Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2024 Aug 1.
Published in final edited form as: DNA Repair (Amst). 2023 Jun 24;128:103528. doi: 10.1016/j.dnarep.2023.103528

Single-molecule Imaging of Genome Maintenance Proteins Encountering Specific DNA Sequences and Structures

Elizabeth Marie Irvin 1, Hong Wang 1,2,3,*
PMCID: PMC10989508  NIHMSID: NIHMS1914619  PMID: 37392578

Abstract

DNA repair pathways are tightly regulated processes that recognize specific hallmarks of DNA damage and coordinate lesion repair through discrete mechanisms, all within the context of a three-dimensional chromatin landscape. Dysregulation or malfunction of any one of the protein constituents in these pathways can contribute to aging and a variety of diseases. While the collective action of these many proteins is what drives DNA repair on the organismal scale, it is the interactions between individual proteins and DNA that facilitate each step of these pathways. In much the same way that ensemble biochemical techniques have characterized the various steps of DNA repair pathways, single-molecule imaging (SMI) approaches zoom in further, characterizing the individual protein-DNA interactions that compose each pathway step. SMI techniques offer the high resolving power needed to characterize the molecular structure and functional dynamics of individual biological interactions on the nanoscale. In this review, we highlight how our lab has used SMI techniques – traditional atomic force microscopy (AFM) imaging in air, high-speed AFM (HS-AFM) in liquids, and the DNA tightrope assay – over the past decade to study protein-nucleic acid interactions involved in DNA repair, mitochondrial DNA replication, and telomere maintenance. We discuss how DNA substrates containing specific DNA sequences or structures that emulate DNA repair intermediates or telomeres were generated and validated. For each highlighted project, we discuss novel findings made possible by the spatial and temporal resolution offered by these SMI techniques and unique DNA substrates.

2. Introduction

Cellular DNA damage repair involves a series of complex pathways each serving a function required to repair the lesion, including remodeling the surrounding chromatin, recognizing the damage site, and recruiting the appropriate downstream proteins for repair of the type of damage (16). The DNA repair research field has historically relied on biochemical assays and, more recently, time course or live-cell fluorescence imaging to study the protein-nucleic acid interactions that drive these pathways (712). While these techniques are plentiful and powerful, biochemistry assays generally share the limitation of characterizing ensemble molecular activities instead of individual interactions between biomolecules (13,14); consequently, these assays likely fail to capture important transient interactions. Further, while the advantages of live-cell imaging, including it being in cellulo and offering resolution typically on the hundreds of nanometers and approaching sub-nanometers with super-resolution imaging (15,16), are not to be overlooked, it can be challenging to generate a specific form of DNA damage in cellulo at defined locations and pinpoint direct interactions between a protein and DNA damage.

In vitro single-molecule imaging (SMI) approaches enable the study of the structure and dynamics of individual biomolecular interactions with nanometer resolution, often in real-time (1727). The purpose of this review is to discuss how two SMI techniques, atomic force microscopy (AFM) and the DNA tightrope assay, can be used to study the molecular structure and dynamics of proteins of interest on DNA substrates emulating DNA repair or replication intermediates. The basic principles of these two approaches are first described, followed by a discussion of the generation and validation of select DNA substrates. Lastly, examples of how the high spatial and temporal resolution of SMI in combination with defined DNA substrates expanded our understanding of the cohesin complex, mitochondrial DNA replication, and telomere maintenance are highlighted. Additional information regarding sample preparation and instrumentation can be found in the referenced sources and previously published review papers (2832).

3. AFM and the DNA tightrope assay: two complementary SMI techniques confer nanometer resolution

3.1. AFM reveals structure and dynamics of protein-nucleic acid interactions

AFM is a powerful SMI technique that had been used to study numerous DNA repair pathways, including nucleotide excision repair (3335), mismatch repair (3638), base excision repair (19,3941), and homologous recombination (42,43). AFM imaging in air collects static images of a dry sample (2830,4446). The sample, containing proteins and nucleic acids diluted in a buffer, is typically deposited onto a freshly cleaved muscovite mica surface. The affinity between DNA molecules in the sample and the negatively charged mica surface can be supported by a variety of buffer chemistries or surface modification, including divalent magnesium ions and aminopropyl silatrane (APS) treatment. The sample deposition protocols require striking the careful balance of sufficiently adhering proteins and nucleic acids to the surface without interfering with their interactions and achieving equilibration on the surface as in an ideal two-dimensional solution (47,48). Once deposited, the sample is dried under a stream of nitrogen gas and typically imaged by oscillating an AFM cantilever through a vibration piezo at its first resonant frequency using the tapping (intermittent contact) mode (Fig. 1A) (44). Interactions between the cantilever and biomolecules on the surface cause deflection of the cantilever, damping the amplitude and modulating the phase of its oscillation. This change in oscillation pattern is monitored through a laser reflected off the back of the cantilever and directed at a four-quadrant photodetector (Fig. 1A). The associated sample height information for constructing a topographic image is obtained through a feedback system that maintains a constant set point for the amplitude of cantilever oscillation (49). Proteins on double stranded DNA (dsDNA) can be identified based on their heights and the AFM cantilever oscillation phase differentials: protein-DNA complexes are generally taller than DNA alone and display a lower phase (meaning they appear darker than nucleic acids in the phase channel of the image) (50). AFM can be used to study the oligomeric state of a protein (50,51), its binding affinity for sequences of interest (5153), and its activities on various substrates, including but not limited to DNA looping (5456), bridging (53,56), compaction (56,57), and unwinding (31,58).

