure 3. Exclusion of pre-RCs from specific regions could create the need for origin focusing mechanisms. In this model, passage of the transcription apparatus before replication depletes transcription units (black boxes) of pre-RCs (yellow stars). (A) Without transcription, there is no selective pressure to focus initiation to specific sites as the assembly of pre-RCs at many sites ensures timely genome replication. (B) When transcription units are sparse, the assembly of multiple pre-RCs at many sites within the intergenic regions is sufficient to accomplish genome replication. (C) When intergenic regions are few and far between, specific DNA sequence recognition elements are required to ensure that at least one pre-RC is assembled at intervals appropriate to accomplish the timely replication of the DNA segment.

    Fig. 1. Pre-RCs are formed and activated on immobilized linear DNA fragments. Equivalent amounts of magnetic beads (lane 1) or magnetic beads coupled to a 3-kb DNA fragment of pBluescript (lanes 2-6) were incubated with egg cytosol for 20min, isolated, washed, and analyzed by Western blotting. In lane 4, the egg cytosol contained geminin. In lanes 5and 6, the incubation in egg cytosol was followed by 15min of incubation in aphidicolin-supplemented NPE that contained (lane 6) or lacked (lane 5) 1M p27Kip. Western blots were probed with MCM3 (all panels), ORC2 (left and center panels), or Cdc45 (right panel).

    Fig. 2. ORC and MCM are efficiently recruited to an 82-bp DNA fragment. A and B, a 251-bp PCR product spanning the pBluescript polylinker was coupled to beads. The beads were divided into equal-sized pools and mock-digested or digested with different enzymes to generate 251(lane 1), 154(lane 2), 120(lane 3), 94(lane 4), and 67(lane 5)-bp DNA fragments. Lane 6contained beads lacking DNA. After digestion, DNA attached to beads was analyzed on a 2% agarose gel (A), or equal quantities of the various beads were incubated with egg cytosol and the attached proteins analyzed by Western blotting (B). C, a similar set of DNA beads as those used in A were incubated with egg cytosol containing buffer (lanes 1-5) or geminin (lanes 6-10), and the bound proteins were analyzed by Western blotting. D, proteins bound to equivalent amounts of 94-bp beads (lane 1) and 251-bp beads (lane 2) were analyzed by Western blotting alongside 2-fold dilutions of a 1:1 molar ratio of purified MCM3 and ORC2 (lanes 3-6). The amount of ORC2 and MCM3 in lane 3is 16and 24.6ng, respectively. E, binding of proteins to digested DNA beads as in B but with the addition of 82-(lane 3) and 54-bp (lane 5) fragments. As in B, all lanes contained equivalent amounts of beads that were all derived by digestion from the 251-bp parental DNA fragment.

    Fig. 3. MCM complex binding is proportional to the length of DNA fragments attached to beads, whereas ORC binding is independent of DNA fragment length. A, a 6-kb PCR product was coupled to magnetic beads and digested to generate 3-,1-,and 0.35-kb DNA fragments. Binding of MCM3 and ORC2 to equal amounts of digested beads was examined in egg cytosol in the presence (lanes 5-8) and absence (lanes 1-4) of geminin. B, an immobilized 6-kb DNA fragment (lane 1) or sperm (lane 2) was incubated in egg cytosol, and binding of MCM3 and ORC2 was examined. C, a 1-kb PCR product was attached to beads and digested to generate a 100-bp DNA fragment. Equivalent amounts of the 1-kb DNA beads (lanes 1and 4), 100-bp DNA beads (lanes 2and 5), and beads lacking DNA (lane 3) were incubated in egg cytosol (EC) containing (lanes 4and 5) or lacking (lanes 1-3) geminin, and the attached proteins were analyzed by Western blotting. D, proteins bound to an immobilized 1-kb DNA fragment (lane 4) were analyzed on a Western blot alongside known quantities of purified MCM3 and ORC2. Lanes 1-3, 0.5ng of ORC2; lane 1, 15.3ng of MCM3; lane 2, 7.7ng of MCM3; lane 3, 1.5ng of MCM3.

    Fig. 4. A, lateral MCM complexes are bound tightly to chromatin. 100-bp DNA beads derived by digestion from 251-bp DNA beads (lanes 1-4) or 6-kb DNA beads prepared separately (lanes 5-8) were incubated with egg cytosol, diluted with buffer containing Triton X-100 and increasing concentrations of salt as indicated, isolated, washed in the same buffer, and probed with MCM3, MCM7, or ORC2 antibodies. B, sperm chromatin was incubated with egg cytosol, diluted with buffer containing Triton X-100 and 0.1M KCl, isolated, washed in buffer containing 0.1(lane 1) or 1M (lane 2) KCl, and probed with ORC2 antibody (lower panel) or a mixture of MCM3 and MCM7 antibodies (upper panel). C, Cdc45 binding is independent of DNA length. Equal quantities of 6-kb DNA beads (lanes 1, 5, and 9), 1-kb DNA beads (lanes 2, 6, and 10), or 100-bp DNA beads derived from the 1-kb DNA beads by digestion (lanes 3, 7, and 11), as well as beads lacking DNA (4, 8, 12), were incubated with egg cytosol (EC) (lanes 1-4), egg cytosol followed by NPE (lanes 5-8), or egg cytosol followed by NPE containing 1M p27Kip (lanes 9-12). The NPE contained aphidicolin. Chromatin-bound proteins were analyzed by Western blotting using MCM3 antibody (upper panel) or a mixture of ORC2 and Cdc45 antibodies (lower panel).

