Single-Stranded DNA Binding Protein in DNA Metabolism: Comparison
Please note this is a comparison between Version 2 by Lindsay Dong and Version 1 by Longfu Xu.

Single-stranded DNA-binding proteins (SSBs) play vital roles in DNA metabolism. Proteins of the SSB family exclusively and transiently bind to ssDNA, preventing the DNA double helix from re-annealing and maintaining genome integrity. In the meantime, they interact and coordinate with various proteins vital for DNA replication, recombination, and repair. Although SSB is essential for DNA metabolism, proteins of the SSB family have been long described as accessory players, primarily due to their unclear dynamics and mechanistic interaction with DNA and its partners. Recently-developed single-molecule tools, together with biochemical ensemble techniques and structural methods, have enhanced our understanding of the different coordination roles that SSB plays during DNA metabolism. 

 

  • single-stranded DNA-binding proteins
  • single-molecule technique
  • DNA replication
  • DNA repair
  • DNA recombination

1. Introduction

Central to many genome-maintenance machineries are single-stranded DNA binding proteins (SSBs). These SSB proteins play a vital role in the maintenance of genomes by binding exclusively and transiently to ssDNA intermediates during DNA replication, recombination, and repair. By further interacting with different proteins crucial to all aspects of genome maintenance and recruiting them to their targets on DNA, the SSB protein plays a prominent role in bridging genome maintenance pathways and modulating their activity. Due to their interaction with DNA, they influence many other downstream processes, which include all the possible protein-mediated biological functions. Biochemical studies have demonstrated that SSB plays an essential role in DNA metabolism. However, the real-time interaction dynamics between SSB with DNA and its partner proteins have proven elusive owing to the limited averaged population and time resolution. The recent development of single-molecule assays, in combination with robust ensemble biochemical techniques and structural methods, have contributed significantly to our understanding of the molecular mechanisms of SSB, interaction dynamics with other protein partners, and the mechanistic interactions with partner proteins.

2. Classification of SSB

2.1. Properties of SSB

As suggested by the name, single-stranded DNA binding proteins bind to ssDNA. While this is mostly understood in the context of preventing re-annealing during lagging strand synthesis or during DNA damage repair, SSB proteins are vital in many other processes. All proteins of this broad class interact with single-stranded DNA, and some also interact with dsDNA [1,2,3,4,5,6,7][1][2][3][4][5][6][7].

2.2. Classification of SSB

As far as is known, all classified organisms with available genomes encode SSBs, suggesting that the role they play is essential to life processes at a fundamental level [8]. The role of SSBs is most saliently communicated as their role in replication, to prevent the re-annealing of single-stranded DNA, so that the template strand can be copied. The SSBs differ significantly from one another, and variations between different kingdoms of life trump intra-kingdom differences [9].

3. Single-Molecule Toolbox to Study SSB

3.1. Single-Molecule Force Studies of SSB–ssDNA Interactions

Generally, single-molecule techniques fall into two classes, namely those that measure force, displacement, and torque, and those that detect fluorescence. The first category, called single-molecule force spectroscopy, has become increasingly important for understanding the tensions, motions, and torques associated with biological molecules and their enzymatic activity. Studies have been conducted using single-molecule force spectroscopy to determine the interaction between different SSBs with dsDNA, ssDNA, or both [1,2[1][2][3][4][5][6][7],3,4,5,6,7], attempting to examine, for example, whether SSBs can destabilize duplex DNA. With optical tweezers, DNA molecules have been manipulated to investigate the kinetics and thermodynamics of the binding of T7 SSB (gp2.5) and T4 SSB (gp32) to dsDNA and ssDNA [4,5,6,7][4][5][6][7]. An optical tweezer assay involves attaching one end of a DNA molecule to an optically trapped bead. On the other end, one of the following methods is employed: a micropipette ([41][10]), the surface of a microfluidic device ([42,43,44][11][12][13]), or a second optically trapped bead (as referenced in [45,46,47,48][14][15][16][17]) which is commonly referred to as a dual-trap optical tweezers setup. To study the effect of SSB on DNA molecules, double-stranded DNA is usually melted by force to obtain ssDNA [49][18]. Alternatively, SSB can be directly observed destabilizing duplex DNA. In the latter experiments, the dsDNA melting force was monitored in relation to the SSB concentration and pulling rate measured by elongation of end-to-end distance for the trapped DNA per time unit; Additionally, models were created to calculate the size of the SSB binding sites, referred to as the “footprint size”, as well as the association rates and equilibrium dissociation constants (KD) of SSB proteins binding to both single-stranded and double-stranded DNA [18,28,29][19][20][21]. Another commonly used assay to study SSB at the single-molecule level is the magnetic tweezer. In a magnetic tweezer assay, a biomolecule is tethered to a micron-sized superparamagnetic bead and a microchannel surface through antigen–antibody interactions. The corresponding force applied to biomolecules can be calibrated by analyzing the Brownian motion of the beads obtained through the bright-field images [53,54,55,56][22][23][24][25]. The relevant distance between the magnetic bead and the surface is determined by measuring the change in the diffraction pattern of the bead with respect to the magnet height [53,54,55,56][22][23][24][25]. By varying the magnet strength and the experimental design, forces of between 0.001 and 100 pN can typically be achieved [54,55,57][23][24][26]. When compared with optical tweezers to study SSB, which is often limited by its lower throughput, magnetic tweezers allow many single DNA molecules to be tethered with separate beads and probed in parallel, thus, achieving high throughput of data collection. Example studies using magnetic tweezers were to determine whether the gp32 and E. coli SSB proteins could prevent DNA strand rezipping [2,3][2][3].

3.2. Image Measurement of SSB-ssDNA Complex

Another significant category of single-molecule tools is based on the detection of fluorescence. Several single-molecule fluorescence approaches have proven to be particularly useful for studying the SSB–DNA complex, for which smFRET has provided a high-resolution dynamic picture of how SSB interacts with ssDNA. The smFRET technique involves the use of two fluorescent dyes, which are covalently attached to specific locations within the DNA molecule or its interacting protein. The smFRET assay can be performed using either confocal microscopy of freely diffusing molecules or TIRF microscopy of molecules attached to surfaces [64][27]. The smFRET measurements are frequently combined with other stretching techniques to provide a more comprehensive understanding of the DNA-protein interactions. When the distance between the two fluorophores is short (usually less than 10 nanometers), the donor transfers energy without radiation to the acceptor, resulting in the emission of fluorescence by the acceptor instead of the donor [65,66][28][29]. The FRET efficiency, defined as the efficiency of energy transfer from the donor to the acceptor, depends upon the proximity between the two fluorophores; therefore, it can be used to measure shifts in the distance up to ~10 nm. This tool is excellent for tracking real-time conformational and relative position changes within single biological molecules. The generated data, by measuring the dynamic states in a molecular system, can be quite different depending on the investigated system. Data fitting in smFRET data analysis is critical in understanding molecular dynamics and, thus, should be adopted based on the research question. Classic examples of smFRET to study the SSB–ssDNA complex are direct demonstrations of E. coli SSB in its (SSB)65 binding mode diffusing along ssDNA [67][30], which is consistent with early ensemble studies [68,69,70,71][31][32][33][34].

3.3. Hybrid Single-Molecule Tools

Besides the independent use of single-molecule force spectroscopy and fluorescence microscopy, combined force manipulation and fluorescence visualization have been extensively exploited to probe the binding dynamics of SSB to ssDNA and its interaction with other protein partners. These combined approaches are instrumental in understanding, for example, DNA–binding protein interactions that are sequence-dependent [46][15], for monitoring protein translocation along DNA [52[35][36][37],73,74], and for examining the relationship between protein binding and the mechanical properties of DNA [7,32,75][7][38][39].