Figure 1: Atomic force microscopy (AFM) and DNA tightrope assay.

Figure 1:

(A) Diagram of AFM operation mechanism (29). (B) Schematic of Dual-Resonance-frequency-enhanced electrostatic force enhanced microscopy (DREEM) and example images of a nucleosome on dsDNA (76). (C) Diagram of the DNA tightrope assay. Gray spheres represent poly-L-lysine-treated silica beads (52). (D) Schematic of QD conjugation strategies. Left: Conjugation of His-tagged proteins to streptavidin-coated QDs through the biotinylated BT-tris-NTA linker. Right: Conjugation of epitope-tagged proteins to antibody-coated QDs using the “antibody sandwich” method (51,53).

While AFM imaging in air provides snapshots of protein-DNA complexes in action, imaging in liquids enables direct observation of sequential steps and conformational changes during biological processes (5961). In addition, by reducing capillary forces that otherwise occur between the tip and a dry sample under ambient humidity, liquid imaging improves the spatial resolution of the images (62,63). Recent developments in enhancing the response speed of key devices, including the AFM closed feedback loop, small AFM cantilevers with high resonance frequencies, fast amplitude detectors, and high-speed scanners, have enabled high speed AFM (HS-AFM) imaging with up to ~50 frames per second (6469). It is worth noting that for studying protein-DNA complexes, functionalization of the mica surface with the stable APS compound can enhance the reproducibility of time-lapse imaging in liquids (47).

HS-AFM has been broadly applied to study the dynamics of many protein complexes. Using HS-AFM, key protein conformational changes and novel activities have been uniquely revealed that traditional biochemistry techniques could only predict from static models. HS-AFM imaging of the glutamate transporter reconstituted in a lipid mixture revealed ligands are moved by largescale amplitude changes of key transport domains in the complex. This “elevator” mechanism was predicted by crystal structures as the mechanism by which neurotransmitters are removed from the synaptic cleft, but the dynamism-capturing abilities of HS-AFM enabled its validation (70). Similarly, the mechanism by which nucleoporins promote bidirectional macromolecule transit through nuclear pores was visualized in oocytes. This revealed a rapid and repeated lengthen-shorten transition of nucleoporins, which drives the movement of macromolecules through the pore (71). The conformational changes of KIF2 associated with ATP hydrolysis were elucidated with HS-AFM, furthering understanding of how this microtubule-destabilizing kinesin depolymerizes microtubules to facilitate intracellular transport and cellular movement (72). Particularly fascinating HS-AFM videos confirmed the ‘hand-over-hand’ movement of myosin V walking along actin fibrils, providing the first dynamic visual evidence of myosin V’s translocation mechanism (73). Part of the contended mechanism by which F1-ATPase generates ATP was elucidated with HS-AFM, which showed the stator α3β3 ring cooperatively drives the power stroke of the complex independently of the central γ channel (74). Both the conformational changes and the three-dimensional diffusion of the CRISPR-Cas9-RNA complex along DNA were resolved with HS-AFM (75). These few examples out of a plethora highlight the versatility of HS-AFM and its unique ability to reveal the dynamic conformations and activities of protein complexes, both in vitro and in cellular membranes.

In traditional AFM images, DNA within protein-DNA complexes cannot be resolved, which prevents us from learning the DNA conformations in these complexes. Taking advantage of the electrostatic signals from DNA and proteins, we (Erie and Wang groups) developed Dual Resonance Frequency Enhanced Electrostatic Force Microscopy (DREEM) that simultaneously collects both topographic and electrostatic force gradient images (Fig. 1B). Through applying both a constant and modulated bias voltage to the cantilever, DREEM can resolve DNA within protein-DNA complexes due to charge neutralization where the protein and DNA interact (50,76). DREEM revealed dsDNA wrapping around nucleosomes and the path of dsDNA through mismatch repair proteins, establishing it as a powerful imaging technique to study DNA conformations in heterogeneous protein-DNA complexes (76).