    Fig. 5. Unlike binding of MCM2-7, chromatin binding of Cdc45 is rate-limiting for DNA replication. A, sperm chromatin (10,000/l) was incubated with egg cytosol containing (lane 3) or lacking (lanes 1, 2, and 4) geminin and then supplemented with 2volumes of NPE containing 50g/ml aphidicolin (lanes 2-4) and p27Kip (lane 4only). Immediately before (lane 1), or 20min after the addition of NPE (lanes 2-4), sperm chromatin was isolated, and the equivalent of 10,000sperm was analyzed on a Western blot using MCM3 antibodies (top panel) or a mixture of ORC2 and Cdc45 antibodies (bottom panel). The Western blot also contained purified proteins (lanes 5-12). Lane 5, 0.4ng of ORC2, 0.6ng of MCM3; lane 6, 0.4ng of ORC2, 1.8ng of MCM3: lane 7, 0.4ng of ORC2, 6ng of MCM3; lane 8, 0.4ng of ORC2, 18ng of MCM3; lane 9, 0.4ng of Cdc45; lane 10, 1.2ng of Cdc45; lane 11, 4ng of Cdc45; lane 12, 12ng of Cdc45. B, sperm chromatin was incubated with 6l of Cdc45-depleted egg cytosol and then supplemented with 12l of total of mock-depleted and Cdc45-depleted NPE mixed in the following ratios: 1:0 (triangles), 1:4 (circles), 1:9 (diamonds), 1:19 (squares), 0:1 (filled squares). 9l of each sample was mixed with [ -32P]dATP and replication measured 20,40,60,and 80min after NPE addition (graph). The other 9l of each sample was mixed with aphidicolin, and chromatin binding of MCM, Cdc45, and RPA was measured 25min after NPE addition. C, mock-depleted and MCM7-depleted egg cytosol were mixed in 1:0 (lane 1), 1:1 (lane 2), 1:3 (lane 3), 1:7 (lane 5), and 1:15 (lane 6) ratios and incubated with sperm chromatin (10,000/l). Immediately before (top 3panels) or 15min after addition of NPE (bottom panel), sperm chromatin was isolated, and bound proteins were analyzed by Western blotting with antibodies against MCM3, ORC2, or Cdc45. An aliquot of each reaction was supplemented with [ -32P]dATP and replication measured 60min after addition of NPE (bar graph).

    Fig. 6. All chromatin-bound MCM complexes are phosphorylated in a Cdc7-dependent fashion. Xenopus sperm chromatin was incubated with egg cytosol (10,000/l) for 30min and then supplemented with 2volumes of NPE. Aliquots containing 40,000sperm were withdrawn immediately before (lane 1) or at the indicated times after NPE addition (lanes 2-7), isolated, and the bound proteins analyzed by Western blotting using antibodies against MCM4 and MCM7 (top panel) or ORC2 (bottom panel). Lane 6contained the same sample as lane 1, and lane 9contained a 30:1 mixture of MCM3 (22.5ng) to ORC2 (0.5ng). Top panel, probed with MCM3 serum; bottom panel, probed with ORC2 serum.

    Fig. 7. Actinomycin D stimulates Cdc45 binding. A, sperm chromatin was incubated with egg cytosol and then supplemented with NPE containing buffer (lane 1), 10g/ml actinomycin D (lane 2), or aphidicolin (lane 3). After 20min, chromatin was isolated and probed with MCM3 and Cdc45 antibodies. B, sperm chromatin was incubated with egg cytosol containing (lane 4) or lacking (lanes 1-3 and 5) geminin. Subsequently, NPE containing 10g/ml actinomycin D (lanes 2, 4, and 5), aphidicolin (lane 3), and p27Kip (lane 5) was added. After 20min, chromatin was isolated and probed by Western blotting with antibodies against MCM3 and Cdc45 alongside 10ng of purified MCM3 and 0.66,2.2,and 6.6ng of purified his-Cdc45 (lanes 6-8).

    Fig. 8. Model for the initiation of DNA replication. A, ORC, Cdc6, and Cdt1 stimulate MCM2-7 binding to sites widely distributed around ORC. B, at the G1/S transition, Cdc7 binds to and phosphorylates many MCM complexes. C, Cdk2/cyclin E stimulates the association of Cdc45 with a subset of these MCM complexes. Activation of the first MCM complexes by Cdc45 may lead to inactivation of neighboring MCM complexes, thereby restricting initiation to defined intervals.

    Figure 1. A) Schematic illustration of mature HSV-1 virion showing major structural components. B) Diagram of HSV genome. The unique long and short segments (UL, US) are shown flanked by repeat sequences. Genes encoded are shown adjacent to their loci in the schematic genome; those above the schematic are not essential for replication in tissue culture. Gene names are color-coded according to their functions: teal = regulatory; blue = capsid and DNA packaging; red = envelope; orange = tegument; green = DNA replication, and black = uncertain function.

    Figure 2. Flow chart illustrating the cascade of regulatory events that results in ordered sequential expression of genes during HSV-1 lytic infection. As ICP4 and ICP27 are both absolutely required for gene expression to proceed to the E or L stages, viral replication can be blocked by deleting one or other of these essential IE genes.

    Figure 3. Schematic summary of the in vivo life cycle of HSV-1. An epithelial surface, with its innervating sensory neuron is shown. The nucleus of the neuron has been enlarged to depict the intranuclear events occurring during lytic and latent infection.

    Figure 4. Schematic summary of the vectors discussed in the text. The name of each virus is shown to the left of the schematic; the viruses used in the studies reviewed here are referred to by name throughout the text. The diagrammatic genomic map of each vector is aligned with that of the HSV-1 genome in Figure 4A to facilitate comparison between viruses. Each schematic depicts the positions and types of foreign transgenes inserted into each construct, and which subset of viral genes has been inactivated.

    Figure 5. Schematic summary of neurological applications for HSV-1 vectors. For each region of the nervous system, the possible disease applications are listed with types and specific examples of transgenes whose delivery and expression may be beneficial. Data pertaining to many of these are discussed in the text.