3.4. Example Output of Single-Molecule Studies

From single-molecule experiments, several parameters can be extracted. For basic properties, there are the binding properties, which include the binding rate constants kon and koff, corresponding to the on rates and off rates, respectively. In addition to binding properties, one can determine stoichiometries to gain insights into the binding footprint of an individual SSB binding event. Information on the kinetics of binding can also be obtained, such as the presence of one, two, or multistage binding. As well as measuring these parameters, one may also measure their dependence on experimental conditions, such as ionic concentration, template tension, temperature, and pH.

Once bound, an important value is the diffusion constant, which can determine if the SSB is stationary or diffusive (such as E. coli SSB [67][30]). It can also test cooperativity through the concentration-dependent binding kinetics. Lastly, single-molecule techniques also allow for the direct observation of interactions between SSBs and other proteins, such as the role of T7 gp2.5 in replisome coordination. These different parameters come together to describe the system of SSB interactions with DNA and with other proteins.

Additionally, it can calculate maximal coating densities, which can be used to calculate the binding footprint. These can be determined by finding an association between DNA length shortening and fluorescence intensity. This is typically linear, as one SSB will induce a near-constant contour length change by bending nucleotides in its OB fold or wrapping the DNA around a tetramer in the case of E. coli [30][40]. The shortening of the DNA will be directly correlated to the fluorescence intensity, which serves as a proxy for the number of SSB bound. When the system is saturated, one can calculate the total contour length change and divide it by the number of SSB bound (found via fluorescence intensity). This allows one to calculate the length change per bound SSB, which is an important quantity in understanding the binding mechanism.

Single-molecule experiments provide insights into the binding footprint of SSB (single-stranded DNA binding) proteins by analyzing the maximal occupancy, which is the point where an increase in fluorescent intensity stops, even with an increase in concentration along the bound DNA. By determining the number of SSB proteins, the average linear occlusion of DNA per bound SSB can be calculated by dividing the number of nucleotides by the number of SSBs. However, it should be noted that the average linear occlusion is not equivalent to the binding footprint, as some SSBs bind disorderedly. Interestingly, the SSB of phi 29 binds in a consistent manner, forming a “unit cell” with a nearly constant spacing of 3.4 ± 0.3 nucleotides per phi 29 SSB monomer [80][41].

Thermodynamic aspects of binding can be determined through bulk methods, such as isothermal titration calorimetry (ITC) or melting experiments, but they can also be probed by the single-molecule techniques. At a rough level, it is possible to determine the binding and unbinding as a function of force. Most proteins will be evicted from DNA held at high tension. Another glimpse into the thermodynamics and binding mode is by measuring the saturation dependence of certain binding parameters. Since salts shield the negative charges along the DNA backbone, along with certain amino acid sequences, such as the C-terminal tail of gp2.5 [23][42] as well as E. coli [22][43], information about the electrostatics of DNA–protein interactions can be garnered by varying the salt concentration of monovalent, bivalent, and polyvalent anions. Monovalent anions are much less effective at shielding, even when normalized per unit charge than bi- or polyvalent anions. Bivalent anions pack double the charge in a more compact volume [82][44], allowing it to come close to the DNA or protein to screen the negative charge [83,84,85][45][46][47]. The change in binding properties as a function of mono-, bi-, and polyvalent anions may possibly be analyzed to determine the relevant length scales of the electrostatic interaction and the allosteric exclusion emerging from the close contact of the protein with DNA.

4. Examine the Interaction between ssDNA with SSB

4.1. General Binding Dynamics of SSB

SSB proteins play a critical role in binding ssDNA. The mode of SSB binding can vary, with some forming multimers, exhibiting cooperativity or exhibiting strong periodicity (as seen in phi29 [88][48]). Single-molecule experiments offer a deeper understanding of SSB behaviour, allowing the determination of binding and unbinding constants and the exploration of factors such as sequence dependence, DNA-conformation dependence, and conditions such as salt concentrations, temperature, pH, crowding agents, protein concentration, and the presence of co-factors. Additionally, single-molecule methods enable the investigation of multiple binding modes, which can be challenging to study in ensemble assays. Stoichiometries can provide insight into the binding footprint of SSBs and further our understanding of the structure of these proteins. When bound, one can measure the diffusion and lifetime of the SSB. Diffusion can be characterized by the diffusion constant, but it is also helpful to determine if there is a directional bias to SSB motion. Interactions with other proteins can also be studied, such as colocalization, assisted binding, and the potential impact on the function of other enzymes.

4.2. Binding Dynamics of SSB to ssDNA under Tension

In addition to the general binding properties of SSB proteins discussed in Section 4.1, which can be probed with both bulk assay and single-molecule studies, the binding dynamics of SSBs to ssDNA under tension can be studied exquisitely with the single-molecule tools, such as by using optical tweezers [16,32][38][49] and magnetic tweezers [35][50]. The force-dependence of binding often depends on the binding mode, which varies between SSBs, from the monomeric binding in an OB-fold by T7 gp2.5 SSB [95][51] to the wrapping of DNA by E. coli SSB [30][40]. As E. coli SSB is highly sensitive to force, the unwrapping of the DNA from E. coli SSB begins at tensions as low as 1 pN, and complete dissociation occurs between 7 and 12 pN [33][52]. The binding dynamics for T7 gp2.5 have been investigated at different pulling timescales to investigate the prevention of secondary structure formation and the impact of T7 gp2.5 binding on the energetics of DNA stretching. The experiment observed a clear shortening with the addition of T7 gp2.5, and by varying the speed, it could limit the number of SSBs binding [6]. It was determined that under the fast-pulling regime, fewer SSB bind, and the force relationship was similar to that of naked DNA. The real-time dynamics of SSB binding have also been investigated via high-speed AFM imaging [37][53]. The E. coli SSB binds, diffuses, and dissociates, and this is shown in real-time with AFM imaging. In the emerging high-speed AFM technique, high-resolution images of the sample can be obtained in a fully hydrated state, thus, allowing millisecond-scale visualization of the nanoscale dynamics of the system. The buffer conditions, such as cation types, concentration, and pH, as well as the length of the substrate, can be varied in order to gain a better understanding of how environmental factors affect binding dynamics.

4.3. Movement of SSB on ssDNA Probed with Single-Molecule Approaches

The distinction between diffusive and non-diffusive proteins is important. It is possible that diffusive proteins can cover a larger effective footprint (i.e., preventing secondary structure formation in this linear region of DNA). The SSB diffusion is passive and is thought to be driven largely by thermal motions. The diffusion of E. coli SSB has been observed in experiments using smFRET [67][30]. The DNA was labelled with donor and acceptor fluorophores located 69 nt apart, such that when the SSB was bound, there was a fluorescent signal produced. It was found that there is free diffusion of the SSB along the DNA. It is possible to ‘lock’ the E. coli SSB by forming a duplex structure with the bases at the 3′ and 5′ ends of the SSB–DNA complex. Additional experiments were conducted to test two distinct diffusion modes of E. coli SSB, which differ by the relative motions of the SSB with the DNA. The first mode, rolling, involves the DNA at the 5′ or 3′ ends lengthening or shortening by moving around the SSB tetramer. In this case, the relative position of a given DNA base and a given spot on the SSB tetramer do not move in relation to one another. The other diffusion model is that of sliding, where the ssDNA moves in relation to a fixed spot on the SSB tetramer. Experimental results support the sliding mechanism, as the site of the DNA FRET tag does not alter the FRET intensity pattern.

4.4. Sequence-Dependent Properties of SSB

For a nucleic acid binding protein, it is assumed that its interaction with the template is largely non-specific and that sections of the template are largely interchangeable and homogenous, except in the case of specific binding sequences, for example, such as Kozak sequences for translation initiation in eukaryotes [98][54]. While the assumption of a largely homogenous polymer may be useful in some applications, bases are often processed differently in important ways. Hairpins of GC-rich DNA require more force to unfold, but also have faster kinetics than AT-rich DNA [99][55]. GC-rich regions form a more stable secondary structure than AT-rich regions and more quickly. T4 gp32 and E. coli SSB proteins both act through the inhibition of refolding [2], so sequence manifests itself due to the different timescales of folding of GC-rich versus AT-rich hairpins. In the related case of mRNA translation, hairpins often slow and stall the ribosome, as the ribosome must resolve the secondary structure before proceeding [100][56], as the entry tunnel only allows ssRNA [101][57].