3.2. DNA tightrope assay reveals protein binding dynamics

The DNA tightrope assay is a low-cost yet versatile fluorescence-based SMI method that enables a variety of dynamic protein-nucleic acid interactions to be studied (77). To form “DNA tightropes” in a flow chamber, DNA is stretched between poly-L-lysine-treated silica beads by hydrodynamic buffer flow using a syringe pump (Fig. 1C). Proteins labeled with either streptavidin or antibody-coated quantum dots (QDs) (Fig. 1D) are then flowed over the DNA tightropes. Fluorescence imaging of QD-labeled proteins is carried out using oblique angle Total Internal Reflection Fluorescence (TIRF) or highly inclined and laminated optical sheets (HILO) microscopy, which directs the laser slightly below the critical angle and enhances the signal-to-noise ratio compared to epifluorescence imaging (78). Tracking of QD-labeled proteins on DNA with high spatial (16 nm) and temporal resolutions (~50 ms/frame) allows us to study multiple parameters of protein activity. These include a protein’s diffusion constant (the rate at which a protein moves across DNA in µm2/s), alpha factor or diffusion exponent (characterizes a protein’s tendency to move in an unbiased fashion across DNA or to pause, suggesting affinity for a sequence), attachment lifetime (time bound to DNA from initial binding to dissociation), diffusion range (the distance over which a protein moves on the DNA tightrope), and binding position (79). Because DNA tightropes are elevated above the glass surface by the beads, DNA can be visualized in an extended form and protein movement on the DNA is not influenced by interactions with the surface and can be monitored without buffer flow (77).

For the DNA tightrope assay, we use either lambda DNA of ~ 48 kbp or ligated linear DNA substrates of ~20 to 40 kbp. The ligated DNA is prepared by restriction enzyme digestion of a plasmid containing a sequence of interest, followed by tandem end-to-end ligation. This protocol ensures that the DNA substrate is of sufficient length for stringing between the beads and the sequence of interest is positioned at regularly spaced intervals (Fig. 1C); Under the standard flow-stretching conditions (300 µl/min), the observed spacing of QD-labeled proteins on the tightropes is ~90% of what is expected for fully stretched DNA (79). Typically, when collecting data, DNA tightropes are not stained with dyes, such as YOYO1 or DAPI, as intercalating dyes have the potential to interfere with DNA binding by proteins. Rather, to confirm the presence of DNA tightropes, YOYO1 can be added to stain the DNA after collecting the protein-DNA binding data. Furthermore, the tension on DNA and the extent of DNA stretching can be adjusted using different buffer flow rates (79). Previously, we found that diffusion constants of TRF2 protein do not change significantly on DNA tightropes formed using 300 µl/min or 25 µl/min (stretching to ~88% DNA contour) (79). The optimal buffer flow rate that minimizes the impact of DNA tension on protein binding must be empirically determined for each individual protein. In addition, it is worth noting that DNA tightropes in stretched DNA form do not allow various DNA binding pathways, including segmental site transfer (80), DNA looping and compaction. To study DNA looping and compaction, we can anchor YOYO1-stained DNA at one end through biotin-streptavidin interactions and flow stretch the DNA, as demonstrated previously by numerous other groups (21,81,82).

4. SMI substrates: generation and validation of substrates emulating DNA repair and replication intermediates

4.1. Linear DNA with single-stranded “gap” region, single-stranded flap, and replication fork

Single-stranded DNA (ssDNA) is present ubiquitously in the nucleus, both as a structural component at the ends of chromosomes and as a transient intermediate during DNA replication (83), recombination, and repair (8486). ssDNA plays numerous supporting roles in these processes, including helping to stimulate the reversal of stalled replication forks (83) and recruiting DNA repair proteins (86,87). While its formation is essential for these processes to occur, ssDNA is vulnerable to nuclease degradation, and thus it is critical for single-stranded binding (SSB) proteins to sequester ssDNA and protect it while also stimulating proteins that process it, including nucleases and helicases (84,88), in a highly coordinated fashion. Replication Protein A (RPA) is perhaps the most studied SSB in eukaryotes, as it rapidly coats freshly exposed ssDNA to prevent the formation of secondary structures that would prohibit subsequent recombination and repair (83,87).

With great diversity in ssDNA-binding domains (88) and a plethora of pathways in which ssDNA is an intermediate structure, novel single-stranded DNA binding proteins are discovered regularly. Given the multifaceted nature of SSB protein interaction networks, SMI approaches offer an expansive and detailed means of studying SSB-ssDNA interactions, enabling their inherently transient and highly dynamic characteristics to be captured (82,8992).