    Figure 6. Some applications of HSV vectors in the peripheral nervous system. A) The cell bodies and central terminals of transduced afferent nerve fibers are depicted schematically. Vector-mediated expression of pre-proenkephalin in these cells results in appropriate processing, transport, and packaging, such that enkephalin is found at central presynaptic terminals. Several models indicate that painful stimuli cause enkephalin release, resulting in the inhibition of pain neurotransmission. Nerve growth factor (NGF) expression and release can also be demonstrated in these cells following transduction with an appropriate vector; the site of action of this agent is uncertain at present. B) Injection of the rat footpad with formalin results in pain behavior that may be scored. Following an initial transient nociceptive response, pain behavior reappears and lasts for approximately 1 hour, reflecting a more chronic pain mechanism. Pretreatment of rats with a pre-proenkephalin expressing vector reduces the chronic pain behavior in this model without affecting the initial nociceptive response. The asterisks (*) denote results that show a statistically significant difference between the experimental (SHPE) and control (SHZ). C) Dorsal root ganglion cells may be transduced with a replication-defective vector expressing NGF, which protects the cells of the ganglion against the toxicity of a hydrogen peroxide challenge. The neuroprotective effect is most pronounced at 3 days post-transduction when using a construct in which a viral immediate-early promoter drives NGF expression (SHN). HSV LAP2-driven transgene expression increases through day 14; neuroprotection is maximal at this later time point (SLN). The protective effect is similar in magnitude to that seen when recombinant NGF is applied to the culture, and is not seen when a promoterless construct (SN) is used. Stable, chronic expression from the LAP2 promoter may enable long-term therapeutic transgene expression in sensory neurons; in other studies, reporter gene expression has been detected over a year after the initial transduction event.

    Figure 7. Strategies for increasing bystander lysis in antiglioma suicide gene therapy. In HSV transduced cells, the pro-drug GCV is activated by viral thymidine kinase to form an oncolytic monophosphate derivative. The active drug may enter and kill surrounding cells; this may be enhanced by coexpression of Cx43 to encourage gap junction formation between transduced cells and their neighbors. In addition, TNF- secretion from transduced cells makes the tumor more sensitive to radiation. Both of these enhancements to TK-GCV suicide gene therapy are effective in promoting the survival of animals in an in vivo model of malignant glioma. The Kaplan-Meier survival curves on the right side of the figure show the survival data from nude mice that have been intracerebrally inoculated with U87 glioma cells and then treated with the vectors and drugs shown in the chart legends.

    Figure 8. Non-neurological applications of HSV vectors. We are currently studying ex vivo transduction of stem cells for the introduction of therapeutic genes or differentiation factors. In vivo transduction of muscle may allow delivery of dystrophin or other gene products that are absent in various muscle diseases. "Depot" tissues, such as adipose and ligament, may be transduced with genes encoding circulating proteins, with a variety of potential applications.

    Figure 1. A simplified model of the bacterial cell cycle. DNA (dark gray lines), origin (oriC, gray circles), terminus (terC, dark gray square), replisome (overlapping triangles, one for each replication fork), cytokinetic ring (dashed line). In this model, DNA replication initiates at or near mid cell. The sister origins rapidly separate from each other and become anchored on opposite halves of the cell. DNA replication continues followed closely by refolding of newly replicated DNA until there are two complete and separate chromosomes. Finally, the cell divides medially. The model is simplified to ignore multifork replication.

    Figure 2. A model of a circular chromosome that is undergoing multifork replication in a rod-shaped bacterium. The cell contains four copies of the origin region (gray circles). The initial round of replication in the middle is duplicating the parental chromosome (thin black) to generate the gray copies. A second round has begun at the quarters, duplicating the daughters (gray) to generate the granddaughters (thin black). The cell contains four copies of the origin region (gray circles), located approximately at the one-, three-, five-, and seven-eighth positions along the cell length, and a single copy of the terminus region (dark gray square) located near mid cell.

    Figure 3. The extrusion-capture model for bacterial chromosome partitioning. After the origin region is replicated, the two sister origins (light gray circles) are extruded from (arrows pointing toward cell poles) the centrally located replisome (overlapping triangles) and captured on opposite halves of the cell at or near the cell quarters. The terminus region (dark gray square) remains at mid cell until it is duplicated.

    Figure 4. Terminus-specific chromosome partitioning events. (A) Chromosome decatenation. (B) If formed, a chromosome dimer is resolved to two monomers by two site-specific recombinases acting on the dif site in the terminus region. (C) When chromosomes (gray) are trapped in an invaginating division septum (black vertical lines), SpoIIIE (black ovals) pumps the chromosomes out of the way (arrows above cell indicate direction each chromosome will move).

    Fig. 1. DNA replication vs. homologous recombination. Chromosomes are shown as double lines. Parental strands are filled; daughter strands are open. (A) A chromosome. (B) Chromosome replication has been initiated. (C) Chromosome replication is nearing completion. (D) Chromosome replication is complete. (E) Strand degradation in preparation for homologous recombination has started. (F) Strand degradation is nearing completion, whereas annealing of the complementary strands is going on.

    Fig. 2. The pathways of replication fork stalling/disintegration with subsequent resetting/repair. DNA duplexes are shown as double lines; a protein tightly bound to DNA is shown as a bricked circle. For all Holliday junctions, one of the two possible resolution directions is indicated by the small arrows. (A) A replication fork. (B) The replication fork approaching a single-strand interruption in template DNA. (C) The replication fork has collapsed at the interruption. (D) Double-strand end invasion to restore the replication fork structure. (E) A stalled replication fork. (F) Regression of the stalled replication fork forms a double-strand end and a Holliday junction. (G) Double-strand end invasion to restore the replication fork structure. Resolution of the Holliday junction in F leads to replication fork breakage (C). Resolution of the Holliday junctions in D or G, or exonucleolytic degradation of the linear tail in F leads to restoration of the replication fork structure (A).

    Fig. 3. Two ways to repair a replication fork by single-strand annealing. DNA duplexes are shown as double lines; sister chromatid cohesion is indicated by thin dumbbells. (A) A replication fork approaching a single-strand interruption in template DNA. (B) The replication fork has collapsed. (C) Rad52-promoted reannealing of the detached end with the complementary single-strand gap on the full-length chromatid. (D) Repair of the single-strand interruption. (E) Filling-in the single-strand gap on the full-length chromatid. Sister chromatid alignment is shown. The thin arrows indicate a hypothetical signal from the Rad52-bound double-strand end to the intact sister chromatid. (F) Rad54-catalyzed unwinding of the intact chromatid in the vicinity of the double-strand end. (G) Rad52-catalyzed annealing of the double-strand end with the open sister duplex to generate a replication fork structure.

