Fact-checked by Grok 2 weeks ago

DNA replication

DNA replication is the biological process by which a double-stranded DNA molecule is copied to produce two identical DNA molecules, ensuring that each daughter cell receives a complete set of genetic information prior to cell division. This semi-conservative mechanism, first experimentally demonstrated in bacteria, results in each new DNA double helix containing one parental (template) strand and one newly synthesized complementary strand, allowing for high-fidelity duplication through base-pairing rules ( with , with ). The process occurs during the of the eukaryotic and is orchestrated by a multiprotein complex at specialized replication origins, achieving remarkable speed and accuracy to copy billions of base pairs with an error rate as low as one in 10^9 . In eukaryotes, replication initiates at multiple origins of replication—sequences rich in adenine-thymine base pairs—where the enzyme unwinds the DNA double helix, creating a Y-shaped replication fork and generating single-stranded templates stabilized by single-strand binding proteins. then synthesizes short RNA primers to provide a 3'-OH group for DNA polymerases to begin synthesis, which proceeds exclusively in the 5' to 3' direction; the leading strand is synthesized continuously toward the fork, while the lagging strand is formed discontinuously in short (100–200 nucleotides long) away from the fork. DNA polymerases (such as alpha, , and in eukaryotes) extend these strands by adding deoxyribonucleotides, with exonuclease activity correcting mismatches during synthesis, and topoisomerases relieving torsional stress ahead of the fork. Completion of replication involves the enzyme removing RNA primers and filling gaps with DNA, followed by sealing nicks to form continuous strands; replication termination occurs when converging forks meet, and in eukaryotes with linear , extends telomeres to counteract the end-replication problem at chromosome ends. is further enhanced by post-replicative mismatch repair systems that scan for and correct errors, reducing the overall mutation rate dramatically. Prokaryotes, like , employ a similar core mechanism but with a single origin and enzymes such as III, enabling faster replication suited to their simpler genomes. Disruptions in this process, such as enzyme deficiencies, can lead to genomic instability, mutations, and diseases including cancer.

Fundamentals of DNA Replication

DNA Structure and Topology

The double-helix structure of DNA, proposed by James D. Watson and Francis H. C. Crick in 1953, consists of two antiparallel right-handed helical strands composed of deoxyribonucleotide subunits, with the sugar-phosphate backbones forming the outer rails and the nitrogenous bases projecting inward. The strands are stabilized by specific hydrogen bonding between complementary base pairs—adenine (A) pairing with thymine (T) via two hydrogen bonds, and guanine (G) pairing with cytosine (C) via three—ensuring faithful transmission of genetic information during replication. This configuration creates a uniform diameter of approximately 2 nm and a helical pitch of 3.4 nm with 10 base pairs per turn, while the asymmetric positioning of the glycosidic bonds results in major and minor grooves along the helix; the major groove is wider (about 1.2 nm) and shallower, allowing proteins to access the base edges for sequence-specific recognition without disrupting the double helix. The topological properties of DNA, particularly supercoiling, arise from the (Lk), defined as the number of times one strand crosses the other in a , which remains unless broken; underwinding (negative supercoiling) or overwinding (positive supercoiling) introduces torsional that compacts the or hinders processes like unwinding. In prokaryotes, the is organized as a single circular , as visualized by John Cairns in 1963 using autoradiography of Escherichia coli DNA, which revealed a theta-shaped structure during replication and confirmed the circular , with the chromosome spanning about 700–900 μm when linearized. During replication, unwinding generates positive supercoils ahead of the fork, which topoisomerases relieve by transiently breaking and rejoining strands; type I topoisomerases, first identified by James C. Wang in 1971 as the E. coli ω protein, relax supercoils through single-strand nicks, while type II enzymes handle catenanes in circular genomes. The semiconservative nature of replication—each parental strand serving as a template for a new complementary strand—was experimentally verified by and Franklin W. Stahl in 1958 using density-labeled E. coli DNA, which showed hybrid density after one generation and segregated densities thereafter. In eukaryotes, chromosomes are linear, presenting distinct topological challenges: the ends, capped by telomeres, cannot be fully replicated by conventional DNA polymerases due to the requirement for an primer and the 5'-to-3' synthesis direction, leading to progressive shortening known as the end-replication problem, as independently proposed by James D. Watson in 1972 and Alexey Olovnikov in 1973. Centromeres, characterized by highly repetitive α-satellite DNA, pose replication hurdles due to their heterochromatic and propensity for secondary structures, which delay fork progression and increase breakage risk, necessitating specialized mechanisms like increased origin density and checkpoint activation to maintain stability. These structural features ensure that replication proceeds accurately while accommodating the topological constraints imposed by supercoiling and chromosomal architecture.

Enzymatic Machinery: DNA Polymerases and Associated Proteins

DNA replication relies on a suite of specialized enzymes and proteins that coordinate the unwinding, priming, and synthesis of new DNA strands. Central to this process is , an enzyme that catalyzes the addition of deoxyribonucleotides to a growing DNA chain in a template-directed manner. The first DNA polymerase was isolated and characterized from by and colleagues in 1956, marking a pivotal advancement in understanding enzymatic . DNA polymerases exhibit strict directionality, synthesizing new DNA strands exclusively in the 5' to 3' direction, which aligns with the antiparallel nature of DNA strands. They cannot initiate synthesis and require a short RNA or DNA primer with a free 3'-OH group to begin polymerization. Key properties include high processivity—the ability to add many before dissociating—and , achieved primarily through selective base-pairing mechanisms that discriminate correct from incorrect during incorporation. Processivity is dramatically enhanced by accessory factors such as sliding clamps; in , the β-clamp forms a ring around DNA, tethering the polymerase for extended synthesis, enabling the addition of thousands of per binding event. In prokaryotes like E. coli, multiple s exist with distinct roles, but DNA polymerase III (Pol III) serves as the primary replicative enzyme, forming a multi-subunit holoenzyme complex that ensures efficient chromosomal duplication. Pol III incorporates at rates exceeding 500 per second with error rates below 10^{-5} per base, bolstered by its proofreading 3'→5' exonuclease activity. In contrast, (Pol I) primarily fills gaps after RNA primer removal and performs repair functions, while Pol II contributes to replication restart and translesion synthesis under stress conditions. Eukaryotic cells employ a more complex set of replicative polymerases, including α (Pol α), δ (Pol δ), and ε (Pol ε). Pol α, associated with , initiates synthesis by extending short primers with DNA. Pol δ primarily handles lagging-strand synthesis, while Pol ε is dedicated to leading-strand elongation, both achieving high processivity through interactions with the PCNA sliding clamp analogous to the bacterial β-clamp. These polymerases collectively ensure accurate genome duplication across the larger eukaryotic chromosomes. Accessory proteins are indispensable for creating and maintaining a suitable environment for polymerase activity. Helicases, such as DnaB in E. coli, unwind the DNA double helix ahead of the replication fork in a 5'→3' direction, powered by ATP hydrolysis, to expose single-stranded templates. Primase, exemplified by DnaG in bacteria, synthesizes short RNA primers (typically 10-12 nucleotides) complementary to the DNA template, providing the necessary 3'-OH for polymerase initiation. Single-strand binding proteins (SSBs), like the SSB tetramer in E. coli, coat unwound single-stranded DNA to prevent reannealing, protect against nucleases, and facilitate the recruitment of other replication factors. Together, these components form a dynamic replisome, enabling coordinated and high-fidelity DNA synthesis.