5. Coordination Role of SSB in DNA Metabolism

5.1. Overview of Single-Molecule Studies on SSB Interacting with Helicase

Helicases carry out many essential genome maintenance processes within the cell, such as replication, recombination, and repair [102,103,104,105,106][58][59][60][61][62]. It has been reported that the same helicase can carry out several of these functions [107,108,109][63][64][65]. Helicase activity must, therefore, undergo strict regulation. While there is no clear evidence regarding how this regulation occurs, growing evidence indicates that interactions with protein partners may be one of the mechanisms involved [105,107,108,109][61][63][64][65]. On the other hand, DNA helicases function to unwind dsDNA into ssDNA intermediate or to translocate along ssDNA, suggesting a frequent encounter with protein binding to ssDNA, such as SSBs, during various DNA processing events. Single-strand binding proteins have been shown to improve the unwinding efficiency of many helicases [110,111,112,113][66][67][68][69].

5.1.1. Interplay with Replicative Helicase CMG Complex

Replicative DNA helicases play an essential role in duplicating the genome in every cell cycle. Replicative DNA helicases are usually protein complexes with multi-subunit structures, such as the replicative helicase of eukaryotes, which is composed of 11 subunits and requires 2 subcomplexes and 1 protein to function. This heterohexameric helicase, the Cdc45-Mcm2-7-GINS (CMG) complex, is initiated through the formation of a complex with Cdc45 and the heterotetrameric GINS complex [38][70]. This CMG complex translocates in the direction of 3′–5′ along the leading-strand template and unwinds DNA at the replication fork powered by ATP hydrolysis [38,114][70][71]. In vitro single-molecule studies reveal that translocation on ssDNA of the yeast CMG helicase shows a rate at 5–10 bp s−1[115][72], while the observed dsDNA unwinding rate to be 0.1–0.5 bp s−1, possibly slowed by a frequent long-lived pausing state [116,117][73][74].

5.1.2. Interplay with Recombinational Repair Helicase XPD

Xeroderma pigmentosum group D (XPD) helicase belongs to subfamily 2B of helicases, including yeast Rad3 and human FANCJ, CHLR1, and RTEL [105,109,123,124[61][65][75][76][77],125], and is involved in a variety of DNA repair pathways. The XPD helicase mutation can affect nucleotide excision repair (NER) [126][78]. Human XPD is also associated with the transcription factor IIH and plays a significant role in the repair of nucleotide excisions [127,128,129,130][79][80][81][82]. Further evidence indicates that it also plays a role in chromosome segregation [131][83] and defence against retroviral infection [132][84]. While functioning on single-stranded DNA, XPD is likely to come into contact with other proteins, such as cognate SSB replication protein A (RPA). When the XPD encounters a bound RPA, it can bypass the RPA without dislodging it or facilitating its dissociation.

5.1.3. Interplay with Recombinational Repair Helicase RecQ

In the case of recombinational repair, another well-studied example of helicase is RecQ. During recombinational repair, RecQ plays a role in repairing ssDNA gaps and dsDNA breaks in E. coli after the primary repair pathway, RecBCD, is inactivated [137][85]. It also has been demonstrated that RecQ is responsible for suppressing the production of illegitimate recombinants [138[86][87],139], resolving replication fork stalled events [140][88] and stimulating the SOS response in Escherichia coli. [141,142,143,144][89][90][91][92]. Additionally, E. coli defective in RecQ is susceptible to ultraviolet light, resulting in a decrease in the frequency of recombination that leads to impaired cell growth and death [142,145][90][93]. Single-molecule approaches have been used, complementary to bulk biochemical and structural tools, to analyze the interactions of single molecules with high spatial and temporal resolutions and reveal dynamic heterogeneities missing in ensemble experiments due to population averaging [146,147,148,149,150,151,152,153][94][95][96][97][98][99][100][101].

5.1.4. Interplay with Replication Restart Helicase PriA

The DNA replication protein complex can be dissociated before replication is completed by collisions with damaged DNA or immovable protein barriers [122,154,155,156,157][102][103][104][105][106]. Cells can resolve this potentially lethal problem by reloading the replisome by activating “replication restart” reactions [157][106]. In bacteria, the PriA DNA helicase orchestrates this vital activity by binding to structure-specific DNA and interacting with replication-associated SSBs [158,159,160][107][108][109].

5.2. SSB Interacting with Replicative DNA Polymerase during Primer Extension

Bulk studies have shown that the presence of SSB significantly enhances DNA replication in vitro [19,166,167,168,169,170][110][111][112][113][114][115]. This enhancement may be attributed to the multiple roles that SSBs perform [19,30,171][40][110][116], including the prevention of degradation of ssDNA, the removal of secondary structures, the increase in recognition and initiation of primers, a decrease in non-specific DNA polymerase binding to the template, and an increase in DNA polymerase’s activity in displacing strands and extending primers [19,30,166,167,168,169,170,171][40][110][111][112][113][114][115][116]. However, it remains unclear whether the seemingly conflicting roles of polymerase and SSB on ssDNA can be coordinated during the lagged strand replication.

5.3. Single-Molecule Studies on SSB with DNA Polymerase during Strand Exchange

The effect of SSB on strand displacement DNA replication has also been examined using single-molecule tools [176,177][117][118]. It was previously demonstrated that strand displacement DNA synthesis, such as by Polγ, accomplishes replication by utilizing stable secondary structures [178[119][120],179], ensuring that the D-loop DNA structure is maintained at the origin of heavy strands [180][121], and removing primers through the coordination of primer processing factors [181,182,183][122][123][124]. However, the efficiency of Polγ is limited to a few nucleotides [182[123][125][126][127][128],184,185,186,187], in accordance with other DNA polymerases involved in strand displacement synthesis [176,188,189][117][129][130].

5.4. Single-Molecule Studies of SSB Interplay with Recombinase

During DNA metabolism, replication forks can stall or collapse, resulting in extensive single-strand gaps [194,195,196,197][131][132][133][134]. Consequently, the SSB protein binds to the ssDNA in these gaps, preventing other proteins from accessing the ssDNA. The RecA protein from E. coli is essential to repair broken DNA and maintain genomic integrity through homologous recombination. In order to function, RecA filaments are required to nucleate and grow on single-stranded DNA concurrently with SSB, which sequesters ssDNA continuously and thereby causes it to compete with and prevent RecA assembly [198,199][135][136]. Because of the complexity resulting from dynamic competition with SSB during self-assembly on ssDNA lattices, our knowledge of RecA filament assembly and its role in DNA recombination has been compromised. Despite extensive and varied efforts, ensemble measurements based on an averaged population are not able to distinguish between nucleation and growth in a reliable manner [198,199,200,201][135][136][137][138].