    Fig. 1. Three systems required for the precise duplication of chromosomal DNA replication. Top panel, a small segment of DNA carrying three replication origins passing through the cell cycle. Red boxes denote ORC and a plus denotes a licensed origin. The metaphase chromosome decondenses, assembles licensed origins, is assembled into a nucleus, replicates, and then recondenses. Below are the activities of the origin recognition system, the licensing system and SPF during the course of the cell cycle. See text for further details.

    Fig. 2. Events occurring at a replication origin in the Xenopus early embryo from late mitosis to late G1. (A)Assembly of ORC onto origin DNA. (B)Binding of Cdc6 and RLF-B/Cdt1 onto ORC. (C)Multiple Mcm(27) heterohexamers are assembled onto each origin to license it. (D)Once licensing is complete Cdc6 is removed, whilst ORC binds less tightly (unshaded ORC denotes weak binding to DNA). (E)Cdc7 is recruited to the licensed origin and phosphorylates Mcm(27) complexes (purple circles denote phosphorylation). (F)Following nuclear assembly, CDK activity induces the assembly of Cdc45 onto the origin.

    Fig. 1. Model of SOS translesion replication by DNA polymerase V.The two DNA strands are shown as green lines, and the replication-blocking lesion is represented by the red rectangle. The three major steps in TLR are pre-initiation (2), in which the RecA nucleoprotein filaments assembles; initiation (3and 4), which involves binding of pol V to the primer-template and loading of the subunit clamp; and lesion bypass by pol V holoenzyme (5). SSB is suggested to help in displacing RecA from DNA both at the initiation and lesion bypass steps.

    Fig. 1. Pathways for recombinational DNA repair of a stalled replication fork. A pathway involving fork regression is shown for gap repair (a-f), and a double-strand break repair path is shown for the repair of a fork collapsed at the site of a DNA strand break (g-l). The pathways shown are intended to be generic and do not incorporate all current ideas for fork repair. The dashed line represents a pathway in which the Holliday junction is converted to a double-strand break by the action of RuvABC, as observed by Michel and colleagues (180) and others. Arrowheads on DNA strands denote 3' ends. If a DNA lesion (other than a strand break) is responsible for halting the progress of a replication fork, note that replication fork repair does not entail repair of the lesion itself. Instead, the recombination and replication steps set up the lesion for repair by providing an undamaged complementary DNA strand. The degree to which excision repair and other DNA repair processes are integrated with the recombinational repair pathways is unknown.

    Fig. 2. A replication intermediate from Drosophila embryos, photographed by Inman (49). One of the apparently regressed forks has a single-stranded tail (arrow). (Reprinted from Biochim. Biophys. Acta, 783,Inman, R.B., "Methodology for the study of the effect of drugs on development and DNA replication in Drosophila melanogaster embryonic tissue," pp. 205-215, Copyright 1984,with permission from Excerpta Medica, Inc.; ref. 49.)

    Figure 1. Models for replication restart by recombination. (Left) Holliday junction (HJ) formation by fork regression allows bypass of unreplicatable DNA damage. (Right) A stalled fork can be restored after breakage by D-loop formation and strand invasion.

    Fig. 1. Recombination repair of broken replication forks. (A) Rescue of blocked replication forks (adapted from ref. 17). The replication fork is blocked at the Ter site in the presence of Tus. A DSB occurs in the lagging-strand template. The RecBCD enzyme enters at the double-strand end and initiates homologous recombination catalyzed by RecA. Completion of the recombination reaction by resolution of the Holliday junction leads to restoration of a replication fork. Binding of the primosome allows loading of a new replisome to promote DNA replication restart. To account for the viability of a strain carrying an ectopic Ter site in a recombination-proficient background, one needs to assume that the newly reconstituted replication fork is not arrested again, and hence that Tus has been removed from Ter during the recombination reaction. DSB on the lagging strand is shown, but a similar model can apply to breakage and repair of the leading strand. (B) Replication fork collapse (adapted from ref. 31). The progressing fork encounters a single-strand interruption in the leading-strand template, because of a defect in closure of the lagging strand at the previous replication round. Reincorporation of the broken DNA strand by homologous recombination and replication restart are catalyzed by the same enzymes as on breakage of the fork. The full lines represent the two DNA strands, the dashed lines represents newly synthesized DNA strands. The arrowhead corresponds to DNA 3' ends.

    Fig. 2. RuvAB/RecBCD-mediated rescue of blocked replication forks (adapted from ref. 44). In the first step (A) the replication fork is blocked and the two newly synthesized strands anneal, forming a Holliday junction (see Fig. 3 for the different pathways proposed to promote this step). In a second step (B) the junction is stabilized by RuvAB binding. (C) In recombination proficient strains, RecBCD binds to the double-strand tail (C1); degradation takes place until the first recognized CHI site (CHI is an octameric sequence that switches RecBCD from an exonuclease to a recombinase enzyme) and is followed by a genetic exchange mediated by RecA (C2); RuvC resolves the first Holliday junction bound by RuvAB (C3). In C2 and C3, the double-strand end is reincorporated into the circular chromosome by homologous recombination and the Holliday junction is resolved, which results in the reconstitution of a replication fork. This pathway is presumably used in recombination-proficient cells. (D) RecBCD-mediated degradation of the tail progresses up to the RuvAB-bound Holliday junction. Replication can restart when RecBCD has displaced the RuvAB complex. D can take place in recombination-proficient strains if RecBCD reaches RuvAB before encountering a CHI site; it is the only pathway that leads to a viable chromosome in recA and ruvC mutants. (E) RuvC resolves the RuvAB-bound Holliday junction. This pathway is used in the absence of RecBCD and leads to the RuvABC-dependent DSBs observed in recBC mutants. Continuous and discontinuous lines represent the template and the newly synthesized strand of the chromosome respectively; the arrowheads indicate the 3'end of the growing strands.