To study ssDNA binding proteins, we generated a linear double-stranded DNA substrate containing a short single-stranded region (37 nt) referred to as the gap site. Treatment of a starting plasmid (originally generated in the Peggy Hsieh lab (93)) with a nicking enzyme generates a region with four closely-spaced nicks (Fig. 2A top panel). The gap site is exposed by adding an excess of DNA oligonucleotides complementary to the ssDNA between nicked sites (that dissociate from the duplex DNA during heating) (94). We quantify the gapping efficiency (typically ~90%) by restriction digestion at the gapped region. Subsequent restriction digestion of the gapped plasmid positions the single-stranded region at a defined location (23% from DNA end, Fig. 2A). Specific binding of mitochondrial single-stranded DNA binding protein (mtSSB) at the gap region further validates the substrate (Fig. 2A bottom panel). To exclude that the DNA sequence at the gap site does not contribute to protein binding preference, a control DNA substrate is generated by directly linearizing the plasmid (without gapping) (50,95).

Figure 2: Generation and validation of DNA substrates that emulate DNA repair and replication intermediates.

Figure 2:

(A) Generation of linear gap DNA and its validation by mtSSB, which preferentially binds to the ssDNA (gap) region (50). (B) Generation of the DNA replication fork substrate and binding of cohesin SA2 at the fork junction (51). (C) Generation of the linear R-loop DNA through in vitro transcription and its validation using RNA:DNA hybrid-binding S9.6 antibody (52).

The protocol for generating the linear gap DNA can be modified to produce other substrates, including a “flap” or “fork” structure. By incubating the linear gap DNA with an oligo complementary at one end to the gap site, the non-complementary end of the oligo protrudes from the duplex once annealed to the gap site. This produces linear DNA with a single-stranded flap analogous to DNA repair intermediate structures (86). This single-stranded flap substrate can also be easily made into a linear double-stranded fork (analogous to a replication fork stalled when DNA damage is sustained during S-phase (96)), simply by incubating with an additional oligo complementary to the protruding flap (51).

4.2. DNA replication fork substrate

When DNA damage is sustained, replication pauses, leading to an accumulation of stalled replication forks (9699). Various DNA repair proteins have been shown to bind stalled replication forks to protect them from nuclease degradation and promote their regression and restart (100102). These factors are thought to be just the tip of the iceberg in a complex network of proteins involved in fork recognition and restart (100). To expand what is known about this protein network, in addition to the linear fork substrate described above, we also generated a circular DNA replication fork using a protocol established by the Jack Griffith group (Fig. 2B) (103). Briefly, the pGLGAP plasmid containing a 398-bp G-less cassette was nicked using a Nickase, followed by nick translation in the presence of dATP, dTTP, and dGTP using the Klenow fragment (exo-) to generate an ssDNA tail. To create a dsDNA tail, a primer was annealed to the ssDNA tail, followed by the extension using the Klenow fragment and purification. While we generated a 25 nt single-stranded gap region at the fork junction in our previous study (51), the length of the ssDNA gap region at the fork junction can be tuned by positioning the primer at different locations on the ssDNA tail. Typically, ~80% of the circular DNA displays a dsDNA tail, and the completion of the DNA replication fork substrate is validated using AFM imaging by measuring the length of the dsDNA tail (~129 nm) and DNA binding of cohesion SA2 protein, which binds at fork junctions (Fig. 2B bottom) (51).

4.3. R-loop DNA substrate

R-loops are RNA-DNA hybrid structures that form when a nascent RNA anneals to the template DNA strand, displacing the non-template strand as a ssDNA loop (104,105). R-loops appear to play diverse and important roles in normal cellular function, including regulating gene expression (105), telomere maintenance (104,106), and DNA double-strand break repair (104106). However, aberrant R-loop formation is also associated with genome instability: their formation can be induced by DNA damage (106,107), and if not promptly resolved, their accumulation can be genotoxic (104107). The multifaceted roles of R-loops in both DNA damage and repair make studying them complex, and understanding of how they are resolved remains limited, with an array of RNA:DNA helicases, RNases and DNA repair proteins that appear to contribute to their resolution through regulatory pathways that are not fully understood (106,107).

To study how DNA repair proteins function on R-loops, we generated a linear R-loop substrate by performing in vitro transcription on a plasmid originally created by the Chédin laboratory (108). This substrate contains the Airn gene cloned downstream of the T3 promotor (Fig. 2C top panel). A previous study established that Airn contains two R-loop forming sites (38% and 42% from one DNA end on our linearized DNA), with one overlapping with a G-stretch forming more stable R-loops. RNase treatment removes unpaired single-stranded RNA. Linearization by a restriction enzyme positions the R-loop regions at defined positions. Interestingly, AFM imaging revealed that the extruded non-template strand in the R-loop adopts distinct three-dimensional architectures, including ‘loop,’ ‘spur,’ and ‘blob’ conformations (108). These structures vary significantly in their sizes and shapes, and are distinguishable from dsDNA by their relative heights (Fig. 2C bottom panel) and volumes (108) using AFM. As the ‘spur’ and ‘loop’ conformations protrude outwards from the DNA backbone, there is uncertainty in determining their positions along the linear DNA. Thus, we focused our analysis on the ‘blob’ shape, which falls along the DNA backbone and is conveniently also the most abundant R-loop conformation (52). The presence and position of the R-loop can be further validated with an S9.6 antibody, which binds RNA-DNA hybrids and can be identified in AFM images based on its height above dsDNA (Fig. 2C bottom panel) (52,109). AFM imaging allows us to discern whether a protein preferentially binds to and is activated by R-loops based on its binding positions along the R-loop DNA (52,109). By measuring the maximum height of the R-loop from many molecules, we first established a threshold height for distinguishing the R-loop from surrounding dsDNA (0.89 nm, Fig. 2C bottom panel). We then can determine if a protein or complex of interest binds the R-loop position by measuring the height of that protein-R-loop complex. If bound to the R-loop position, the protein-R-loop complex should display a height significantly greater than R-loop alone.