Stages of the Replication Process

Initiation at Origins of Replication

In prokaryotes like , DNA replication initiates at a single chromosomal origin known as , a ~245 sequence characterized by an AT-rich DNA unwinding element () and multiple high- and low-affinity binding sites for the initiator protein , termed DnaA boxes. The consists of three 13-mer repeats that facilitate initial strand separation due to their low melting temperature. , bound to ATP, recognizes these boxes with , wrapping the DNA into a right-handed helical filament that promotes localized unwinding at the . The assembly of the initiation complex at oriC begins with DnaA oligomerization on the DnaA boxes, which distorts the DNA and exposes single-stranded regions in the DUE. This unwound region allows DnaA to recruit the DnaB helicase (in complex with DnaC) via direct protein-protein interactions, loading two DnaB hexamers onto the separated strands in a head-to-head orientation. Once loaded, DnaB further unwinds the DNA, and the DnaG primase associates with DnaB to synthesize short RNA primers, marking the transition to elongation. This process ensures precise and regulated initiation at oriC. In eukaryotes, replication origins are distributed across chromosomes, with the human genome featuring 30,000 to 50,000 such sites to accommodate the large genome size and complete replication within the cell cycle. Unlike prokaryotes, eukaryotic origins often lack a strict sequence consensus; in budding yeast (Saccharomyces cerevisiae), they are defined by autonomously replicating sequences (ARS) containing a conserved 17-bp ARS consensus sequence (ACS) that serves as a binding platform. In higher eukaryotes, origin selection is more flexible, influenced by chromatin accessibility, histone modifications, and non-sequence-specific factors rather than rigid DNA motifs. Eukaryotic initiation assembly centers on the (ORC), a conserved heterohexameric protein (Orc1-6) that binds origins in an ATP-dependent manner to mark potential start sites. ORC recruits the Cdc6 and the licensing factor Cdt1, which together load two head-to-head MCM2-7 hexameric complexes onto the double-stranded DNA, encircling it without initial unwinding and forming the pre-replicative complex (pre-RC). The MCM complexes serve as the replicative , and their loading "licenses" the origin for future activation. To prevent re-licensing and re-replication within the same , Cdt1 is inhibited by binding to geminin, a cell cycle-regulated protein that accumulates in S, G2, and M phases. Pre-RC formation occurs primarily during the of the , ensuring that origins are licensed before entry, while activation— involving kinase-mediated unwinding and primer synthesis by α-primase— is restricted to the to coordinate with progression. This temporal separation maintains genomic stability by limiting replication to once per cycle.

and Fork Progression

During the elongation phase of DNA replication, the replication adopts a characteristic Y-shaped structure, consisting of two diverging arms where the parental DNA double helix is unwound to form a bubble of single-stranded DNA. This unwinding creates a region of exposed template strands that serve as scaffolds for new DNA synthesis, with the progressing bidirectionally away from the in both prokaryotes and eukaryotes. In prokaryotes such as , this process was first visualized through autoradiography as theta (θ)-shaped intermediates, confirming the bidirectional nature of movement.80070-4) DNA synthesis at the fork proceeds through the incorporation of deoxynucleoside triphosphates (dNTPs) by DNA polymerases, which catalyze the formation of phosphodiester bonds while releasing as a , driving the reaction forward energetically. The leading strand is synthesized continuously in the 5' to 3' direction toward the advancing fork, whereas the lagging strand is synthesized discontinuously in short segments called , each initiated by an primer. These fragments, typically 1000–2000 nucleotides long in and 100–200 in eukaryotes, were discovered by Reiji Okazaki and colleagues in 1968 through pulse-labeling experiments on E. coli DNA. The overall rate of fork progression varies by organism: approximately 1000 nucleotides per second in bacterial systems like E. coli, enabling rapid duplication, compared to about 50 nucleotides per second in eukaryotes, reflecting the added complexity of and larger genomes. As the replication fork advances, the unwinding of the DNA helix generates positive supercoils ahead of the fork, which must be relieved to prevent stalling and breakage. Topoisomerases manage this topological stress: type I topoisomerases, such as I, introduce transient single-strand breaks to relax supercoils without ATP, while type II topoisomerases, like in bacteria or topoisomerase II in eukaryotes, use ATP-dependent double-strand breaks to remove supercoils and decatenate intertwined daughter strands. This coordinated action ensures smooth fork progression and maintains genomic integrity throughout elongation.

Termination and Completion

In prokaryotes, particularly in Escherichia coli, DNA replication termination is precisely controlled by a replication fork trap mechanism involving the Tus protein and specific Ter sites arranged in the terminus region opposite the origin of replication (oriC). The E. coli chromosome is circular, with bidirectional replication forks initiating at oriC and progressing until they converge in the terminus region, located approximately 180° opposite oriC on the circular map. This region contains 10 Ter sites (TerA through TerJ) organized into two oppositely oriented clusters that create a trap: Tus binds tightly to these 21-base-pair Ter sequences, forming polar barriers that halt approaching replication forks in a direction-specific manner while permitting passage in the opposite direction. The first Ter sites (TerA and TerB) were identified in 1988, and the tus gene, encoding the 309-amino-acid Tus protein, was cloned and characterized in 1989, revealing its role as a DNA-binding terminator that interacts with the DnaB helicase to arrest fork progression. This system ensures efficient fork convergence and prevents over-replication, with Tus-Ter complexes trapping the final forks to coordinate termination. In eukaryotes, termination lacks dedicated Ter-like barriers and instead occurs stochastically when replication forks from adjacent origins converge at random inter-origin sites along linear chromosomes. Fork convergence requires the resolution of topological constraints, primarily through decatenation by topoisomerase II (Topo II), which removes intertwinings (catenanes) between newly replicated to allow their separation. Inactivation or depletion of Topo II in leads to incomplete replication, as unresolved catenanes stall forks and prevent replisome disassembly, highlighting its essential role in termination.00303-1) Linear eukaryotic chromosomes face an additional challenge at their ends: the end-replication problem, where the lagging-strand RNA primer at the terminus cannot be fully replaced, leading to progressive shortening with each . This issue, first proposed in 1971, is mitigated by , a ribonucleoprotein enzyme discovered in 1985 that extends telomeres by adding TTAGGG repeats using its template. Following fork convergence in both prokaryotes and eukaryotes, post-termination processing completes duplication. RNA primers from on the lagging strand are removed—by activity in prokaryotes and by flap endonuclease 1 (FEN1) coordinated with δ in eukaryotes—and the resulting nicks are sealed by to form continuous strands. In prokaryotes, NAD+-dependent accomplishes this, while eukaryotic I, associated with PCNA, performs the final ligation during . reassembly then restores epigenetic marks and structure on the duplicated DNA; in eukaryotes, this is mediated by assembly factor 1 (CAF-1), which deposits H3-H4 tetramers onto newly synthesized DNA in a replication-coupled manner, ensuring faithful transmission of organization.00129-8) These steps are critical for genome stability, with defects leading to chromosomal aberrations or arrest.

Strand-Specific Synthesis Mechanisms

Leading Strand Synthesis

The leading strand is synthesized continuously in the 5' to 3' direction, aligning with the movement of the replication fork, where DNA helicase unwinds to expose the template strand. This process begins at the with a single RNA primer synthesized by , which then extends without interruption, adding to the 3' end of the growing chain as the fork progresses. In bacteria, such as Escherichia coli, the DNA polymerase III (Pol III) holoenzyme serves as the primary replicative polymerase for leading strand synthesis, achieving high processivity through association with the β sliding clamp. The holoenzyme's core, comprising the α catalytic subunit for nucleotide addition, the ε proofreading subunit for error correction, and the θ stabilizing subunit, is tethered to the DNA by the ring-shaped β clamp, which encircles the duplex DNA and slides along it, preventing dissociation and enabling synthesis of thousands of nucleotides per binding event. The clamp is loaded onto the primed template in an ATP-dependent manner by the γ complex (clamp loader), ensuring efficient, continuous elongation at rates up to 1000 nucleotides per second. In eukaryotes, ε (Pol ε) performs the bulk of leading strand synthesis, forming a tetrameric holoenzyme with catalytic Pol2, accessory subunits Dpb2, Dpb3, and Dpb4. Pol ε physically couples with the CMG complex via interactions mediated by Dpb2's OB-fold domain, which channels the emerging single-stranded template directly to the active site for coordinated unwinding and . This association enhances processivity, with Dpb3–Dpb4 stabilizing the on double-stranded DNA through a mooring , allowing high-fidelity replication across large chromosomal regions. The continuous of leading strand synthesis confers advantages in efficiency and , requiring only one priming event per replicon and minimizing initiation-related compared to discontinuous mechanisms. This continuity facilitates tight coordination with activity, where progression matches unwinding speed to maintain stability and expose without excessive single-stranded gaps, ultimately achieving rates as low as 1 per 10^9 through integrated .