    Fig. 3. Models for formation of Holliday junctions at arrested replication forks by RFR. (A) RecA binds to the single-stranded region of the lagging-strand template, polymerizing in the 5' to 3' direction. Pairing of the lagging-strand template with the leading-strand template renders the leading strand free to anneal with the 5' end of the lagging strand. This results in the formation of a Holliday junction that can be bound directly by RuvAB. (Adapted from ref. 64.) (B) RecG binds to the replication fork and migrates toward the chromosomal replication origins, displacing the 5' end of the lagging strand. RecG activity ultimately creates a four-stranded junction. (Adapted with modifications from ref. 76.) (C) (+) Topological stress that accumulates downstream of the fork on arrest is relaxed by unwinding of the two newly synthesized strands from the template strands and their annealing. (Schematic representation based on results in ref. 78.) Full and dashed lines represent the template and the newly synthesized DNA, respectively; the arrowheads indicate the 3' end of the growing strands.

    Fig. 4. RFR in UV-irradiated cells (adapted from refs. 76, 80, and 92). The replication fork is blocked by a UV photo-product (black triangle) in the leading-strand template. RFR, proposed to be catalyzed by RecG in E. coli (76), or by Rad51 (the yeast RecA homologue) in S. pombe (80), renders the damaged DNA double stranded and thereby allows direct repair by nucleotide excision repair enzymes (A). If the lagging-strand polymerase has continued synthesis past the lesion, leading-strand DNA synthesis using the lagging strand as template followed by reverse branch migration [proposed to be catalyzed by RecG in E. coli (76) and by Rqh in S. pombe (80)] reconstitutes a fork on which the lesion has been bypassed (B; ref. 94). Full and dashed lines represent the template and the newly synthesized DNA, respectively; the arrowhead indicates the 3' end of the growing strand.

    Fig. 1. A scheme to illustrate stalled replication fork rebuilding by replication fork regression, by using recombination enzymes in the E. coli chromosome (adapted from refs. 17, 18, 24, and 25). Pathways that lead to a single crossover (or an odd number of crossovers) generate dimeric chromosomes, and those that act without crossing over (or even numbers of crossovers) retain the monomeric status of the chromosome. Pathways considered to be major routes to retaining the monomeric chromosome status are overlaid onto a beige background. Similarly, the arrows bounded by a bold line are intended to indicate major pathways, with relative contributions being indicated by arrow breadth. Those arrows in dark green indicate pathways that would be expected to be RecABCD-dependent, whereas those in light green indicate RuvC cleavage from within a RuvABC complex. Arrows in light blue indicate RuvAB helicase action, whereas that in pink indicates similar action by RecG or RecQ helicases. Black lines are unreplicated DNA, and red/pink lines newly replicated daughter strands. (a) Reannealing of daughter strands at a stalled replication fork (closed triangle indicates a nontemplate lesion on the leading strand). Reannealing can be mediated by RecG (18) and facilitated by positive supercoiling ahead of the fork (27). RuvAB and RecA may also promote the growth of the reversed fork once fork reversal has been initiated. Lagging strand synthesis is indicated as proceeding beyond the lesion, thereby providing an opportunity to replicate past the lesion by copying of the switched template in a reaction that uses recombination proteins, but not recombination (pink line; ref. 18). (b) Productive RuvAB branch migration to extend the four-way HJ (RuvB is cartooned as a pair of cylinders on opposed arms of the HJ). Note that RuvB binding to the other two arms of the junction (step l) will lead to abortion of the four-way junction by branch migration (step p). (c and m) Action of RuvABC to cleave the strands 3' of the bound RuvB on the branch point side, to generate broken forks (corresponding to single strand lesions in the leading and lagging parental template strands respectively; ref. 24). RecABCD-mediated reinvasion of the broken ends leads to rebuilt replication forks that most readily yield noncrossover (d-f) and crossover (n) chromosomes, respectively. Note that after reinvasion, a further round of RuvABC action is required; in each case the "productive" orientation of RuvB binding gives the majority species shown (f and n). Binding of RuvB in the abortive configuration either will act to reverse the invasion or will lead to RuvC cleavage to give the minority products, dimers and monomers, respectively (k, o). (g-i). RecABCD-mediated invasion of the end created by fork reversal into its homologous region to generate a molecule containing two HJs (or a HJ and a three-way junction). Such an intermediate can be processed by crossover (i) or noncrossover (h) pathways by several ways that involve the simultaneous or sequential action of proteins at each of the branch points; we predict dimers to predominate over monomers in these events. (j) Processing of the reversed replication fork intermediate by RecG or RecQ helicases (equivalent to step p, promoted by RuvAB).

    Fig. 2. Alignment of E. coli DnaX, YcaJ, and RuvB (CLUSTAL X multiple sequence alignment program, v. 1.8,http://www.ebi.ac.uk/clustalw/). The sequences of YcaJ, RuvB, and DnaX all contain well-conserved nucleotide-binding sites with Walker A (GxxxxGKT/S) and Walker B (Dexx) motifs. The Zn-binding motif of DnaX is absent in YcaJ, but the putative ATPase sensor motifs (29) are present. Colors represent types of amino acids.