5.5. Chemo-Mechanical Pushing of

E. coli

SSB by a Translocating Protein Partner

Both bulk and single-molecule assays have shown that SSB binds exclusively to ssDNA with very high (pM to fM) affinities [32,33,71,217][34][38][52][139]; however, these tightly bound complexes must be displaced, bypassed, or redistributed along ssDNA to complete replication, recombination, and repair. As discussed in Section 5.1.2 [34][140], a pre-bound RPA on ssDNA can be dislodged by a translocating helicase XPD or bypassed without dissociation. The PriA binding to E. coli SSB can modulate the binding mode from SSB65-to-SSB35 to expose more ssDNA for DNA replication restart. Considering the diffusive property of E. coli SSB, one other potential mechanism for reorganization can be pushed along ssDNA by a translocating protein. The DNA translocases are motor proteins capable of translocating ssDNA at high rates powered by the hydrolysis of ATP [107,218,219][63][141][142]. To delve into the impact of a directional translocase encountering an E. coli SSB tetramer bound to single-stranded DNA, a smFRET assay was utilized to detect such pushing events [220][143]. A fluctuating FRET signal is observed when Cy5-labeled E. coli SSB is bound to surface-immobilized 3′-Cy3–labelled ssDNA, indicating that SSB is randomly diffusing on ssDNA. When adding Saccharomyces cerevisiae Pif1, a translocase for ssDNA 5′ to 3′, irregular-spaced saw-tooth FRET spikes are observed with ATP.

6. Conclusions

As transient and exclusive binders to ssDNA intermediates, SSBs are crucial for genome maintenance. Further interacting with various proteins vital for DNA maintenance, the SSB protein bridges genome maintenance pathways and modulates their activity by recruiting them to their DNA sites of action. Although it has been demonstrated that SSB is essential for DNA metabolism, the dynamics of SSB interaction with DNA and its partners remain unclear. Recent developed single-molecule assays have provided essential insights into the molecular mechanisms of SSB and interacting dynamics with other protein partners. By combining robust ensemble biochemical techniques and structural methods, a more comprehensive understanding of these SSBs has been gained.

 