Lagging Strand Synthesis

The lagging strand is synthesized discontinuously in short segments known as , a process necessitated by the antiparallel nature of DNA strands and the unidirectional 5' to 3' synthesis by s. This contrasts with the continuous extension of the leading strand. Each Okazaki fragment begins with an primer synthesized by , followed by DNA polymerase extension until it reaches the previous fragment's primer region. The discovery of Okazaki fragments in 1968 by Reiji Okazaki and colleagues provided key evidence for this discontinuous mechanism during bacteriophage T4 DNA replication in Escherichia coli. In prokaryotes, these fragments are typically 1,000 to 2,000 nucleotides long, while in eukaryotes, they average 100 to 200 nucleotides. The shorter eukaryotic fragments reflect differences in primase efficiency and replication fork speed.00093-6) In eukaryotes, δ (Pol δ) primarily extends the RNA primers on the lagging strand, associating with the for processivity. After synthesis, the RNA primers are removed through coordinated nuclease activity: RNase H2 cleaves most of the RNA, leaving a flap that is processed by the 5' flap endonuclease FEN1, often in conjunction with .00157-X) The resulting nick is then sealed by DNA ligase I, completing the fragment and forming a continuous strand.00157-X) To coordinate synthesis, the lagging strand polymerase recycles via the trombone model, where the template DNA loops out, allowing the to remain tethered to the while synthesizing multiple fragments without dissociating. This model, first proposed by for the T4 phage system, ensures efficient coupling with leading strand progression. The discontinuous nature of lagging strand synthesis, involving multiple priming and joining events, contributes to a higher compared to the leading strand, particularly at lesion sites, due to increased opportunities for replication errors during fragment initiation and processing.

Coordination at the Replication Fork

The is a that coordinates DNA unwinding, priming, and synthesis at the replication fork to ensure efficient and directional progression. In prokaryotes, such as , the assembles as a coupled unit comprising , , and , with the encircling the lagging strand to drive fork movement while polymerases synthesize both strands. Single-stranded DNA-binding protein (SSB) stabilizes the unwound single-stranded DNA regions, preventing reannealing and secondary structure formation, thereby maintaining fork integrity and facilitating access for Okazaki fragment initiation. In eukaryotes, the core centers on the CMG (Cdc45-MCM2-7-GINS) , which translocates along the leading strand in a 3' to 5' direction, coupled with (Pol ε) for leading-strand synthesis and (Pol δ) for lagging-strand synthesis, supported by Pol α- for primer synthesis. Coordination between leading- and lagging-strand synthesis is achieved through physical tethering and looping mechanisms within the , ensuring synchronized progression despite the discontinuous nature of lagging-strand synthesis. The lagging strand enhances overall replisome processivity by approximately 61%, extending the coupled synthesis distance from 52 kb (leading strand alone) to 86 kb, likely due to dual anchoring via sliding clamps that provide increased DNA grip. However, this coupling reduces fork speed by about 23%, from 317 nt/s to 246 nt/s, reflecting the periodic repriming and looping required for . Fork stalling, often triggered by DNA lesions on the leading strand, is managed through recovery pathways involving lesion skipping by repriming enzymes like PrimPol in eukaryotes or translesion synthesis (TLS) s that bypass damage while maintaining replisome integrity via interactions with clamps and repair factors. In eukaryotes, the proliferating cell nuclear antigen (PCNA) sliding clamp, loaded onto DNA by the replication factor C (RFC) complex at primer-template junctions, enhances polymerase processivity and facilitates switching between replicative and TLS polymerases during stalling events. RFC captures the 3' ss/dsDNA junction, partially melts the duplex, and loads PCNA in a multistep, ATP-dependent process that closes the clamp around DNA without hydrolysis in the final step, promoting efficient Okazaki fragment extension. The CMG helicase integrates with Pol ε to form a stable 15-subunit holoenzyme (CMG^E), where the Dpb2 subunit of Pol ε binds the GINS component of CMG, ensuring leading-strand specificity and fork directionality at rates up to 1.92 kb/min when coupled with PCNA. Replisome speed is regulated to align with cell cycle demands, slowing during nutrient limitation or stationary phase to adapt elongation rates (e.g., from exponential to stationary growth), which delays completion without invoking damage responses. Recent cryo-EM structures post-2010, such as those of Drosophila CMG at 7.4–9.8 Å resolution, reveal dynamic ATPase states that grip or release DNA, supporting monomeric CMG translocation on the leading strand while implying loose dimeric tethering for bidirectional forks, thus elucidating replisome stability and uncoupling mechanisms.

Regulation and Control Mechanisms

Prokaryotic Replication Control

In prokaryotes, DNA replication is tightly regulated to ensure it occurs once per per , coordinating with rapid bacterial growth and division. This control is streamlined for unicellular organisms, relying on mechanisms that link initiation to accumulation and prevent over-replication through feedback loops. Unlike the complex, multi-phase licensing in eukaryotes, bacterial systems emphasize efficiency, with initiation primarily governed at the () in model organisms like . Initiation of replication is controlled by the protein, which binds to in its ATP-bound form (DnaA-ATP) to unwind the DNA and assemble the , while the ADP-bound form (DnaA-) is inactive. The DnaA-ATP/ADP cycle is regulated by regulatory inactivation of (), where Hda protein, associated with the β-clamp on newly replicated DNA, stimulates on , reducing active DnaA levels post-initiation to prevent re-initiation. Additionally, the locus titrates DnaA and promotes its hydrolysis, further fine-tuning the timing. This cycling ensures replication initiates only when DnaA-ATP levels are sufficient, typically tied to rate, as faster-growing E. coli cells accumulate more DnaA per origin, allowing initiations at smaller cell sizes and enabling multifork replication during rapid division (doubling times as short as 20 minutes). Post-replication, oriC is sequestered to block premature re-initiation. Immediately after fork passage, the newly duplicated becomes hemimethylated at GATC sites, as Dam methylase lags behind replication. The SeqA protein preferentially binds these hemimethylated sequences, preventing DnaA from accessing and sequestering it for about one-third of the (roughly 10-15 minutes in E. coli). Full remethylation by Dam methylase then restores accessibility, completing the sequestration cycle and ensuring a refractory period. This mechanism is essential for maintaining replication timing, as seqA mutants exhibit asynchronous initiations and over-replication. Prokaryotes lack the elaborate checkpoints of eukaryotes, with minimal cell cycle pauses beyond basic damage sensing. Instead, the RecA protein mediates the SOS response to replication stress or damage, forming nucleoprotein filaments on single-stranded DNA at stalled forks to halt progression, induce error-prone repair, and facilitate fork restart via homologous recombination. This response prioritizes survival over strict fidelity, allowing replication to resume quickly in dynamic environments. In E. coli, these controls enable the 4.6 Mb to replicate in approximately 40 minutes via bidirectional forks moving at ~1,000 base pairs per second each, despite generation times shorter than this in fast growth, achieved through overlapping replication rounds.