    Fig. 3. (A) An outline of the Xer recombination reaction. XerCD bind cooperatively at dif, psi, or cer recombination sites, ensuring synapsis (with the help of accessory sequences and proteins in the case of psi/cer) (i). XerC initiates catalysis (ii) to form a HJ intermediate, which undergoes a conformational change (iii) to provide a substrate for catalysis by XerD, which can then complete the recombination reaction (iv). There is normally a barrier to this conformational change, and XerC frequently catalyzes the conversion of the HJ back to substrate (ii). In recombination at psi, the proteins PepA and ArgA-P facilitate the HJ conformational change, whereas in recombination at dif, FtsKc is thought to facilitate this change (34, 35). (B, C) Species specificity of FtsK action. FtsK cells (DS9041) were transformed with pBAD expression vectors (48) carrying full-length FtsK proteins (B) or the C-terminal domains (C) of different species. To assay for Xer recombination, they were transformed with a plasmid containing two dif sites and grown in conditions of repression ( ; 0.2% glucose) or induction (+; 0.2% arabinose) of the expression vectors (34). Induction was checked by Western blot analysis by using an antibody directed against a FLAG epitope fused to the N termini of the constructs, after resolution of the protein extracts on a 6% (B) or an 8% (C) SDS/PAGE. (D) Alignment of the C-terminal domains of FtsK homologues. Identical residues are indicated by stars, conservative substitutions by dots. Open boxes underline regions predicted to adopt an -helix conformation by PREDICTPROTEIN PHD software v. 1.96,http://www.embl-heidelberg.de/predictprotein/predictprotein.html whereas black arrows underline those predicted to form sheets. Ec: E. coli FtsK, Hi: H. influenzae FtsK and Bs: B. subtilis SpoIIIE.

    Fig. 4. The C-terminal domain of FtsK is randomly distributed throughout the cytoplasm. FtsK cells (DS9041) were transformed with pBAD expression vectors carrying an N-terminal fusion of the green fluorescent protein (GFP) to full-length FtsK or to the C-terminal domain of FtsK (FtsKc). Cells were grown to midexponential phase in LB supplemented with 0.1% (full-length) or 0.2% (FtsKc) arabinose. Nucleoids were stained by using 4',6-diamidino-2-phenylindole (DAPI). Phase-contrast and fluorescent images were acquired by using a cooled charge-coupled device camera (Princeton Instruments, Trenton, NJ) and METAMORPH IMAGE ACQUISITION software (Universal Imaging, Media, PA) from an Olympus BX50 (New Hyde Park, NY) fluorescence microscope. Shown are overlays of the DAPI image in red and the GFP image in green. Full-length FtsK frequently localizes to the septum, whereas FtsKc is always distributed throughout the cytosol and never found at the septum.

    Fig. 5. A model for chromosome segregation in E. coli. In contrast to the replication origins (large dark gray circles), the replication terminus region, which contains the dif site (open triangle), stays localized at midcell (49). As a consequence, sister dif sites can be synapsed by the XerCD recombinases (small white and light gray circles) whether the chromosomes form a dimer or not. This leads to cycles of HJ formation and resolution by XerC. If the chromosomes are monomeric, segregation will eventually break the synaptic complex and move the sites away from midcell before septum closure. If the chromosomes form a dimer, the synaptic complex will stay trapped at midcell, which allows access to FtsK. FtsK mediates the HJ conformation change needed to activate catalysis by XerD, thus coordinating resolution of chromosome dimers to cell division.

    Fig. 1. Conversion of plectonemic supercoils into knot and catenane nodes by Int site-specific recombination. (A) A ( ) supercoiled Int substrate (black line) is shown with the att recombination sites represented by red and blue arrows. When the att sites are in inverse (head-to-head) orientation in the primary sequence, the recombination products are right-handed torus knots; these knots can be drawn without crossings on the surface of a torus- or doughnut-shaped object (Upper). When the sites are directly (head-to-tail) repeated, the product is a right-handed torus catenane (Lower). (B) An electronmicrograph of a 13-noded Int knot produced in vitro. The DNA was coated with RecA protein to help visualize the crossings. (Reprinted from Cell 276,Spergler, S.J., Stasiak, A.&Cozzarelli, N.R., "The stereostructure of knots and catenanes produced by phage lambda integrative recombination: 1985,Implications for mechanism and DNA structure," 325-334, 1985,with permission from Elsevier Science; ref. 15.)

    Fig. 2. Monte Carlo simulations of catenanes between relaxed and supercoiled DNA molecules. Simulation of a singly linked catenane between two relaxed plasmids (A) or between a relaxed and supercoiled plasmid (B). The yellow chain represents a 3.5-kb DNA, and the red chain a 7-kb DNA. Simulations courtesy of Alexander Vologodskii.

    Fig. 3. Conformations of replicating DNA. (A) A replication fork is depicted at Left with the parental strands in black and the daughter strands in red. Red arrows denote 3' ends. During replication, the denaturation of the parental duplex causes a (+) Lk in the replicating molecule (Center). This (+) Lk can be expressed either as (+) supercoiling of the parental duplex in front of the replication fork or (+) precatenanes between the replicated duplexes behind the fork. We have shown the (+) Lk in the usual fashion, which assumes that it is initially ahead of the fork and must diffuse past the replisome to generate precatenanes. It is, however, possible that the converse is true and that (+) precatenanes are the primary consequence of the (+) Lk from replication. Upon replisome dissociation (Right), the ends of the nascent strands will be free to base pair with each other, forming a four-way junction at the replication fork, the chickenfoot, that allows the replication fork to regress until the molecule is relaxed. (B) Electron micrograph of an in vitro replication intermediate with replication stalled by the Tus/ter complex. The molecule displays both supercoils in the unreplicated region (thick line) and precatenanes in the replicated region (thin line). (Reprinted from ref. 47, with permission from Elsevier Science.) (C) Scanning force microscopy of an in vivo replication intermediate incubated in ethidium bromide. This molecule displays the linear duplexes of the middle toe of the chickenfoot (white arrows) emerging from both the unidirectional origin of replication and the terminus. (Reprinted from ref. 60, with permission from the American Society for Biochemistry and Molecular Biology.) (For B and C, the scale bar is 100nm.)