References

  1. De Vlaminck, I.; Vidic, I.; van Loenhout, M.T.J.; Kanaar, R.; Lebbink, J.H.G.; Dekker, C. Torsional regulation of hRPA-induced unwinding of double-stranded DNA. Nucleic Acids Res. 2010, 38, 4133–4142.
  2. Hatch, K.; Danilowicz, C.; Coljee, V.; Prentiss, M. Direct measurements of the stabilization of single-stranded DNA under tension by single-stranded binding proteins. Phys. Rev. E 2007, 76, 021916.
  3. Hatch, K.; Danilowicz, C.; Coljee, V.; Prentiss, M. Measurement of the salt-dependent stabilization of partially open DNA by Escherichia coli SSB protein. Nucleic Acids Res. 2007, 36, 294–299.
  4. Pant, K.; Karpel, R.L.; Rouzina, I.; Williams, M.C. Mechanical Measurement of Single-molecule Binding Rates: Kinetics of DNA Helix-destabilization by T4 Gene 32 Protein. J. Mol. Biol. 2004, 336, 851–870.
  5. Pant, K.; Karpel, R.L.; Rouzina, I.; Williams, M.C. Salt Dependent Binding of T4 Gene 32 Protein to Single and Double-stranded DNA: Single Molecule Force Spectroscopy Measurements. J. Mol. Biol. 2005, 349, 317–330.
  6. Shokri, L.; Marintcheva, B.; Eldib, M.; Hanke, A.; Rouzina, I.; Williams, M.C. Kinetics and thermodynamics of salt-dependent T7 gene 2.5 protein binding to single- and double-stranded DNA. Nucleic Acids Res. 2008, 36, 5668–5677.
  7. Shokri, L.; Marintcheva, B.; Richardson, C.C.; Rouzina, I.; Williams, M.C. Single Molecule Force Spectroscopy of Salt-dependent Bacteriophage T7 Gene 2.5 Protein Binding to Single-stranded DNA. J. Biol. Chem. 2006, 281, 38689–38696.
  8. Broderick, S.; Rehmet, K.; Concannon, C.; Nasheuer, H.-P. Eukaryotic Single-Stranded DNA Binding Proteins: Central Factors in Genome Stability. In Genome Stability and Human Diseases; Springer: Dordrecht, The Netherlands, 2009; Volume 50, pp. 143–163.
  9. Szczepankowska, A.K.; Prestel, E.; Mariadassou, M.; Bardowski, J.K.; Bidnenko, E. Phylogenetic and Complementation Analysis of a Single-Stranded DNA Binding Protein Family from Lactococcal Phages Indicates a Non-Bacterial Origin. PLoS ONE 2011, 6, e26942.
  10. Bryant, Z.; Stone, M.D.; Gore, J.; Smith, S.B.; Cozzarelli, N.R.; Bustamante, C. Structural transitions and elasticity from torque measurements on DNA. Nature 2003, 424, 338–341.
  11. La Porta, A.; Wang, M.D. Optical Torque Wrench: Angular Trapping, Rotation, and Torque Detection of Quartz Microparticles. Phys. Rev. Lett. 2004, 92, 190801.
  12. Lang, M.J.; Asbury, C.L.; Shaevitz, J.W.; Block, S.M. An Automated Two-Dimensional Optical Force Clamp for Single Molecule Studies. Biophys. J. 2002, 83, 491–501.
  13. Hohng, S.; Zhou, R.; Nahas, M.K.; Yu, J.; Schulten, K.; Lilley, D.M.J.; Ha, T. Fluorescence-Force Spectroscopy Maps Two-Dimensional Reaction Landscape of the Holliday Junction. Science 2007, 318, 279–283.
  14. Gross, P.; Farge, G.; Peterman, E.J.; Wuite, G.J. Combining Optical Tweezers, Single-Molecule Fluorescence Microscopy, and Microfluidics for Studies of DNA–Protein Interactions. Methods Enzymol. 2010, 475, 427–453.
  15. Schakenraad, K.; Biebricher, A.S.; Sebregts, M.; ten Bensel, B.; Peterman, E.J.G.; Wuite, G.J.L.; Heller, I.; Storm, C.; van der Schoot, P. Hyperstretching DNA. Nat. Commun. 2017, 8, 2197.
  16. van Mameren, J.; Modesti, M.; Kanaar, R.; Wyman, C.; Wuite, G.J.; Peterman, E.J. Dissecting Elastic Heterogeneity along DNA Molecules Coated Partly with Rad51 Using Concurrent Fluorescence Microscopy and Optical Tweezers. Biophys. J. 2006, 91, L78–L80.
  17. Greenleaf, W.J.; Woodside, M.T.; Abbondanzieri, E.A.; Block, S.M. Passive All-Optical Force Clamp for High-Resolution Laser Trapping. Phys. Rev. Lett. 2005, 95, 208102.
  18. Candelli, A.; Hoekstra, T.P.; Farge, G.; Gross, P.; Peterman, E.J.G.; Wuite, G.J.L. A toolbox for generating single-stranded DNA in optical tweezers experiments. Biopolymers 2013, 99, 611–620.
  19. Jose, D.; Weitzel, S.E.; Baase, W.A.; von Hippel, P.H. Mapping the interactions of the single-stranded DNA binding protein of bacteriophage T4 (gp32) with DNA lattices at single nucleotide resolution: gp32 monomer binding. Nucleic Acids Res. 2015, 43, 9276–9290.
  20. Lindner, C.; Nijland, R.; van Hartskamp, M.; Bron, S.; Hamoen, L.W.; Kuipers, O.P. Differential Expression of Two Paralogous Genes of Bacillus subtilis Encoding Single-Stranded DNA Binding Protein. J. Bacteriol. 2004, 186, 1097–1105.
  21. Frickey, T.; Lupas, A. CLANS: A Java application for visualizing protein families based on pairwise similarity. Bioinformatics 2004, 20, 3702–3704.
  22. Strick, T.R.; Allemand, J.-F.; Bensimon, D.; Croquette, V. The Elasticity of a Single Supercoiled DNA Molecule. Science 1996, 271, 1835–1837.
  23. Neuman, K.C.; Nagy, A. Single-molecule force spectroscopy: Optical tweezers, magnetic tweezers and atomic force microscopy. Nat. Methods 2008, 5, 491–505.
  24. Lipfert, J.; Hao, X.; Dekker, N.H. Quantitative Modeling and Optimization of Magnetic Tweezers. Biophys. J. 2009, 96, 5040–5049.
  25. Gosse, C.; Croquette, V. Magnetic Tweezers: Micromanipulation and Force Measurement at the Molecular Level. Biophys. J. 2002, 82, 3314–3329.
  26. Sarkar, R.; Rybenkov, V.V. A Guide to Magnetic Tweezers and Their Applications. Front. Phys. 2016, 4, 48.
  27. Hellenkamp, B.; Schmid, S.; Doroshenko, O.; Opanasyuk, O.; Kühnemuth, R.; Adariani, S.R.; Ambrose, B.; Aznauryan, M.; Barth, A.; Birkedal, V.; et al. Precision and accuracy of single-molecule FRET measurements—A multi-laboratory benchmark study. Nat. Methods 2018, 15, 669–676.
  28. Maleki, P.; Budhathoki, J.B.; Roy, W.A.; Balci, H. A practical guide to studying G-quadruplex structures using single-molecule FRET. Mol. Genet. Genom. 2017, 292, 483–498.
  29. Ha, T.; Enderle, T.; Ogletree, D.F.; Chemla, D.S.; Selvin, P.R.; Weiss, S. Probing the interaction between two single molecules: Fluorescence resonance energy transfer between a single donor and a single acceptor. Proc. Natl. Acad. Sci. USA 1996, 93, 6264–6268.
  30. Roy, R.; Kozlov, A.G.; Lohman, T.M.; Ha, T. SSB protein diffusion on single-stranded DNA stimulates RecA filament formation. Nature 2009, 461, 1092–1097.
  31. Kozlov, A.G.; Lohman, T.M. Stopped-Flow Studies of the Kinetics of Single-Stranded DNA Binding and Wrapping around the Escherichia coli SSB Tetramer. Biochemistry 2002, 41, 6032–6044.
  32. Kuznetsov, S.V.; Kozlov, A.G.; Lohman, T.M.; Ansari, A. Microsecond Dynamics of Protein–DNA Interactions: Direct Observation of the Wrapping/Unwrapping Kinetics of Single-stranded DNA around the E. coli SSB Tetramer. J. Mol. Biol. 2006, 359, 55–65.
  33. Roemer, R.; Schomburg, U.; Krauss, G.; Maass, G. Escherichia coli single-stranded DNA binding protein is mobile on DNA: Proton NMR study of its interaction with oligo- and polynucleotides. Biochemistry 1984, 23, 6132–6137.
  34. Kozlov, A.G.; Lohman, T.M. Kinetic Mechanism of Direct Transfer of Escherichia coli SSB Tetramers between Single-Stranded DNA Molecules. Biochemistry 2002, 41, 11611–11627.
  35. King, G.A.; Burla, F.; Peterman, E.J.G.; Wuite, G.J.L. Supercoiling DNA optically. Proc. Natl. Acad. Sci. USA 2019, 116, 26534–26539.