Eukaryotic Replication Licensing and Timing

In eukaryotic cells, DNA replication is tightly regulated to ensure that the is duplicated exactly once per , a process that begins with the licensing of replication origins during the . Licensing involves the assembly of the (pre-RC) at origins, where the (ORC) binds to DNA and recruits Cdc6 and Cdt1 proteins, which in turn load the MCM2-7 complex as a double hexamer around the DNA. This MCM loading renders origins "licensed" and competent for future activation, occurring exclusively in when (CDK) activity is low. To prevent re-replication, high CDK levels in , , and phases phosphorylate pre-RC components, inhibiting their rebinding to origins and promoting their degradation or nuclear export, thus ensuring licensing is restricted to post-mitotic . Once licensed, origins fire stochastically during , with only a activating while many remain dormant as a mechanism against replication stress. Origin firing is triggered by S-phase CDKs and Dbf4-dependent (DDK), which phosphorylate MCM and associated factors to unwind DNA and recruit polymerases, but the timing and efficiency vary due to local context and inter-origin spacing. In cells, the contains approximately 30,000 to 50,000 potential origins, spaced about 50-100 kb apart, allowing efficient coverage of the 6 billion base pairs within the ~8-hour . Active replication forks cluster into "replication factories" or foci, where multiple origins within a chromosomal coordinate to form these immobile sites, facilitating processive synthesis and enabling dormant origins to fire locally if nearby forks stall. S-phase progression is temporally regulated to replicate early-firing euchromatic regions before late-firing , coordinating with to ensure complete duplication before chromosome segregation. This timing is influenced by epigenetic marks and nuclear positioning, with checkpoints halting progression if forks are impeded. Telomere maintenance during replication involves specialized mechanisms, as the linear ends pose an "end-replication problem"; while standard origins fire inefficiently here, proteins and alternative lengthening pathways ensure telomere integrity without relying on in all cases. Recent studies from the highlight origin plasticity under stress, where dormant origins are dynamically recruited to counteract fork stalling from DNA damage or depletion, adapting the replication program to maintain genome stability.

Fidelity, Errors, and Repair

Sources of Replication Errors

DNA replication errors arise primarily from intrinsic limitations in the fidelity of DNA s and from spontaneous or induced chemical alterations to the DNA template. During incorporation, base mismatches occur when the selects an incorrect dNTP, often due to transient tautomerization of bases, where keto-enol shifts in like or lead to non-standard Watson-Crick pairing, such as G pairing with T instead of C. These mismatches contribute to transition mutations (purine-to-purine or pyrimidine-to-pyrimidine substitutions). Insertions and deletions (indels) are another common error type, particularly in repetitive sequences like microsatellites or homopolymer runs, where slippage during strand synthesis causes frameshift mutations; this is exacerbated by imbalanced dNTP pools that favor misalignment. The intrinsic error rate of replicative polymerases, such as Pol δ and Pol ε in eukaryotes, is approximately 10^{-4} to 10^{-5} errors per incorporated , though rates are slightly lower at around 10^{-7} due to contextual factors like replication fork speed. Spontaneous chemical damage to DNA also serves as a major source of replication errors, independent of polymerase activity. Depurination, the hydrolysis of the N-glycosyl bond releasing or , occurs at a rate of about 5,000 purine bases lost per per day, leaving an apurinic ( that, if unreplicated, can result in base deletions or transversions during as the inserts an opposite the void. , another frequent event, affects (converting it to uracil at ~100 sites per per day) or (to hypoxanthine), leading to C·G to T·A transitions if the altered base is used as a template, since uracil pairs with . External factors amplify these errors; (UV) radiation induces cyclobutane (e.g., T-T dimers) that stall polymerases and promote error-prone bypass, while chemical mutagens like alkylating agents form adducts (e.g., O^6-methylguanine) that mispair with , increasing G·C to A·T transitions. Certain genomic regions, such as replication origins and repetitive elements, act as error hotspots due to structural features like secondary structures or high , and these hotspots show evolutionary conservation across species, suggesting selective pressures maintain them for functions like recombination despite mutagenic risk. Overall, despite these sources, the net in humans is remarkably low at approximately 1.2 × 10^{-8} per per generation, reflecting the baseline error burden after all processes. These replication errors play dual roles: in , they generate essential for and , while in , elevated error rates contribute to genomic , driving somatic mutations in cancers such as colorectal and endometrial tumors where polymerase variants like mutations increase .

Proofreading and Error Correction

During DNA replication, proofreading is an intrinsic error-correction mechanism performed by replicative DNA polymerases, which possess a 3'→5' exonuclease activity that excises mismatched nucleotides immediately after incorporation. This activity allows the polymerase to reverse its polymerization step, removing the incorrect base from the 3' end of the growing strand before resuming synthesis. In eukaryotes, the leading-strand polymerase ε (Pol ε) relies on its catalytic subunit's exonuclease domain for this proofreading, enhancing replication fidelity by detecting and correcting base-pairing errors with high efficiency. Without proofreading, the intrinsic error rate of nucleotide incorporation by DNA polymerases is approximately 10^{-4} to 10^{-5}, but the exonuclease activity improves accuracy by a factor of 10^{2} to 10^{3}, reducing errors to about 10^{-7} per base pair. Post-replication, mismatch repair (MMR) provides an additional layer of fidelity by scanning the newly synthesized DNA for persistent mismatches that escaped proofreading. In bacteria, the MutS protein recognizes mismatched bases, forming a complex that recruits MutL to initiate repair; strand discrimination occurs via dam methylation, where the unmethylated daughter strand is targeted for excision. In eukaryotes, homologs such as MSH2-MSH6 (MutSα) detect mismatches, while MLH1-PMS2 (MutLα) coordinates excision; strand discrimination relies on nicks or gaps in the nascent strand, often introduced by ribonucleotide incorporation or Okazaki fragment processing. MMR excises a segment of the error-containing strand (typically 100-1000 nucleotides) using helicase and exonuclease activities, followed by resynthesis and ligation, further reducing the error rate by 10^{2} to 10^{3}-fold. Combined with base selection and proofreading, MMR achieves an overall replication fidelity of 10^{-9} to 10^{-10} errors per base pair. Defects in MMR genes, such as MLH1 or MSH2 mutations, underlie , leading to and a dramatically elevated that predisposes carriers to colorectal and other cancers. Beyond MMR, other repair pathways address replication-associated damage: (BER) removes oxidized or alkylated bases via glycosylases, creating single-strand breaks that are processed during or shortly after replication to prevent fork stalling. For non-instructive lesions that block high-fidelity polymerases, translesion synthesis employs specialized low-fidelity polymerases (e.g., Pol κ or Pol ζ) to bypass the damage, allowing replication to continue while deferring accurate repair. These mechanisms collectively ensure high-fidelity duplication, though translesion bypass introduces errors at rates up to 10^{-3} per lesion.

Implications for Genome Stability

DNA replication plays a crucial role in maintaining stability by minimizing mutations that could lead to oncogenic transformations. Random errors during account for approximately two-thirds of the mutations driving human cancers, independent of environmental or hereditary factors, as these arise from the inherent stochasticity of activity in proliferating cells. In tumor cells, replication stress—often induced by oncogene activation such as or —exacerbates fork stalling and collapse, promoting genomic instability and facilitating tumor evolution through the accumulation of chromosomal aberrations. This stress response, if unchecked, can trigger or as a barrier to tumorigenesis, but evasion of these safeguards allows cancer progression. Telomere shortening during successive replication cycles contributes to cellular aging and organismal lifespan limits by eroding protective chromosomal ends, ultimately leading to replicative . In 1961, Hayflick and Moorhead observed that human diploid fibroblasts undergo approximately 50 population doublings before entering , a phenomenon now linked to progressive telomere attrition of about 50-100 base pairs per division in the absence of activity. This "" underscores how replication-imposed constraints prevent indefinite proliferation, thereby safeguarding against immortalization in aging tissues. Pathologies like exemplify the consequences of impaired replication fork protection; defects in the Fanconi anemia pathway lead to hypersensitivity to interstrand crosslinks, causing frequent fork collapse and double-strand breaks that heighten cancer risk, particularly leukemias. From an evolutionary perspective, replication errors serve as a primary source of , enabling through the introduction of beneficial mutations under selective pressure. Studies in microbial systems reveal that replication-induced copy number variations and point mutations drive adaptive evolution, such as antibiotic resistance in , by generating heritable variation at rates tuned by . Core replication mechanisms, including the architecture and origin recognition, are highly conserved across prokaryotes and eukaryotes, reflecting their ancient origins and essentiality for integrity from to humans. In modern , CRISPR-Cas9 off-target effects in the 2020s have highlighted replication's vulnerability, as unintended double-strand breaks can induce fork stalling and , compromising stability in edited cells and necessitating improved strategies.