5. Highlights: Novel findings from the AFM and the DNA tightrope assay

5.1. Mitochondrial DNA replication machinery

The mitochondrial genome is organized into many copies of circular, double-stranded DNA that encodes essential cellular respiration enzymes (110), as well as transfer and ribosomal RNAs required for translation within the mitochondria (110,111). Numerous heritable diseases are associated with the failure of the mitochondrial genome to be replicated and repaired properly (112). The mitochondrial replisome consists of the DNA polymerase γ holoenzyme, a single-stranded DNA binding protein (mtSSB), and a helicase known as Twinkle (50,95,113). mtSSB assists the mitochondrial replisome by suppressing secondary structures that form in unwound ssDNA (114) and stimulating the activities of DNA Polymerase γ (114,115) and Twinkle (116,117). When mtSSB is knocked down in cultured human cells, mtDNA synthesis decreases, leading to a reduction in mitochondrial genome copy number over time (118). Twinkle might also function in recombination-mediated double-stranded break (DSB) repair of mitochondrial DNA (119). The functions and activities of the E. coli replisome are well-characterized with significant contributions having been made by AFM and HS-AFM studies. Thanks to the ability to visualize the dynamics of protein-nucleic acid interactions, AFM has uniquely revealed that the RecG and PriA helicases of the E. coli replisome translocate away from the replication fork junction, in processes facilitated by E. coli SSB (120,121) and ATP (122,123), respectively. Additionally, AFM imaging has captured the collaborative binding and oligomerization of E. coli SSB tetramers on ssDNA (124,125). However, our understanding of how the mitochondrial genome is replicated and repaired, and how that process can be disrupted in human diseases, is limited. AFM and DREEM imaging offer the resolution needed to help address these gaps.

Previously, typical DNA substrates utilized to assess DNA unwinding by Twinkle in vitro contain only short duplex regions (~60 bp) (126). Thus, the dynamics and processivity of DNA unwinding by Twinkle were largely unknown. To fill this knowledge gap, we used AFM imaging in air to characterize the helicase activity of human Twinkle (hTwinkle) that generates ssDNA (31,95). We imaged hTwinkle on circular gap DNA in the presence of ATP and observed an accumulation of ssDNA with lower heights than dsDNA by approximately 200 pm (Fig. 3A). AFM imaging of hTwinkle-gap DNA samples deposited on mica surfaces after defined incubation times revealed that hTwinkle unwinds several hundred base pairs of duplex DNA at an average rate of ~240 bp/min.

Figure 3: Structure and dynamics of DNA binding and unwinding by Twinkle revealed by AFM imaging in air and HS-AFM imaging in liquids.

Figure 3:

(A) Representative AFM image capturing hTwinkle’s helicase activity. ssDNA accumulates around hTwinkle as it unwinds duplex DNA (left panel). ssDNA (denoted by gray line) can be distinguished from dsDNA (denoted by blue line) by their height difference (~200 pm, right panel) (95). (B) Representative AFM images of hTwinkle in the presence of mtSSB on circular DNA with two gap sites (left panels), and quantification of hTwinkle’s DNA unwinding rate (bp/min) with increasing concentrations of mtSSB (right panels). hTwinkle partially unwinds one or both gap sites, or fully unwinds the duplex between the gap sites (95). (C) Time-lapse HS-AFM images showing LcTwinkle captures DNA through N-protrusion (denoted by green arrow) (129). Also see Video S1. (D) Time-lapse HS-AFM images showing LcTwinkle transiently binds to DNA through N-protrusion (denoted by green arrow), followed by the transfer to its central channel (denoted by pink arrow) (129). Also see Video S2. (E) Model of Twinkle DNA binding based on cryo-EM structure and AFM imaging of LcTwinkle-DNA (129).