  36. Heller, I.; Sitters, G.; Broekmans, O.D.; Farge, G.; Menges, C.; Wende, W.; Hell, S.W.; Peterman, E.J.G.; Wuite, G.J.L. STED nanoscopy combined with optical tweezers reveals protein dynamics on densely covered DNA. Nat. Methods 2013, 10, 910–916.
  37. Brouwer, I.; Sitters, G.; Candelli, A.; Heerema, S.J.; Heller, I.; De, A.J.M.; Zhang, H.; Normanno, D.; Modesti, M.; Peterman, E.J.G.; et al. Sliding sleeves of XRCC4–XLF bridge DNA and connect fragments of broken DNA. Nature 2016, 535, 566–569.
  38. Zhou, R.; Kozlov, A.G.; Roy, R.; Zhang, J.; Korolev, S.; Lohman, T.M.; Ha, T. SSB Functions as a Sliding Platform that Migrates on DNA via Reptation. Cell 2011, 146, 222–232.
  39. Pant, K.; Karpel, R.L.; Williams, M.C. Kinetic Regulation of Single DNA Molecule Denaturation by T4 Gene 32 Protein Structural Domains. J. Mol. Biol. 2003, 327, 571–578.
  40. Antony, E.; Lohman, T.M. Dynamics of E. coli single stranded DNA binding (SSB) protein-DNA complexes. Semin. Cell Dev. Biol. 2019, 86, 102–111.
  41. Soengas, M.; Esteban, J.A.; Salas, M.; Gutierrez, C. Complex Formation between Phage Phi φ29 Single-stranded DNA Binding Protein and DNA. J. Mol. Biol. 1994, 239, 213–226.
  42. Ghosh, S.; Marintcheva, B.; Takahashi, M.; Richardson, C.C. C-terminal Phenylalanine of Bacteriophage T7 Single-stranded DNA-binding Protein Is Essential for Strand Displacement Synthesis by T7 DNA Polymerase at a Nick in DNA. J. Biol. Chem. 2009, 284, 30339–30349.
  43. Kozlov, A.G.; Weiland, E.; Mittal, A.; Waldman, V.; Antony, E.; Fazio, N.; Pappu, R.V.; Lohman, T.M. Intrinsically Disordered C-Terminal Tails of E. coli Single-Stranded DNA Binding Protein Regulate Cooperative Binding to Single-Stranded DNA. J. Mol. Biol. 2015, 427, 763–774.
  44. Clementi, E.; Raimondi, D.L.; Reinhardt, W.P. Atomic Screening Constants from SCF Functions. II. Atoms with 37 to 86 Electrons. J. Chem. Phys. 1967, 47, 1300–1307.
  45. Kriegel, F.; Ermann, N.; Forbes, R.; Dulin, D.; Dekker, N.; Lipfert, J. Probing the salt dependence of the torsional stiffness of DNA by multiplexed magnetic torque tweezers. Nucleic Acids Res. 2017, 45, 5920–5929.
  46. Anthony, P.C.; Sim, A.Y.L.; Chu, V.B.; Doniach, S.; Block, S.M.; Herschlag, D. Electrostatics of Nucleic Acid Folding under Conformational Constraint. J. Am. Chem. Soc. 2012, 134, 4607–4614.
  47. Bizarro, C.V.; Alemany, A.; Ritort, F. Non-specific binding of Na + and Mg 2+ to RNA determined by force spectroscopy methods. Nucleic Acids Res. 2012, 40, 6922–6935.
  48. Meijer, W.J.J.; Horcajadas, J.A.; Salas, M. φ29 Family of Phages. Microbiol. Mol. Biol. Rev. 2001, 65, 261–287.
  49. Suksombat, S.; Khafizov, R.; Kozlov, A.G.; Lohman, T.M.; Chemla, Y.R. Structural dynamics of E. coli single-stranded DNA binding protein reveal DNA wrapping and unwrapping pathways. eLife 2015, 4, e08193.
  50. Bagchi, D.; Zhang, W.; Hodeib, S.; Ducos, B.; Croquette, V.; Manosas, M. Magnetic Tweezers-Based Single-Molecule Assays to Study Interaction of E. coli SSB with DNA and RecQ Helicase. Methods Mol. Biol. 2021, 2281, 93–115.
  51. Hollis, T.; Stattel, J.M.; Walther, D.S.; Richardson, C.C.; Ellenberger, T. Structure of the Gene 2.5 Protein, a Single-Stranded DNA Binding Protein Encoded by Bacteriophage T7. Proc. Natl. Acad. Sci. USA 2001, 98, 9557–9562.
  52. Honda, M.; Park, J.; Pugh, R.A.; Ha, T.; Spies, M. Single-Molecule Analysis Reveals Differential Effect of ssDNA-Binding Proteins on DNA Translocation by XPD Helicase. Mol. Cell 2009, 35, 694–703.
  53. Shlyakhtenko, L.S.; Lushnikov, A.Y.; Miyagi, A.; Lyubchenko, Y.L. Specificity of Binding of Single-Stranded DNA-Binding Protein to Its Target. Biochemistry 2012, 51, 1500–1509.
  54. Kozak, M. Initiation of translation in prokaryotes and eukaryotes. Gene 1999, 234, 187–208.
  55. Woodside, M.T.; Behnke-Parks, W.M.; Larizadeh, K.; Travers, K.; Herschlag, D.; Block, S.M. Nanomechanical measurements of the sequence-dependent folding landscapes of single nucleic acid hairpins. Proc. Natl. Acad. Sci. USA 2006, 103, 6190–6195.
  56. Doma, M.K.; Parker, R. Endonucleolytic cleavage of eukaryotic mRNAs with stalls in translation elongation. Nature 2006, 440, 561–564.
  57. Kurkcuoglu, O.; Doruker, P.; Sen, T.; Kloczkowski, A.; Jernigan, R.L. The ribosome structure controls and directs mRNA entry, translocation and exit dynamics. Phys. Biol. 2008, 5, 046005.
  58. McGlynn, P. Helicases at the Replication Fork. Adv. Exp. Med. Biol. 2012, 767, 97–121.
  59. Daley, J.M.; Niu, H.; Sung, P. Roles of DNA Helicases in the Mediation and Regulation of Homologous Recombination. Adv. Exp. Med. Biol. 2012, 767, 185–202.
  60. Kuper, J.; Kisker, C. DNA Helicases in NER, BER, and MMR. Adv. Exp. Med. Biol. 2012, 767, 203–224.
  61. Stekas, B.; Yeo, S.; Troitskaia, A.; Honda, M.; Sho, S.; Spies, M.; Chemla, Y.R. Switch-like control of helicase processivity by single-stranded DNA binding protein. eLife 2021, 10, e60515.
  62. Jin, S.; Bueno, C.; Lu, W.; Wang, Q.; Chen, M.; Chen, X.; Wolynes, P.G.; Gao, Y. Computationally exploring the mechanism of bacteriophage T7 gp4 helicase translocating along ssDNA. Proc. Natl. Acad. Sci. USA 2022, 119, e2202239119.
  63. Lohman, T.M.; Tomko, E.J.; Wu, C.G. Non-hexameric DNA helicases and translocases: Mechanisms and regulation. Nat. Rev. Mol. Cell Biol. 2008, 9, 391–401.
  64. Wu, C.G.; Spies, M. Overview: What Are Helicases? Adv. Exp. Med. Biol. 2012, 767, 1–16.
  65. Beyer, D.C.; Ghoneim, M.K.; Spies, M. Structure and Mechanisms of SF2 DNA Helicases. Adv. Exp. Med. Biol. 2012, 767, 47–73.
  66. Umezu, K.; Nakayama, H. RecQ DNA Helicase of Escherichia coli: Characterization of the Helix-unwinding Activity with Emphasis on the Effect of Single-stranded DNA-binding Protein. J. Mol. Biol. 1993, 230, 1145–1150.
  67. Lee, J.; Chastain, P.; Griffith, J.D.; Richardson, C.C. Lagging strand synthesis in coordinated DNA synthesis by bacteriophage T7 replication proteins. J. Mol. Biol. 2002, 316, 19–34.
  68. Harmon, F.G.; Kowalczykowski, S.C. RecQ Helicase, in Concert with RecA and SSB Proteins, Initiates and Disrupts DNA Recombination. Genes Dev. 1998, 12, 1134–1144.
  69. Rajagopal, V.; Patel, S.S. Single Strand Binding Proteins Increase the Processivity of DNA Unwinding by the Hepatitis C Virus Helicase. J. Mol. Biol. 2008, 376, 69–79.
  70. Kose, H.B.; Xie, S.; Cameron, G.; Strycharska, M.S.; Yardimci, H. Duplex DNA engagement and RPA oppositely regulate the DNA-unwinding rate of CMG helicase. Nat. Commun. 2020, 11, 1–15.
  71. Fu, Y.V.; Yardimci, H.; Long, D.T.; Ho, T.V.; Guainazzi, A.; Bermudez, V.P.; Hurwitz, J.; van Oijen, A.; Schärer, O.D.; Walter, J.C. Selective Bypass of a Lagging Strand Roadblock by the Eukaryotic Replicative DNA Helicase. Cell 2011, 146, 931–941.
  72. Wasserman, M.R.; Schauer, G.D.; O’Donnell, M.E.; Liu, S. Replication Fork Activation Is Enabled by a Single-Stranded DNA Gate in CMG Helicase. Cell 2019, 178, 600–611.e16.
  73. Raghuraman, M.K.; Winzeler, E.A.; Collingwood, D.; Hunt, S.; Wodicka, L.; Conway, A.; Lockhart, D.J.; Davis, R.W.; Brewer, B.J.; Fangman, W.L. Replication Dynamics of the Yeast Genome. Science 2001, 294, 115–121.