Applications and Techniques

In Vitro Replication Methods

In vitro replication methods enable the study of outside living cells using purified enzymes and cell-free extracts, providing insights into the biochemical mechanisms of replication. The foundational achievement came in 1957 when and colleagues isolated from and demonstrated its ability to synthesize DNA from a DNA template, marking the first enzymatic replication in a . This system required a primed DNA template, deoxynucleoside triphosphates, and magnesium ions, but initially produced short DNA fragments due to the enzyme's low processivity. During the 1970s, Kornberg's group advanced bacterial in vitro replication by reconstituting more complete systems with purified proteins, notably replicating the single-stranded DNA genome of bacteriophage φX174. This involved assembling a replisome-like complex including , , , and , allowing semi-conservative replication starting from an intact phage template. These efforts revealed key accessory factors, such as the β-sliding clamp (the core subunit of the γ complex), which dramatically enhanced processivity from ~10 to thousands, enabling efficient fork progression. For eukaryotic systems, cell-free extracts from mammalian cells infected with simian virus 40 () provided a model to study viral DNA replication dependent on host machinery. In 1984, Li and Kelly established an system using cell extracts, the viral , and the , which recruits cellular replication factors like α-primase and to initiate bidirectional synthesis. This setup recapitulated theta-mode replication intermediates and required ATP, but relied on crude extracts rather than fully purified components. More recent advances have achieved full reconstitution of eukaryotic replisomes with purified proteins. In 2015, a (Saccharomyces cerevisiae) system was developed using 31 distinct polypeptides, enabling coupled leading- and lagging-strand synthesis while suppressing nucleoprotein filament formation. In 2022, a was reconstituted with 11 purified factors, demonstrating fast and efficient replication of DNA templates at rates comparable to processes. Key advances in the 2010s included reconstitution of the E. coli chromosome replication cycle using 14 purified enzymes (25 polypeptides), enabling exponential propagation of circular DNA templates and multiple rounds of replication without added primers. This system, developed by Su’etsugu and colleagues, incorporated for origin unwinding and demonstrated coordinated leading- and lagging-strand synthesis at rates approaching speeds. Such reconstitutions facilitated structural studies via cryo-electron microscopy (cryo-EM), revealing dynamic architectures, including polymerase-clamp interactions during fork progression in bacteriophage T7 systems. However, these methods face limitations, particularly for complex eukaryotes, where chromatin assembly, histone modifications, and numerous accessory factors are not fully recapitulated, leading to incomplete fidelity and regulation compared to prokaryotic models. In vitro approaches laid the groundwork for techniques like , which amplify specific DNA segments through repeated thermal cycling.

Polymerase Chain Reaction (PCR)

The Polymerase Chain Reaction (PCR) is an in vitro technique that selectively amplifies specific DNA segments through repeated thermal cycling, enabling the generation of billions of copies from minute starting amounts for analysis in research, diagnostics, and other fields. Developed by biochemist Kary Mullis in 1983 during his tenure at Cetus Corporation, PCR was first demonstrated in a 1985 publication and marked a pivotal advancement in nucleic acid manipulation. Mullis received the Nobel Prize in Chemistry in 1993 for this invention, recognizing its transformative impact on biology and medicine. The PCR process relies on three sequential steps cycled 20–40 times: denaturation, annealing, and extension, powered by a thermostable DNA polymerase. Denaturation heats the reaction to 94–98°C for 20–30 seconds, separating the double-stranded DNA template into single strands by disrupting hydrogen bonds. Annealing cools the mixture to 50–65°C for 20–40 seconds, permitting two synthetic oligonucleotide primers—short DNA sequences designed to flank the target region—to hybridize specifically to their complementary sites on the template strands. Extension then occurs at 72°C for 30 seconds to 2 minutes, during which the DNA polymerase, typically Taq derived from the thermophilic bacterium Thermus aquaticus, synthesizes new DNA strands by incorporating deoxynucleotide triphosphates (dNTPs) along the template starting from the 3' end of each primer. This cyclic process results in exponential amplification, producing theoretically $2^n copies of the target sequence after n cycles, though actual yields are slightly lower due to inefficiencies. Taq polymerase, originally isolated from T. aquaticus cells grown in hot springs, remains stable at high temperatures, eliminating the need to replenish the enzyme after each denaturation step. Key reaction components include the target DNA template (often nanograms or less), forward and reverse primers (typically 18–22 long), a of the four dNTPs (dATP, dCTP, dGTP, dTTP), a buffered with Mg²⁺ ions to optimize activity and primer annealing, and the thermostable . These elements are assembled in a small volume (10–50 μL) and subjected to automated temperature control in a device. The specificity of amplification is dictated by primer design, allowing precise targeting of genes or regions of interest. Variants of PCR address limitations of the standard method and expand its scope. Reverse transcription PCR (RT-PCR) incorporates an initial reverse transcription step using reverse transcriptase enzyme to convert into (cDNA), followed by amplification; this facilitates studies of RNA expression levels, viral genomes, and transcriptomics. Quantitative PCR (qPCR), or real-time PCR, integrates fluorescent reporter molecules (such as SYBR Green dye or probes) to detect and quantify product accumulation during each cycle via measurement, providing data on initial template abundance through the threshold cycle (Ct) value where signal exceeds background. These techniques underpin diverse applications, including for amplifying degraded or trace DNA from evidence like bloodstains or touch samples to generate DNA profiles for suspect identification, and clinical diagnostics for rapid pathogen detection (e.g., in or testing), screening, and monitoring disease progression through assessment. Compared to cellular DNA replication, standard PCR with Taq polymerase exhibits higher infidelity, with error rates around $10^{-4} to $10^{-5} mutations per base pair per cycle, attributable to the absence of robust proofreading mechanisms. Such errors can accumulate, particularly in long amplicons or numerous cycles, potentially introducing artifacts in downstream analyses like sequencing. High-fidelity polymerases, such as those engineered with 3'–5' exonuclease activity (e.g., Pfu from or blends like Phusion), reduce error rates to approximately $10^{-6} or better, enhancing accuracy for applications requiring precise sequence fidelity, such as or variant detection.