While mtSSB is clearly established as an essential player in mitochondrial genome replication, full understanding of how it binds to DNA to complete these roles has been limited (50), compared to E. coli SSBs (117,120122,124,127). We first used AFM imaging in air to establish the native conformation of mtSSB in the absence of DNA. Volumetric analysis of individual mtSSB complexes correlated with mtSSB forming tetramers, as assessed using the standard calibration curve detailed above (50). Knowing the oligomeric state of apo-mtSSB, we next applied AFM imaging in air as well as DREEM imaging to characterize the association of mtSSB with DNA. When incubated with single-stranded circular DNA, mtSSB binding density increases proportionally to its concentration. However, each bound mtSSB is clearly separate from adjacent ones on the DNA, indicating mtSSB does not bind ssDNA cooperatively, making it distinctly different from E. coli SSB (50). When incubated on the linear gap DNA, mtSSB preferentially binds the single-stranded gap site (Fig. 2A), while it is randomly distributed across control DNA. Volumetric analysis again indicated mtSSB binds the gap site as a tetramer (see (50)). Interestingly, mtSSB binding appeared to shorten the linear gap DNA by approximately 20 nm or 40 nucleotides, equating to the girth of one mtSSB tetramer. This raised the question if mtSSB wraps ssDNA around itself. To address this question, we probed the molecular structure of mtSSB bound at the gap position using DREEM imaging. Previously, we established that DREEM can reveal ssDNA wrapping around histone proteins (Fig. 4A) (128). Using DREEM, we observed DNA (visible as a thin light-colored band wrapped around dark structures) wrapping around individual mtSSB tetramers (the darker structures) with one turn at the position of the gap site (Fig. 4B), consistent with the measured shortening of the gap DNA.

Figure 4: DREEM reveals DNA paths in protein-DNA complexes.

Figure 4:

(A) ssDNA wraps twice around histone proteins (128). (B) dsDNA wraps around mtSSB once (50). (C) DNA appears at the edge of large TRF2 complexes (208). (D) dsDNA wraps around the TRFH domain from TRF2 (171). Left panels: Topographical AFM images. Right panels: DREEM images. Inserts in A, B, and D: models of protein-DNA complexes based on DREEM images.

Beyond characterizing the structure of mtSSB bound to DNA, we also used AFM imaging in air to study how mtSSB affects the helicase activity of Twinkle. We co-incubated hTwinkle with increasing concentrations of mtSSB. By tracking the sample incubation time and measuring the length of ssDNA produced by hTwinkle at increasing concentrations of mtSSB, we discovered that the addition of mtSSB results in a ~5-fold stimulation of the apparent DNA-unwinding rate (Fig. 3B). mtSSB also increases the estimated translocation processivity from 1750 to >9000 bp before hTwinkle disassociation (95). AFM analyses indicated that mtSSB both enhances loading of Twinkle onto defined DNA substrates and stabilizes the unwound ssDNA product (95).

How Twinkle self-loads onto DNA without a loader protein was a long-standing question. While AFM imaging in air provides the resolution needed to observe Twinkle’s oligomeric state and helicase activity, it lacks the temporal resolution for capturing Twinkle’s conformational changes required for self-loading onto DNA. By tuning the 1-(3-aminopropyl)silatrane (APS) (47) concentration on a mica surface, we developed robust sample deposition conditions to achieve the intricate balance of keeping protein and DNA molecules partially anchored onto a surface while still being mobile (129). HS-AFM imaging in liquids (0.8 to 2.3 frames/s) enabled Twinkle’s DNA binding dynamics to be observed in real-time. For HS-AFM imaging, we used a Twinkle homolog that is similar to hTwinkle, Lates calcarifer (Lc) Twinkle, which is amendable to cryo-EM structure determination (129). At a distance (> 200 nm) from DNA, LcTwinkle switches between open and closed ring conformations, but exhibits limited conformational changes. Interestingly, when DNA is in close proximity (distance <20 nm), conformations of LcTwinkle change dramatically with small protrusions (~5 nm) directed towards the DNA (Fig. 3C, green arrow; Video S1) observed from all LcTwinkle molecules. LcTwinkle appears to use this protrusion to initially capture DNA, followed by the loading of DNA into its central channel in the open-ring form (Fig. 3D, pink arrow; Video S2) (129). Taken together, the results from HS-AFM and cryo-EM (129) revealed two separate DNA binding regions on LcTwinkle: the N-terminal domain (NTD) for initial transient DNA capture and the helicase central channel for DNA unwinding. Our results suggest that electrostatic interactions between the positive charges on the NTD and the negatively charged DNA drive the initial loading of Twinkle on DNA (Fig. 3E). We proposed that the flexible N-C linker between the NTD and CTD enables the NTD to protrude in the presence of DNA.