    Fig. 4. Physiological implications of the chickenfoot. Replication forks are shown with parental strands in black and daughter strands in red, and with the 3' end tipped with an arrowhead. (A) Recombination-mediated replication restart. The four-way junction of the chickenfoot can be cleaved by Holliday junction resolving enzymes such as RuvC (white arrowheads) [1]. Cleavage will sever one of the replicated arms, which can then be processed by the RecBCD complex (Pac-man) [2], allowing it to become a substrate for homologous recombination [3]. Recombination with the sister replicated arm re-forms the replication fork, allowing replication restart [4]. (B) Bypass of a lesion. An unpaired lesion (black rectangle) on the parental strand blocks leading strand replication, but lagging strand replication can continue for more than 500nt [1]. Upon replisome dissociation, the two nascent strands can base pair, requiring the denaturation of more than 500bp at the lagging strand [2]. Once a chickenfoot is formed, the lagging strand becomes a template for the leading strand, allowing leading strand replication to continue past the lesion [3]. Reabsorption of the chickenfoot re-forms the replication fork, allowing replication to continue and a second opportunity to repair the lesion [4].

    Fig. 5. Model for topology of the replicating chromosome. A segment of chromosomal DNA is depicted, with black lines as parental strands and red lines as nascent strands. In the bacterial chromosome, domain barriers (yellow boxes) isolate the topology around the fork from the rest of the chromosome, which is ( ) supercoiled by DNA gyrase, as shown in the domains on either side of the replication domain. Replication creates a (+) Lk in the replicating domain (Center), which can cause (+) supercoils ahead of the fork and (+) precatenanes behind it. Thus, either gyrase or topo IV could support replication by removing (+) supercoils in front of the fork, and topo IV could also support replication by removing precatenanes behind the replication fork.

    Fig. 1. Resolution of Holliday junctions in early replication intermediates can generate a number of different products. (i) The ERI. (ii) An ERI in which one end (the origin-proximal end) of the nascent DNA has regressed. This could happen with equal probability at the other end as well. (iii) Cleavage of the Holliday junction in ii generates an structure. (iv) An ERI in which both ends of the nascent DNA have regressed. Because resolution of each Holliday junction can occur in one of two ways, two sets of products are generated: a nicked circle (form II) and a short duplex DNA corresponding to the distance on the replicated portion of the template that is between the sites of resolution (which will vary depending on the amount of nascent DNA that has regressed) (v); and a linear molecule that is longer than the original template by the distance on the template that is between the sites of resolution (vi). Red, the leading-strand template; blue, the lagging-strand template; green, nascent DNA.

    Fig. 2. Both RusA and RuvC can cleave ERIs. Standard oriC replication reactions containing either RusA (lanes 2-5, concentration varied by a factor of 10in each lane, left to right), RuvC (lanes 7-10, concentration varied by a factor of 10in each lane, left to right), or having no addition (lanes 1and 6) were incubated at 37C for 5min. The reactions then were terminated by the addition of EDTA, and the DNA products then were analyzed by electrophoresis through either a neutral 1% agarose gel (A) or a denaturing 1% alkaline agarose gel (B) as described in Materials and Methods. , structure; FII, form II (nicked, circular DNA). The mobility difference between the ERI and the structure is subtle and can be seen more clearly in Figs. 5-7.

    Fig. 3. RuvA inhibits cleavage of the ERI by RusA. Standard oriC replication reactions containing either 10nM RusA (lane 2), 10nM RusA and RuvA (lanes 3-5, RuvA concentration varied by a factor of 10in each lane from left to right), 100nM RuvA (lane 6), or having no addition (lane 1) were incubated first for 2min at 37C after RuvA addition and then for an additional 5min at 37C after RusA addition. The reactions were terminated by the addition of EDTA, and the DNA products then were analyzed by electrophoresis through a neutral 1% agarose gel.

    Fig. 4. Holliday junctions form in inactivated ERI. Inactivated ERI was formed as follows. Nine standard reactions for ERI formation as described in Materials and Methods were incubated at 37C for 10min. The reactions then were terminated by heating to 65C for 5min. After cooling, the reactions were pooled together. Nine and one-half microliters of the inactivated ERI pool was incubated in new reaction mixtures (10l) containing either no RuvA or RusA (lane 1), RusA (lanes 2-7, concentration in lanes 2-4 varied by a factor of 10from left to right and was 100nM in lanes 5-7), and RuvA (lanes 5-8, concentration varied by a factor of 10in lanes 5-7 from left to right and was 500nM in lane 8) and incubated for 2min at 37C after RuvA addition and for an additional 5min at 37C after RusA addition. The reactions then were terminated by the addition of EDTA, and the DNA products were analyzed by neutral 1% agarose gel electrophoresis. F III', form III' (see text for definition); F III, form III (full length, linear form).

    Fig. 5. Negative supercoiling inhibits RusA cleavage of the ERI. Inactivated ERI was formed as in the legend to Fig. 4 except that 12reaction mixtures were pooled. Fresh ATP was added to this pool to a concentration of 2mM. Nine microliters of the inactivated ERI pool was incubated in new reaction mixtures (10l) containing either no RusA (lanes 1,7,and 12) or 100nM RusA (all other lanes), and either Topo IV (lanes 3-7, concentration varied by a factor of 5in lanes 3to 6from left to right and was 25nM in lane 7) or DNA gyrase (lanes 8-12, concentration varied by a factor of 5in lanes 7-11 from left to right and was 25nM in lane 12) for 5min at 37C. The reactions were terminated by the addition of EDTA, and the DNA products were analyzed by neutral 1% agarose gel electrophoresis. The autoradiogram of the gel is shown in A and a photograph of the ethidium bromide stain gel is shown in B. FI, negatively supercoiled template DNA; F I', form I' (intact template DNA with no supercoils); SC, supercoiled ERI.

    Fig. 6. RecG stimulates RusA-catalyzed cleavage of the ERI. Inactivated ERI was formed as in the legend to Fig. 4 except that six reaction mixtures were pooled. Fresh ATP was added to this pool to a concentration of 2mM. Six and three-tenths microliters of the inactivated ERI pool was then incubated in new reaction mixtures (7l) containing either no RecG or RusA (lane 1), RecG (lanes 2-4 and 6-8, concentration varied by a factor of 10in lanes 2-4 and 6-8, left to right), and 10nM RusA (lanes 5-8) for 5min at 37C. The reactions were terminated by the addition of EDTA, and the DNA products were analyzed by neutral 1% agarose gel electrophoresis.