References

  1. [1]
    DNA Replication - National Human Genome Research Institute
    DNA replication is the process by which the genome's DNA is copied in cells. Before a cell divides, it must first copy (or replicate) its entire genome.Missing: mechanism | Show results with:mechanism
  2. [2]
    DNA Replication Mechanisms - Molecular Biology of the Cell - NCBI
    An enzyme called DNA ligase then joins the 3′ end of the new DNA fragment to the 5′ end of the previous one to complete the process (Figures 5-13 and 5-14).The DNA Replication Fork Is... · The High Fidelity of DNA... · Special Nucleotide...
  3. [3]
    Biochemistry, DNA Replication - StatPearls - NCBI Bookshelf
    Mechanism. The actual mechanism is brought by the interaction of a multitude of proteins and enzymes and occurs during the S phase. · Initiation · Elongation.Introduction · Molecular Level · Function · Mechanism
  4. [4]
    A Structure for Deoxyribose Nucleic Acid - Nature
    The determination in 1953 of the structure of deoxyribonucleic acid (DNA), with its two entwined helices and paired organic bases, was a tour de force in ...
  5. [5]
    The twisted circular form of polyoma viral DNA. - PNAS
    The twisted circular form of polyoma viral DNA. J Vinograd, J Lebowitz, R Radloff, +1 , R Watson, and P Laipis -1Authors Info & Affiliations. May 15, 1965.
  6. [6]
    The bacterial chromosome and its manner of replication as seen by ...
    The chromosome of E. coli consists of a single piece of two-stranded DNA, 700 to 900 μ long. (2) This DNA duplicates by forming a fork.
  7. [7]
    The replication of DNA in Escherichia coli* | PNAS
    In this paper reference will be made to the apparent molecular weight of DNA samples determined by means of density-gradient centrifugation.Missing: URL | Show results with:URL
  8. [8]
    The dark side of centromeres: types, causes and consequences of ...
    Oct 18, 2018 · In conclusion, replicating centromeric DNA is a difficult task for the cells due to the accumulation of large regions of repetitive DNA. ...
  9. [9]
    Fidelity of DNA replication—a matter of proofreading - PMC
    The fidelity of DNA replication relies on nucleotide selectivity of replicative DNA polymerase, exonucleolytic proofreading, and postreplicative DNA mismatch ...
  10. [10]
    The bacterial DNA sliding clamp, β-clamp: structure, interactions ...
    May 30, 2024 · An essential factor in the bacterial replication complex is the ring-shaped DNA sliding clamp, β-clamp, ensuring processive DNA replication and ...
  11. [11]
    DNA replication fidelity in Escherichia coli: a multi-DNA polymerase ...
    The major replicative DNA polymerase in E. coli cells is DNA Pol III HE that can synthesize DNA with high processivity (over 50 kb per binding event) (Yao et al ...DNA Pol III HE · Role of DNA polymerase II... · Role of DNA polymerase I (Pol I)
  12. [12]
    Roles for E coli DNA polymerases I, II, and III in DNA replication
    May 10, 1974 · Chain elongation during E. coli DNA replication proceeds in three distinct stages. Participatory roles of the three E. coli DNA polymerases ...
  13. [13]
    Eukaryotic DNA polymerases in DNA replication and DNA repair
    DNA polymerase delta functions as a dimer and, therefore, may be responsible for both leading and lagging strand DNA replication. In addition, this enzyme is ...
  14. [14]
    The Escherichia coli dnaB replication protein is a DNA helicase
    Apr 5, 1986 · We have found that the dnaB protein is a DNA helicase that is capable of unwinding extensive stretches of double-stranded DNA.
  15. [15]
    Primase, the dnaG protein of Escherichia coli. An enzyme ... - PubMed
    Feb 10, 1978 · This stage of G4 DNA replication requires action of three proteins: the E. coli DNA-binding protein, the dnaG protein, and the DNA polymerase ...
  16. [16]
    The single-stranded DNA-binding protein of Escherichia coli - NIH
    The single-stranded DNA-binding protein (SSB) of Escherichia coli is involved in all aspects of DNA metabolism: replication, repair, and recombination.
  17. [17]
    The DnaA Cycle in Escherichia coli: Activation, Function ... - Frontiers
    Dec 20, 2017 · The 245 bp minimal oriC region has multiple binding sites for the chromosomal replication initiator protein DnaA (DnaA boxes), and a single ...
  18. [18]
    The bacterial replication initiator DnaA. DnaA and oriC, the bacterial ...
    oriC of E. coli contains in 260 bp five DnaA boxes, and at its left end an AT-rich region consisting of three 13-mer repeats and the so-called AT cluster (see ...Domain functions · DnaA oligomerization and... · Variations of the theme: DnaA...
  19. [19]
    Mechanism of origin unwinding: sequential binding of DnaA to double
    The AT-rich region of E.coli oriC contains three 13mer repeats, each of them starting with a GATC site that is part of a potential ATP-DnaA box. In total, there ...
  20. [20]
    Two discriminatory binding sites in the Escherichia coli replication ...
    The second possibility is that oriC contains binding sites specific for DnaA-ATP. DnaA-ATP binding to such sites would be analogous to the interaction of ...Sign Up For Pnas Alerts · Materials And Methods · Results
  21. [21]
    DnaB helicase is recruited to the replication initiation complex via ...
    DnaB helicase is recruited to the replication initiation complex via binding of DnaA domain I to the lateral surface of the DnaB N-terminal domain.
  22. [22]
    Concerted actions of DnaA complexes with DNA-unwinding ...
    May 18, 2022 · Unwinding of the replication origin and loading of DNA helicases underlie the initiation of chromosomal replication. In Escherichia coli ...
  23. [23]
    Eukaryotic DNA replication origins: many choices for ... - PubMed
    At each cell division in humans, 30,000-50,000 DNA replication origins are activated, and it remains unclear how they are selected and recognized by replication ...
  24. [24]
    Prediction of Saccharomyces cerevisiae replication origins
    Mar 4, 2004 · At the core of every yeast replication origin is a replicator sequence containing a conserved 17 bp stretch known as the ARS consensus sequence ...Missing: seminal | Show results with:seminal
  25. [25]
    The origin recognition complex: a biochemical and structural view
    Aside from the fact that origins are located near to ORC binding sites and promote initiation of DNA replication, there is little consensus among eukaryotic ...
  26. [26]
    ATP-dependent recognition of eukaryotic origins of DNA replication ...
    May 14, 1992 · ATP-dependent recognition of eukaryotic origins of DNA replication by a multiprotein complex. Stephen P. Bell &; Bruce Stillman. Nature volume ...Missing: discovery | Show results with:discovery
  27. [27]
    A double-hexameric MCM2-7 complex is loaded onto origin DNA ...
    Cdc6 ATP hydrolysis is required for MCM2-7 loading, and Orc1 ATP hydrolysis promotes the release of the MCM2-7 complex from ORC to finish the process of pre-RC ...
  28. [28]
    Inhibition of eukaryotic DNA replication by geminin binding to Cdt1
    Here we show that geminin interacts tightly with Cdt1, a recently identified replication initiation factor necessary for MCM loading.
  29. [29]
    Initiation of Eukaryotic DNA Replication: Molecular Cell
    At the G1/S transition, S phase–specific cyclin-dependent protein kinase (Cdk) and Cdc7/Dbf4 kinase transform the pre-RC into an active replication fork. During ...<|control11|><|separator|>
  30. [30]
    Origins of DNA replication - PMC - PubMed Central
    DNA synthesis of daughter strands starts at discrete sites, termed replication origins, and proceeds in a bidirectional manner until all genomic DNA is ...
  31. [31]
    The E. coli DNA Replication Fork - ScienceDirect.com
    Once assembled, the E. coli replisome unwinds and duplicates DNA at remarkable speeds, approaching 1000 bp/s with an error rate of approximately one mutation ...Missing: per | Show results with:per
  32. [32]
    The Initiation and Completion of DNA Replication in Chromosomes
    To replicate such a DNA molecule from end to end with a single replication fork moving at a rate of 50 nucleotides per second would require 0.02 × 150 × 106 = ...Missing: eukaryotic | Show results with:eukaryotic
  33. [33]
    Human topoisomerases and their roles in genome stability ... - Nature
    Feb 28, 2022 · During replication fork progression, topoisomerases dissipate the supercoiling generated by DNA unwinding. Without topoisomerase activity ...
  34. [34]
    Top1- and Top2-mediated topological transitions at replication forks ...
    DNA topoisomerases solve topological problems during chromosome metabolism. We investigated where and when Top1 and Top2 are recruited on replicating ...