5.2. R-loops recruit cohesin SA1/SA2 and PARP1

Emerging evidence supports the notion that the three-dimensional (3D) structure of the interphase genome plays key roles in regulating DNA replication and repair (130). In addition to its role in maintaining sister chromatids cohesion during S-phase, the cohesin complex, assisted by NIPBL, is the primary orchestrator of the 3D genome organization (131134). Cohesin contains SMC1, SMC3, RAD21, and mutually exclusive SA1 or SA2 (135). Differences in cohesin-SA1 and cohesin-SA2 relative abundances in human cells suggest these complexes might have different functions conferred by SA1 and SA2 (136).

While cohesin helps maintain the architecture of the entire genome, it is also recruited to DNA damage sites (137142). Recent research shows cohesin performs chromatin loop extrusion at these damage sites, and this activity appears to promote the formation of DNA damage response foci (143) and subsequent DNA repair (140,144146). However, the mechanisms by which cohesin recognizes various DNA damage sites and promotes their repair is not well understood. Additionally, research has shown that SA1 and SA2 are differentially essential for maintaining sister chromatids cohesion between telomeric and centromeric regions, respectively (136). How SA1 and SA2 might recognize these sequences and DNA damage sites and differentially respond to them has not been well studied. Using AFM and the DNA tightrope assay, we showed that SA1 and SA2 are DNA-binding proteins that independently recognize several hallmarks of DNA damage and display differential binding dynamics on DNA sequences of interest (51,53).

AFM imaging revealed both SA1 and SA2 directly bind dsDNA and target DNA ends in particular (51,53). However, SA1 and SA2 display distinct binding behaviors on DNA substrates containing telomeric or centromeric sequences. AFM imaging of SA1 on telomeric T270 DNA showed a higher percentage of SA1 binding events occurring at the telomeric sequence repeat region compared to that same region of control DNA (53). SA2, meanwhile, does not specifically bind to telomeric or centromeric sequences (51). This finding highlights a distinct benefit offered by AFM compared to ensemble biochemistry assays: while techniques like electrophoretic mobility shift assay (EMSA) and fluorescence anisotropy are powerful tools for gauging if a protein can bind a nucleic acid substrate, they offer no information on where binding occurs, such as at DNA ends versus internally. Furthermore, the DNA tightrope assay offered insight into how SA1 finds the telomeric region and revealed unique binding modes that SA1 and SA2 display. The majority of QD-labeled SA1 statically binds the T270 tightropes at regularly spaced intervals corresponding to the position of telomeric repeats for extensive time periods often exceeding the 2-minute recording window, suggesting a high-affinity interaction. Some SA1 proteins diffuse across the tightropes at ranges spanning several telomeric repeats and sometimes paused at telomeric regions. SA1 displays a much shorter dwell time and diffuses freely across tightropes containing genomic DNA and centromeric sequences, further supporting SA1’s preference for telomeric sequences (53). QD-labeled SA2 diffuses freely over DNA tightropes containing either genomic, centromeric or telomeric sequences, supporting our findings with AFM that SA2 shows no binding preference for these sequences in vitro (51).

While SA2 displays no clear sequence specificity for the aforementioned dsDNA substrates, it does preferentially bind to ssDNA (51). Using AFM imaging in air, we found that SA2 primarily binds at the single-stranded region of linear gap DNA, while it is randomly distributed across control dsDNA (51). AFM imaging revealed that DNA binding (dsDNA as well as ssDNA) stimulates SA2 to transition from a mostly monomeric form to a higher-order oligomeric state (51,52). QD-labeled SA2 displays one-dimensional search behavior, diffusing over DNA and binding statically to regularly spaced ssDNA gap regions. On control non-gapped DNA, SA2 is primarily diffusive. We also found using AFM imaging that SA2 preferentially binds to DNA flap and fork substrates, with most binding events occurring at the DNA junction on the DNA replication fork substrate (Fig. 2B bottom panel) (51). Taken together, these results show a clear preference by SA2 for intermediate structures during DNA repair, recombination, and replication. These findings led us to propose a novel model in which cohesin SA1 and SA2 serve as “anchor” proteins to localize cohesin at specific DNA sequences and structures. Recently, the cryo-EM structure of cohesin-NIPBL (147) and biochemical studies (64) further establish the interface between SA1 and DNA.

Given the increasingly recognized role RNA plays in DNA repair (148152), we explored the behavior of SA1 and SA2 on various RNA and RNA:DNA hybrid substrates. AFM height analysis showed SA1 and SA2 oligomerize on ssRNA oligos, long ssRNA, and R-loop DNA (52). On the linear R-loop DNA, both SA1 and SA2 preferentially target the R-loop region (Fig. 5A) yet randomly distribute across control DNA. This was validated with the tightrope assay, which demonstrated both SA1 and SA2 bind at regularly spaced intervals corresponding to the positions of the tandemly ligated R-loops (Fig. 5B). A recent study confirmed that in cells SA1/SA2 localize to R-loops and are required for cohesin loading onto DNA in the absence of NIPBL. Furthermore, SA1 may play a key role in R-loop homeostasis. (153).