  74. Anglana, M.; Apiou, F.; Bensimon, A.; Debatisse, M. Dynamics of DNA Replication in Mammalian Somatic Cells: Nucleotide Pool Modulates Origin Choice and Interorigin Spacing. Cell 2003, 114, 385–394.
  75. E Fairman-Williams, M.; Guenther, U.-P.; Jankowsky, E. SF1 and SF2 helicases: Family matters. Curr. Opin. Struct. Biol. 2010, 20, 313–324.
  76. Byrd, A.K.; Raney, K.D. Superfamily 2 Helicases. Front. Biosci. (Landmark Ed.) 2012, 17, 2070–2088.
  77. White, M.F.; Dillingham, M.S. Iron–sulphur clusters in nucleic acid processing enzymes. Curr. Opin. Struct. Biol. 2012, 22, 94–100.
  78. van Brabant, A.J.; Stan, R.; Ellis, N.A. DNA Helicases, Genomic Instability, and Human Genetic Disease. Annu. Rev. Genom. Hum. Genet. 2000, 1, 409–459.
  79. Egly, J.-M.; Coin, F. A history of TFIIH: Two decades of molecular biology on a pivotal transcription/repair factor. DNA Repair 2011, 10, 714–721.
  80. Fuss, J.O.; Tainer, J.A. XPB and XPD helicases in TFIIH orchestrate DNA duplex opening and damage verification to coordinate repair with transcription and cell cycle via CAK kinase. DNA Repair 2011, 10, 697–713.
  81. Kuper, J.; Braun, C.; Elias, A.; Michels, G.; Sauer, F.; Schmitt, D.R.; Poterszman, A.; Egly, J.-M.; Kisker, C. In TFIIH, XPD Helicase Is Exclusively Devoted to DNA Repair. PLoS Biol. 2014, 12, e1001954.
  82. Van Houten, B.; Kuper, J.; Kisker, C. Role of XPD in cellular functions: To TFIIH and beyond. DNA Repair 2016, 44, 136–142.
  83. Ito, S.; Tan, L.J.; Andoh, D.; Narita, T.; Seki, M.; Hirano, Y.; Narita, K.; Kuraoka, I.; Hiraoka, Y.; Tanaka, K. MMXD, a TFIIH-Independent XPD-MMS19 Protein Complex Involved in Chromosome Segregation. Mol. Cell 2010, 39, 632–640.
  84. Yoder, K.; Sarasin, A.; Kraemer, K.; McIlhatton, M.; Bushman, F.; Fishel, R. The DNA repair genes XPB and XPD defend cells from retroviral infection. Proc. Natl. Acad. Sci. USA 2006, 103, 4622–4627.
  85. Handa, N.; Morimatsu, K.; Lovett, S.T.; Kowalczykowski, S.C. Reconstitution of initial steps of dsDNA break repair by the RecF pathway of E. coli. Genes Dev. 2009, 23, 1234–1245.
  86. Hanada, K.; Ukita, T.; Kohno, Y.; Saito, K.; Kato, J.-I.; Ikeda, H. RecQ DNA helicase is a suppressor of illegitimate recombination in Escherichia coli. Proc. Natl. Acad. Sci. USA 1997, 94, 3860–3865.
  87. Harami, G.M.; Seol, Y.; In, J.; Ferencziová, V.; Martina, M.; Gyimesi, M.; Sarlós, K.; Kovács, Z.J.; Nagy, N.T.; Sun, Y.; et al. Shuttling along DNA and directed processing of D-loops by RecQ helicase support quality control of homologous recombination. Proc. Natl. Acad. Sci. USA 2017, 114, E466–E475.
  88. Mohaghegh, P.; Karow, J.K.; Brosh, R.M.; Bohr, V.A.; Hickson, I.D. The Bloom’s and Werner’s syndrome proteins are DNA structure-specific helicases. Nucleic Acids Res. 2001, 29, 2843–2849.
  89. Hishida, T.; Han, Y.-W.; Shibata, T.; Kubota, Y.; Ishino, Y.; Iwasaki, H.; Shinagawa, H. Role of the Escherichia coli RecQ DNA helicase in SOS signaling and genome stabilization at stalled replication forks. Genes Dev. 2004, 18, 1886–1897.
  90. Courcelle, J.; Hanawalt, P.C. RecQ and RecJ process blocked replication forks prior to the resumption of replication in UV-irradiated Escherichia coli. Mol. Genet. Genom. 1999, 262, 543–551.
  91. Bachrati, C.Z.; Hickson, I.D. RecQ helicases: Guardian angels of the DNA replication fork. Chromosoma 2008, 117, 219–233.
  92. Bagchi, D.; Manosas, M.; Zhang, W.; Manthei, K.A.; Hodeib, S.; Ducos, B.; Keck, J.L.; Croquette, V. Single molecule kinetics uncover roles for E. coli RecQ DNA helicase domains and interaction with SSB. Nucleic Acids Res. 2018, 46, 8500–8515.
  93. Nakayama, H.; Nakayama, K.; Nakayama, R.; Irino, N.; Nakayama, Y.; Hanawalt, P.C. Isolation and genetic characterization of a thymineless death-resistant mutant of Escherichia coli K12: Identification of a new mutation (recQ1) that blocks the RecF recombination pathway. Mol. Genet. Genom. 1984, 195, 474–480.
  94. Harami, G.M.; Nagy, N.T.; Martina, M.; Neuman, K.; Kovács, M. The HRDC domain of E. coli RecQ helicase controls single-stranded DNA translocation and double-stranded DNA unwinding rates without affecting mechanoenzymatic coupling. Sci. Rep. 2015, 5, 11091.
  95. Greenleaf, W.J.; Woodside, M.T.; Block, S.M. High-Resolution, Single-Molecule Measurements of Biomolecular Motion. Annu. Rev. Biophys. Biomol. Struct. 2007, 36, 171–190.
  96. Tanner, N.A.; van Oijen, A.M. Visualizing DNA Replication at the Single-Molecule Level. Methods Enzymol. 2010, 475, 259–278.
  97. Dessinges, M.-N.; Lionnet, T.; Xi, X.G.; Bensimon, D.; Croquette, V. Single-molecule assay reveals strand switching and enhanced processivity of UvrD. Proc. Natl. Acad. Sci. USA 2004, 101, 6439–6444.
  98. Lionnet, T.; Spiering, M.M.; Benkovic, S.J.; Bensimon, D.; Croquette, V. Real-time observation of bacteriophage T4 gp41 helicase reveals an unwinding mechanism. Proc. Natl. Acad. Sci. USA 2007, 104, 19790–19795.
  99. Manosas, M.; Xi, X.G.; Bensimon, D.; Croquette, V. Active and passive mechanisms of helicases. Nucleic Acids Res. 2010, 38, 5518–5526.
  100. Manosas, M.; Spiering, M.M.; Zhuang, Z.; Benkovic, S.J.; Croquette, V. Coupling DNA unwinding activity with primer synthesis in the bacteriophage T4 primosome. Nat. Chem. Biol. 2009, 5, 904–912.
  101. Manosas, M.; Perumal, S.K.; Croquette, V.; Benkovic, S.J. Direct Observation of Stalled Fork Restart via Fork Regression in the T4 Replication System. Science 2012, 338, 1217–1220.
  102. Bhattacharyya, B.; George, N.P.; Thurmes, T.M.; Zhou, R.; Jani, N.; Wessel, S.R.; Sandler, S.J.; Ha, T.; Keck, J.L. Structural mechanisms of PriA-mediated DNA replication restart. Proc. Natl. Acad. Sci. USA 2013, 111, 1373–1378.
  103. Cox, M.M.; Goodman, M.F.; Kreuzer, K.N.; Sherratt, D.J.; Sandler, S.J.; Marians, K.J. The importance of repairing stalled replication forks. Nature 2000, 404, 37–41.
  104. Heller, R.C.; Marians, K.J. Replisome assembly and the direct restart of stalled replication forks. Nat. Rev. Mol. Cell Biol. 2006, 7, 932–943.
  105. Merrikh, H.; Zhang, Y.; Grossman, A.D.; Wang, J.D. Replication–transcription conflicts in bacteria. Nat. Rev. Microbiol. 2012, 10, 449–458.
  106. Yeeles, J.T.; Poli, J.; Marians, K.J.; Pasero, P. Rescuing Stalled or Damaged Replication Forks. Cold Spring Harb. Perspect. Biol. 2013, 5, a012815.
  107. McGlynn, P.; A Al-Deib, A.; Liu, J.; Marians, K.J.; Lloyd, R.G. The DNA replication protein PriA and the recombination protein RecG bind D-loops. J. Mol. Biol. 1997, 270, 212–221.
  108. Nurse, P.; Liu, J.; Marians, K.J. Two Modes of PriA Binding to DNA. J. Biol. Chem. 1999, 274, 25026–25032.
  109. Manhart, C.M.; McHenry, C.S. The PriA Replication Restart Protein Blocks Replicase Access Prior to Helicase Assembly and Directs Template Specificity through Its ATPase Activity. J. Biol. Chem. 2013, 288, 3989–3999.
  110. Hernandez, A.J.; Richardson, C.C. Gp2.5, the multifunctional bacteriophage T7 single-stranded DNA binding protein. Semin. Cell Dev. Biol. 2018, 86, 92–101.