  35. [35]
    Mechanism of termination of DNA replication of Escherichia coli ...
    We show further that replication termination involves protein–protein interaction between Tus and DnaB at a critical region of Tus protein, called the L1 loop.<|control11|><|separator|>
  36. [36]
    Identification of a specific telomere terminal transferase ... - PubMed
    We have found a novel activity in Tetrahymena cell free extracts that adds tandem TTGGGG repeats onto synthetic telomere primers.
  37. [37]
    DNA Replication - The Cell - NCBI Bookshelf - NIH
    Because DNA polymerases extend primers only in the 5′ to 3′ direction, they are unable to copy the extreme 5′ ends of linear DNA molecules. Consequently ...
  38. [38]
    Structure of the polymerase ε holoenzyme and atomic model of the ...
    Jun 22, 2020 · Pol ε performs bulk leading strand synthesis while Pol δ acts on the lagging strand. Pol α-primase contains both RNA primase and DNA polymerase ...
  39. [39]
    CMG helicase and DNA polymerase ε form a functional 15 ... - PNAS
    DNA replication in eukaryotes is asymmetric, with separate DNA polymerases (Pol) dedicated to bulk synthesis of the leading and lagging strands.
  40. [40]
    Synergism between CMG helicase and leading strand DNA ... - Nature
    Sep 20, 2023 · Coordinating the activities of replicative helicase and DNA polymerase(s) at the replication fork is crucial for efficient and accurate DNA ...
  41. [41]
  42. [42]
    Structure of eukaryotic DNA polymerase δ bound to the PCNA clamp ...
    Nov 17, 2020 · Pol δ performs lagging strand synthesis by extending primers generated by Pol α primase every 100 to 200 nt (9). Pol δ also synthesizes a small ...<|control11|><|separator|>
  43. [43]
    Replisome dynamics and use of DNA trombone loops to bypass ...
    Bruce Alberts first proposed the “trombone” DNA looping model of lagging strand DNA replication nearly thirty years ago. DNA looping on the lagging strand is ...Missing: original paper
  44. [44]
    Strand-resolved mutagenicity of DNA damage and repair - Nature
    Jun 12, 2024 · Our initial analysis demonstrated a significantly higher mutation rate for lagging strand synthesis over a lesion-containing template (Pearson's ...
  45. [45]
    How the Eukaryotic Replisome Achieves Rapid and Efficient DNA ...
    Dec 15, 2016 · DNA-combing experiments have shown the majority of replication forks in various human cell lines travel between 1 and 2 kb min−1, with an ...Missing: source | Show results with:source
  46. [46]
    Human single-stranded DNA binding proteins are essential for ...
    SSBs are essential for the stabilisation and restart of stalled replication forks. As mentioned above, RPA is essential for the replication of DNA during S- ...
  47. [47]
    Single-molecule analysis reveals that the lagging strand increases ...
    The lagging strand significantly increases the processivity of the replisome, possibly reflecting the increased grip to DNA provided by 2 DNA polymerases.<|separator|>
  48. [48]
    Lesion Bypass and the Reactivation of Stalled Replication Forks
    It is now clear that replication forks stall frequently as a result of encounters between the replication machinery and template damage.
  49. [49]
    Multistep loading of a DNA sliding clamp onto DNA by replication ...
    Aug 8, 2022 · PCNA is loaded around DNA by the ATP-dependent clamp loader replication factor C (RFC), which acts at single-stranded (ss)/double-stranded DNA (dsDNA) ...
  50. [50]
  51. [51]
    Cryo-EM structures of the eukaryotic replicative helicase bound to a ...
    Feb 18, 2016 · Here we use cryo-electron microscopy to describe two subnanometre resolution structures of the CMG helicase trapped on a DNA fork.
  52. [52]
    Regulatory elements coordinating initiation of chromosome ... - PNAS
    DnaA-ADP could have been a plausible inhibitor of initiation, but, although DnaA-ADP cannot initiate replication by itself, it has been shown to promote rather ...
  53. [53]
    The Initiator Function of DnaA Protein Is Negatively Regulated by ...
    DnaA inactivation occurs through accelerated hydrolysis of ATP bound to DnaA; the resulting ADP-DnaA fails to initiate replication.
  54. [54]
    DnaA binding locus datA promotes DnaA-ATP hydrolysis to ... - PNAS
    Dec 31, 2012 · DnaA-ATP hydrolysis is crucial to regulation of initiation and mainly occurs by a replication-coupled feedback mechanism named RIDA (regulatory ...
  55. [55]
    The Escherichia coli replication initiator DnaA is titrated on ... - Nature
    Aug 21, 2025 · Changes in DnaA ATP/ADP ratio or titration do not impact its function as transcription factor. Next to initiating DNA replication, DnaA is also ...
  56. [56]
    E. coli SeqA Protein Binds oriC in Two Different Methyl ... - Cell Press
    Sep 22, 1995 · Direct interaction of. SeqA protein with the replication origin is likely to be involved in both replication initiation and sequestration.Missing: post- | Show results with:post-
  57. [57]
    High-affinity binding of hemimethylated oriC by Escherichia coli ...
    This is consistent with the view that SeqA plays a role in sequestering the newly replicated origin by making it inaccessible to the replication machinery, and ...
  58. [58]
    E. coli oriC and the dnaA gene promoter are sequestered from dam ...
    We present evidence for three major conclusions: that oriC is a single function unit that is specifically sequestered from dam methyltransferase for a ...
  59. [59]
    Replication is required for the RecA localization response to DNA ...
    Jan 23, 2007 · Our observation that replication of damaged DNA is required in order to elicit RecA-GFP foci after UV treatment is consistent with these ...
  60. [60]
    A Bacterial G Protein-Mediated Response to Replication Arrest
    To define factors in E. coli promoting survival to replication fork stress, we isolated insertion mutants sensitive to replication inhibitors.
  61. [61]
    Replication and segregation of an Escherichia coli chromosome ...
    Jun 13, 2011 · coli harbors a single circular 4.6-Mbp chromosome within which replication initiates once per generation from a single replication origin, and ...Missing: frequency | Show results with:frequency
  62. [62]
    Eukaryotic Origin-Dependent DNA Replication In Vitro Reveals ...
    Eukaryotic DNA replication initiation occurs in two steps: helicase loading in G1, and activation by S-CDK and DDK in S phase.
  63. [63]
    Emerging players in the initiation of eukaryotic DNA replication - PMC
    Oct 17, 2012 · Therefore, the dynamic oscillating pattern ensures that pre-RC is assembled only in G1 phase and is strictly inhibited outside G1 phase, serving ...
  64. [64]
    Review Regulation of Early Events in Chromosome Replication
    CDKs inhibit pre-RC assembly by directly inhibiting each pre-RC component. The Cdc6 protein accumulates during early G1 phase, but is not present from late G1 ...
  65. [65]
    preventing rereplication via multiple mechanisms in eukaryotic cells
    In the G1 phase, origins of DNA replication are rendered competent when the core of the helicase, the MCM2–7 complex, is assembled into pre-RCs. In S phase, ...
  66. [66]
    Replication timing and its emergence from stochastic processes - PMC
    The reproducible replication timing patterns seen in population studies can be explained by stochastic origin firing at the single-cell level.
  67. [67]
    Genome-wide studies highlight indirect links between human ...
    Oct 14, 2008 · We systematically mapped replication origins along 1% of the human genome in HeLa cells. We identified 283 origins, 10 times more than previously known.
  68. [68]
    Integrative analysis of DNA replication origins and ORC-/MCM ...
    Apr 3, 2024 · In each human cell cycle, replication starts from~50,000 genomic locations called replication origins (Hu and Stillman, 2023). At an origin of ...
  69. [69]
    Chk1 inhibits replication factory activation but allows dormant origin ...
    This allows the activation of dormant origins as a consequence of stochastic origin firing within existing replication factories where forks are inhibited ( ...
  70. [70]
    Stochastic association of neighboring replicons creates replication ...
    Sep 23, 2013 · Upon replication stress or DNA damage, a replication factory defines an important boundary, inside of which dormant origins can initiate ...
  71. [71]
    Cell Cycle Regulation of DNA Replication - PMC - PubMed Central
    Eukaryotic DNA replication is regulated to ensure all chromosomes replicate once and only once per cell cycle. Replication begins at many origins scattered ...
  72. [72]