    Fig. 7. RecG promotes nascent strand regression in negatively supercoiled ERI. Inactivated ERI was formed as in the legend to Fig. 4 except that 12reaction mixtures were pooled. Fresh ATP was added to this pool to a concentration of 2mM. Twelve and three-quarter microliters of the inactivated ERI pool was then incubated in new reaction mixtures (15l) containing DNA gyrase (25nM), RecG (10nM), and RusA (100nM) as indicated for 5min at 37C. The reactions were then terminated by the addition of EDTA. One-half of the reaction mixture was analyzed by neutral 1% agarose gel electrophoresis [autoradiogram (A); photograph of the ethidium bromide-stained gel (C)] and the other half was analyzed by denaturing 1% alkaline agarose gel electrophoresis [autoradiogram (B)]. The arrow on the right-hand side of B marks a faint band that may correspond to the linear fragment formed by cleavage of the ERI at both ends of the nascent DNA (see Fig. 1v).

    Fig. 1. Replication of Mu by transposition. In the first stage, the phage-encoded transposition proteins aided by the histone-like protein HU promote transfer of 3'-OH ends of miniMu (red) to each strand of target DNA (green). Two sites, 5bp apart on target DNA, that will be subjected to a nucleophilic attack by each Mu end are indicated by arrows. Strand exchange produces a fork at each Mu end, the target providing 3'-OH ends (indicated by half arrows) that can potentially serve a primers for leading strand synthesis. MuA transposase, which has been assembled into an oligomeric transpososome, remains tightly bound to both Mu ends in the strand exchange product (strand transfer complex, STC1). Host factors then initiate Mu DNA synthesis from one end to duplicate Mu and form the final cointegrate product. The DNA synthesis phase was initially reconstituted in an eight-protein system supplemented with partially purified host factors (MRF), as described in the text.

    Fig. 2. Components of the MRF. MRF was originally identified as host factors needed in addition to the eight-protein system to convert STC1 to cointegrates. Resolution of MRF into enzyme fractions distinguishable by function and into pure components (ClpX, PriA, PriB, and DnaT) is indicated.

    Fig. 3. Transition from transpososome to replisome. The molecular chaperone ClpX converts STC1 (A) to STC2 (B), altering the conformation of the transpososome. MRF 2 then displaces the transpososome to assemble the prereplisome at the Mu forks, forming STC3 (C). PriA binds to the forked DNA structure created by strand exchange (D) and begins the process of assembling a replisome at one Mu end. The mechanism that determines which Mu end is used to initiate DNA synthesis is not yet clear. PriA assembles a preprimosome complex by recruiting PriB, DnaT, and the DnaB-DnaC complex (E). In this process, DnaB must be bound to single-stranded lagging strand template. To create this binding site, PriA unwinds duplex DNA by translocating 3' to 5' along this template. Once bound to DNA, DnaB attracts primase to form a primosome, which catalyzes primer synthesis for lagging strand synthesis, and DnaB promotes binding of the DNA pol III holoenzyme to complete replisome assembly.

    Fig. 4. Fragile property of the STC2 transpososome. Formation of STC1, its conversion to STC2, crosslinking of the transpososome with DSS, and detection of crosslinked MuA by Western blot analysis was conducted as previously described (63). Lane 1: As control, the strand exchange reaction mixture was incubated without HU protein, conditions which do not permit transpososome assembly, and then MuA was subjected to crosslinking with DSS. Lanes 2-6: STC1 was isolated free of unbound proteins by filtration through a Bio-Gel (Bio-Rad) A-15 m column equilibrated with 25mM Hepes-KOH (pH 7.5), 12mM magnesium acetate, and 60mM KCl. Isolated complexes were incubated at 37C for 30min in the presence of ATP, ClpX being included for conversion to STC2. The reaction mixture was adjusted to 300or 500mM NaCl, as indicated, and allowed to stand at room temperature for 15min before addition of DSS. For lane 2,STC1 was not subjected to DSS treatment as control.

    Fig. 5. MRF 2 consists of at least two distinct components. The reconstituted Mu DNA replication reaction (50l) with [ -32P]dNTPs was assembled with ClpX, the 12-protein system, and the indicated MRF 2 components, and products were resolved by alkaline agarose electrophoresis as previously described (75). Crude MRF (fraction II) (30) and MRF 2 (fraction III) (63) were prepared as described. Resolution of MRF 2 into two components, MRF 2A and MRF 2B, will be described in a future publication (V.D. and H.N., unpublished work). Approximately 10units (63) of the indicated MRF 2 components were added. Where indicated, MRF 2A and MRF 2B were heated at 65and 100C, respectively, for 10min. In the reaction catalyzed with crude MRF, greater than 95% of STC1 was converted to cointegrate. When MRF 2 was supplied as two components, typically 50-95% of STC1 was converted to a cointegrate. CO, position of the cointegrate; STC, position of the strand exchange product (not radiolabeled and therefore not visible).

    Fig. 1. Alternative BIR mechanisms. A broken chromosome end will be resected by 5' to 3' exonucleases, allowing the 3' end to interact with various recombination proteins to carry out strand invasion. (A) The 3' end of the invading strand initiates DNA replication, leading to a migrating D-loop "bubble" as described by Formosa and Alberts (5). the displaced newly synthesized DNA strand can then be made double-stranded. (B) Strand invasion sets up a replication fork that will result in semiconservatively synthesized molecules. A Holliday junction will be resolved at some point. (C) Strand invasion sets up a replication fork in which branch migration enzymes displace both newly synthesized DNA strands as the replication structure migrates down the template.

    Fig. 2. BIR-dependent formation of LEU2 recombinants. (A) EcoRI-digested plasmid pWYL37 has homology only at one end to sites in the yeast genome. (B) The "LEU" segment at one end of the DSB may initiate new DNA synthesis, but the completion of the event requires that the newly synthesized DNA is displaced from the template and must rejoin to the other end of the DSB by a nonhomologous end-joining event. (C) A