  111. Ciesielski, G.L.; Bermek, O.; Rosado-Ruiz, F.A.; Hovde, S.L.; Neitzke, O.J.; Griffith, J.D.; Kaguni, L.S. Mitochondrial Single-stranded DNA-binding Proteins Stimulate the Activity of DNA Polymerase γ by Organization of the Template DNA. J. Biol. Chem. 2015, 290, 28697–28707.
  112. Williams, A.J.; Kaguni, L.S. Stimulation of Drosophila Mitochondrial DNA Polymerase by Single-stranded DNA-binding Protein. J. Biol. Chem. 1995, 270, 860–865.
  113. A Korhonen, J.; Pham, X.H.; Pellegrini, M.; Falkenberg, M. Reconstitution of a minimal mtDNA replisome in vitro. EMBO J. 2004, 23, 2423–2429.
  114. Ghosh, S.; Hamdan, S.; Richardson, C.C. Two Modes of Interaction of the Single-stranded DNA-binding Protein of Bacteriophage T7 with the DNA Polymerase-Thioredoxin Complex. J. Biol. Chem. 2010, 285, 18103–18112.
  115. Kim, Y.; Tabor, S.; Churchich, J.; Richardson, C. Interactions of gene 2.5 protein and DNA polymerase of bacteriophage T7. J. Biol. Chem. 1992, 267, 15032–15040.
  116. Shereda, R.D.; Kozlov, A.G.; Lohman, T.M.; Cox, M.M.; Keck, J.L. SSB as an Organizer/Mobilizer of Genome Maintenance Complexes. Crit. Rev. Biochem. Mol. Biol. 2008, 43, 289–318.
  117. Manosas, M.; Spiering, M.M.; Ding, F.; Bensimon, D.; Allemand, J.-F.; Benkovic, S.J.; Croquette, V. Mechanism of strand displacement synthesis by DNA replicative polymerases. Nucleic Acids Res. 2012, 40, 6174–6186.
  118. Ismael, P.-G.A.; Lemishko, K.M.; Crespo, R.; Truong, T.Q.; Kaguni, L.S.; Cao-García, F.J.; Ciesielski, G.L.; Ibarra, B. Mechanism of Strand Displacement DNA Synthesis by the Coordinated Activities of Human Mitochondrial DNA Polymerase and SSB. bioRxiv 2022.
  119. Sullivan, E.D.; Longley, M.J.; Copeland, W.C. Polymerase γ efficiently replicates through many natural template barriers but stalls at the HSP1 quadruplex. J. Biol. Chem. 2020, 295, 17802–17815.
  120. Fusté, J.M.; Shi, Y.; Wanrooij, S.; Zhu, X.; Jemt, E.; Persson, Ö.; Sabouri, N.; Gustafsson, C.M.; Falkenberg, M. In Vivo Occupancy of Mitochondrial Single-Stranded DNA Binding Protein Supports the Strand Displacement Mode of DNA Replication. PLoS Genet. 2014, 10, e1004832.
  121. Takamatsu, C.; Umeda, S.; Ohsato, T.; Ohno, T.; Abe, Y.; Fukuoh, A.; Shinagawa, H.; Hamasaki, N.; Kang, D. Regulation of mitochondrial D-loops by transcription factor A and single-stranded DNA-binding protein. EMBO Rep. 2002, 3, 451–456.
  122. Nicholls, T.J.; Zsurka, G.; Peeva, V.; Schöler, S.; Szczesny, R.J.; Cysewski, D.; Reyes, A.; Kornblum, C.; Sciacco, M.; Moggio, M.; et al. Linear mtDNA fragments and unusual mtDNA rearrangements associated with pathological deficiency of MGME1 exonuclease. Hum. Mol. Genet. 2014, 23, 6147–6162.
  123. Uhler, J.P.; Thörn, C.; Nicholls, T.J.; Matic, S.; Milenkovic, D.; Gustafsson, C.M.; Falkenberg, M. MGME1 processes flaps into ligatable nicks in concert with DNA polymerase γ during mtDNA replication. Nucleic Acids Res. 2016, 44, 5861–5871.
  124. Zheng, L.; Zhou, M.; Guo, Z.; Lu, H.; Qian, L.; Dai, H.; Qiu, J.; Yakubovskaya, E.; Bogenhagen, D.F.; Demple, B.; et al. Human DNA2 Is a Mitochondrial Nuclease/Helicase for Efficient Processing of DNA Replication and Repair Intermediates. Mol. Cell 2008, 32, 325–336.
  125. He, Q.; Shumate, C.K.; A White, M.; Molineux, I.J.; Yin, Y.W. Exonuclease of human DNA polymerase gamma disengages its strand displacement function. Mitochondrion 2013, 13, 592–601.
  126. Farge, G.; Pham, X.H.; Holmlund, T.; Khorostov, I.; Falkenberg, M. The accessory subunit B of DNA polymerase is required for mitochondrial replisome function. Nucleic Acids Res. 2007, 35, 902–911.
  127. Farr, C.L.; Wang, Y.; Kaguni, L.S. Functional Interactions of Mitochondrial DNA Polymerase and Single-stranded DNA-binding Protein: Template-primer DNA binding and initiation and elongation of DNA strand synthesis. J. Biol. Chem. 1999, 274, 14779–14785.
  128. Macao, B.; Uhler, J.P.; Siibak, T.; Zhu, X.; Shi, Y.; Sheng, W.; Olsson, M.; Stewart, J.B.; Gustafsson, C.M.; Falkenberg, M. The exonuclease activity of DNA polymerase γ is required for ligation during mitochondrial DNA replication. Nat. Commun. 2015, 6, 7303.
  129. Canceill, D.; Viguera, E.; Ehrlich, S.D. Replication Slippage of Different DNA Polymerases Is Inversely Related to Their Strand Displacement Efficiency. J. Biol. Chem. 1999, 274, 27481–27490.
  130. Stano, N.M.; Jeong, Y.-J.; Donmez, I.; Tummalapalli, P.; Levin, M.K.; Patel, S.S. DNA synthesis provides the driving force to accelerate DNA unwinding by a helicase. Nature 2005, 435, 370–373.
  131. McInerney, P.; O’Donnell, M. Replisome Fate upon Encountering a Leading Strand Block and Clearance from DNA by Recombination Proteins. J. Biol. Chem. 2007, 282, 25903–25916.
  132. Heller, R.C.; Marians, K.J. The Disposition of Nascent Strands at Stalled Replication Forks Dictates the Pathway of Replisome Loading during Restart. Mol. Cell 2005, 17, 733–743.
  133. Cordeiro-Stone, M.; Makhov, A.M.; Zaritskaya, L.S.; Griffith, J.D. Analysis of DNA replication forks encountering a pyrimidine dimer in the template to the leading strand. J. Mol. Biol. 1999, 289, 1207–1218.
  134. Higuchi, K.; Katayama, T.; Iwai, S.; Hidaka, M.; Horiuchi, T.; Maki, H. Fate of DNA replication fork encountering a single DNA lesion during oriC plasmid DNA replication in vitro. Genes Cells 2003, 8, 437–449.
  135. Kowalezykowski, S.C. Biochemistry of Genetic Recombination: Energetics and Mechanism of DNA Strand Exchange. Annu. Rev. Biophys. Biophys. Chem. 1991, 20, 539–575.
  136. Kowalczykowski, S.C.; Clow, J.; Somani, R.; Varghese, A. Effects of the Escherichia coli SSB protein on the binding of Escherichia coli RecA protein to single-stranded DNA: Demonstration of competitive binding and the lack of a specific protein-protein interaction. J. Mol. Biol. 1987, 193, 81–95.
  137. Thresher, R.J.; Christiansen, G.; Griffith, J.D. Assembly of presynaptic filaments: Factors affecting the assembly of RecA protein onto single-stranded DNA. J. Mol. Biol. 1988, 201, 101–113.
  138. Cazenave, C.; Toulmé, J.; Hélène, C. Binding of RecA protein to single-stranded nucleic acids: Spectroscopic studies using fluorescent polynucleotides. EMBO J. 1983, 2, 2247–2251.
  139. Lohman, T.M.; Ferrari, M.E. Escherichia coli single-stranded DNA-binding protein: Multiple DNA-binding modes and cooperativities. Annu. Rev. Biochem. 1994, 63, 527–570.
  140. Naufer, M.N.; Morse, M.; Möller, G.B.; McIsaac, J.; Rouzina, I.; Beuning, P.J.; Williams, M.C. Multiprotein E. coli SSB–ssDNA complex shows both stable binding and rapid dissociation due to interprotein interactions. Nucleic Acids Res. 2021, 49, 1532–1549.
  141. Crozat, E.; Grainge, I. FtsK DNA Translocase: The Fast Motor That Knows Where It’s Going. Chembiochem 2010, 11, 2232–2243.
  142. Fishburn, J.; Tomko, E.; Galburt, E.; Hahn, S. Double-stranded DNA translocase activity of transcription factor TFIIH and the mechanism of RNA polymerase II open complex formation. Proc. Natl. Acad. Sci. USA 2015, 112, 3961–3966.
  143. Sokoloski, J.E.; Kozlov, A.G.; Galletto, R.; Lohman, T.M. Chemo-mechanical pushing of proteins along single-stranded DNA. Proc. Natl. Acad. Sci. USA 2016, 113, 6194–6199.
More
ScholarVision Creations