    DNA Replication and Checkpoint Control in S Phase - Nature
    Each time the DNA fork opens, the leading strand can be elongated, and a new Okazaki fragment is added to the lagging strand. All Okazaki fragments are ...Missing: length | Show results with:length
  73. [73]
    Telomere maintenance and DNA replication: how closely are these ...
    The maintenance of the DNA at chromosome ends, the telomeres, depends on conventional semiconservative replication and on the action of telomerase, ...
  74. [74]
    DNA replication and replication stress response in the context of ...
    Dec 6, 2023 · Here, we review known mechanisms and factors orchestrating replication initiation, and replication fork progression upon stress.Missing: 2020s | Show results with:2020s
  75. [75]
    Hallmarks of DNA Replication Stress - PMC - NIH
    This review discusses the major sources of DNA replication stress, the direct and indirect effects of replication stress on the genome, and the assays to ...Missing: 2020s | Show results with:2020s
  76. [76]
  77. [77]
    DNA Replication—A Matter of Fidelity - ScienceDirect.com
    Jun 2, 2016 · DNA Replication—A Matter of Fidelity. Author links open overlay panel ... Main Text. Introduction. An appropriate level of fidelity during ...
  78. [78]
    High-accuracy lagging-strand DNA replication mediated by ... - PNAS
    Apr 2, 2018 · Our study demonstrates that the two strands are copied with different accuracies. Specifically, the discontinuous lagging strand is significantly more accurate.
  79. [79]
    DNA Repair - Molecular Biology of the Cell - NCBI Bookshelf
    Figure 5-47. Depurination and deamination. These two reactions are the most frequent spontaneous chemical reactions known to create serious DNA damage in cells.Missing: tautomerization | Show results with:tautomerization
  80. [80]
    Genome-wide mapping of spontaneous DNA replication error ...
    Dec 11, 2024 · This study provides a comprehensive genome-wide map of replication error hotspots, offering a holistic perspective on the intricate interplay between various ...
  81. [81]
    Unravelling roles of error-prone DNA polymerases in shaping ...
    Oct 18, 2021 · This review summarizes the range of experimental approaches that have investigated and connected DNA polymerases with specific mechanisms of mutagenesis.Missing: paper | Show results with:paper
  82. [82]
    Molecular basis for proofreading by the unique exonuclease domain ...
    Dec 14, 2023 · DNA polymerase 3′→5′ exonuclease activity: different roles of the beta hairpin structure in family-B DNA polymerases. DNA Repair 29, 36–46 (2015) ...
  83. [83]
    The proofreading mechanism of the human leading-strand DNA ...
    The eukaryotic leading-strand DNA polymerase ε (Polε) is a dual-function enzyme with a proofreading 3′-5′ exonuclease (exo) site located 40 Å from the DNA ...
  84. [84]
    Strand Discrimination in DNA Mismatch Repair - PMC - NIH
    As for methyl-directed MMR, neither the recruitment of MutS homologs nor recruitment of MutL homologs seem likely mediate strand discrimination, as described ...
  85. [85]
    Mechanisms and functions of DNA mismatch repair | Cell Research
    Dec 24, 2007 · DNA is threaded through the protein complex until the latter reaches a strand discrimination signal in either orientation, a process that forms ...
  86. [86]
    Base selection, proofreading, and mismatch repair during DNA ...
    The results show that base selection discriminates against errors by 200,000-2,000,000-fold, proofreading by 40-200-fold, and mismatch repair by 20-400-fold, ...
  87. [87]
    DNA replication and mismatch repair safeguard against metabolic ...
    May 22, 2017 · Proofreading properties of these polymerases improve their fidelity by 10- to 100-fold, and postreplicative DNA mismatch repair (MMR), which ...
  88. [88]
    Lynch Syndrome (Hereditary Nonpolyposis Colorectal Cancer) - NCBI
    Jun 8, 2024 · Rarely an individual is not born with Lynch syndrome but develops mutations in both alleles, leading to the somatic loss of the MMR proteins.[5] ...Continuing Education Activity · Introduction · Evaluation · Treatment / Management
  89. [89]
    Base Excision Repair - PMC - NIH
    Base excision repair (BER) corrects DNA damage from oxidation, deamination and alkylation. Such base lesions cause little distortion to the DNA helix structure.
  90. [90]
    Catalytic and noncatalytic functions of DNA polymerase κ in ... - Nature
    Sep 19, 2024 · Translesion DNA synthesis (TLS) is a cellular process that enables the bypass of DNA lesions encountered during DNA replication and is ...
  91. [91]
    The expanding cellular functions of translesion DNA polymerases
    Oct 19, 2023 · Translesion synthesis (TLS) DNA polymerases were originally described as error-prone enzymes involved in the bypass of DNA lesions.
  92. [92]
    Mechanisms of Oncogene-Induced Replication Stress
    Oncogene-induced replication stress is an early driver of genomic instability and is attributed to a plethora of factors, most notably aberrant origin firing.
  93. [93]
    Fanconi anemia proteins stabilize replication forks - PMC - NIH
    We show that FANCL depletion results in defective DNA replication restart following treatment with camptothecin, a drug that results in DSBs during DNA ...Missing: paper | Show results with:paper
  94. [94]
    Origins of DNA replication | PLOS Genetics - Research journals
    Sep 12, 2019 · In all kingdoms of life, DNA is used to encode hereditary information. Propagation of the genetic material between generations requires ...
  95. [95]
    Recent Advancements in Reducing the Off-Target Effect of CRISPR ...
    Jan 18, 2024 · The CRISPR-Cas9 systems' off-target effect mainly results from the generation of DNA double-strand breaks (DBSs) at target genomic loci, the ...Missing: stability 2020s
  96. [96]
    Simian virus 40 DNA replication in vitro. - PNAS
    In vitro, SV40 DNA replication uses extracts from infected cells, with the origin of replication, and proceeds via branched intermediates, resembling in vivo  ...
  97. [97]
    Exponential propagation of large circular DNA by reconstitution of a ...
    Sep 28, 2017 · In this study, we have reconstituted the E. coli chromosome replication cycle using purified components, providing a powerful enzymatic tool for ...
  98. [98]
    Cryo-EM structure of the replisome reveals multiple interactions ...
    Feb 21, 2017 · The “trombone” model of DNA replication postulates that the lagging strand forms a loop such that the leading- and lagging-strand ...Missing: original | Show results with:original
  99. [99]
    Mechanisms and regulation of DNA replication initiation in eukaryotes
    The large number of origins and the need to coordinate initiation across these sites represents a fundamental challenge to DNA replication in eukaryotes; other ...
  100. [100]
    Kary B. Mullis – Facts - NobelPrize.org
    Kary Mullis invented the process known as polymerase chain reaction (PCR), in which a small amount of DNA can be copied in large quantities over a short period ...Missing: primary | Show results with:primary
  101. [101]
    Polymerase Chain Reaction (PCR) - StatPearls - NCBI Bookshelf - NIH
    Jul 7, 2025 · The process consists of three major phases: denaturation, annealing, and extension. During denaturation, DNA is heated to 95°C to disrupt ...Introduction · Testing Procedures · Results, Reporting, and... · Clinical SignificanceMissing: original | Show results with:original
  102. [102]
    A beginner's guide to RT-PCR, qPCR and RT-qPCR | The Biochemist
    Jun 15, 2020 · RT-PCR uses RNA to generate cDNA, qPCR amplifies DNA in real-time, and RT-qPCR combines both for RNA level measurement.Missing: source | Show results with:source
  103. [103]
    PCR in Forensic Science: A Critical Review - PMC - NIH
    Mar 29, 2024 · This review examines the evolution of the PCR from its inception in the 1980s, through to its current application in forensic science.
  104. [104]
    Principles and applications of polymerase chain reaction in medical ...
    The aim of the present study is to outline the principles and applications of conventional PCR and Real-Time PCR techniques in some medical sciences.
  105. [105]
    Error Rate Comparison during Polymerase Chain Reaction by DNA ...
    Given the amplification efficiency of each PCR reaction, the error rate (average of 2 experiments) for Taq polymerase is 4.3 × 10−5 ± 1.8 mutations/bp per ...
  106. [106]
    PCR Fidelity of Pfu DNA Polymerase and Other Thermostable DNA ...
    Under these conditions, the error rate of exo −Pfu was ∼40-fold higher (5 × 10 −5 ) than the error rate of Pfu . As the reaction pH was raised from pH 8 ...