Figure 5: R-loops recruit SA1/SA2 and PARP1, and activate PARP1.

Figure 5:

(A) Diagram of the linear R-loop DNA (top) and representative AFM image of SA2 on the R-loop DNA (bottom) (52). (B) Diagrams of the DNA substrates (top left), representative fluorescence images (middle left) and kymographs (bottom left) of QD-labeled SA2, and analysis of the spacing between QD-labeled SA2 on R-loop and control DNA (right) (52). The distribution of distances between adjacent QD-SA2 on R-loop DNA tightropes is consistent with SA2 specifically binds to R-loops (marked by the red lines). (C) Model and representative AFM image of PARP1 binding linear R-loop DNA (left panel) and analysis of positions of R-loops, PARP1 on R-loop DNA and control DNA in the absence of NAD+ (right panel) (109). (D) Model and representative AFM image showing PARP1 auto-PARylation on the R-loop DNA in the presence of NAD+. Arrow: PAR chain (109).

ADP-ribosyltransferases (ARTs), which transfer either one (MARylate) or multiple ADP-ribose units (PARylate) from NAD+ onto substrates, play essential roles in regulating DNA damage response (DDR) (154159). ART family members, including PARP1, PARP2, and PARP3, are activated in the presence of damaged DNA. The resulting linear or branched synthesized PAR polymers are important for assembling and activating many DNA repair proteins at the damaged site (160164). PARP1 is one of the earliest proteins to arrive at a damage site during the DNA damage response (165). Previous AFM studies support this, demonstrating PARP1’s binding preference for DNA damage sites in vitro (19,24,162,166). A growing body of evidence suggests PARP1, R-loops, and R-loop resolving proteins overlap in several pathways (109,167169). However, the mechanisms by which PARP1 and R-loop resolving proteins interact at R-loops and how this contributes to R-loop resolution are not well understood.

AFM imaging in air revealed that PARP1 preferentially binds R-loops on linear dsDNA (Fig. 5C). Most PARP1 binding events without NAD+, occur at the position of the R-loop along dsDNA, yet are distributed randomly across control DNA (109). It is worth noting that the identification of PARP1-R-loop complexes on dsDNA was based on the significantly increased height compared to R-loop alone. In some cases, PARP1 can be identified binding to the “spur” R-loop structure, as seen in Fig. 5C. Interestingly, binding to the R-loop in the presence of NAD+ activated PARP1’s PARylation activity, which is visible on AFM as branches budding out from PARP1 bound at the R-loop (Fig. 5D). Additionally, AFM revealed that PARP1 can bridge two R-loop DNA molecules together at the position of the R-loops (109). These findings were validated using a PARP1 knockout cell line, in which it was found that R-loops accumulate in the absence of PARP1. These results suggest that activation of PARP1 at R-loops could serve to suppress R-loop formation or facilitate the resolution of R-loops (109).

5.3. Dynamics of TRF1 and TRF2 binding to telomeric DNA

Telomeres are nucleoprotein structures that assemble at the ends of linear chromosomes (170) and prevent chromosomal ends from erroneously initiating DNA damage checkpoints, recombination, and end-fusion (170177). Telomeric sequences are bound by the protective shelterin complex, which contains six protein constituents (Fig. 6A), including Telomeric Repeat Binding Factors-1 and −2 (TRF1 and TRF2), TRF1-Interacting Nuclear Protein-2 (TIN2), Protection of Telomeres-1 (POT1), Telomere Protection Protein-1 (TPP1), and Repressor/Activator Protein-1 (RAP1) (178). TRF1 and TRF2 homodimerize and directly bind telomere DNA, which consists of 2–20 kb tandem arrays of the non-coding (TTAGGG)N sequence that terminates with a 50–400 nucleotide-long G-rich 3’ overhang (179). Telomere DNA can assume a variety of higher-order structures that play important roles in regulating telomere maintenance, including the formation of G-quadruplex structures (180183) in the 3’ overhang and T-loops (171,184,185). T-loop formation is dependent on TRF2 (186) and occurs through invasion of the G-rich 3’ overhang into the telomere duplex DNA, forming a lasso-like loop (Fig. 6A) (184), as has previously been captured with AFM (187). Through sequestration of the 3’ overhang, T-loop formation is an important means by which chromosomal ends are distinguished from DNA double-stranded breaks, thus helping prevent aberrant initiation of the DNA damage response (185,186). Additionally, although long thought to be transcriptionally silent, sub-telomeric DNA is transcribed to long noncoding Telomeric Repeat-containing RNA (TERRA) in mammalian cells, a significant portion of which remains associated with telomeres, either those from which it is transcribed (cis interacting) or with telomeres of other chromosomes (trans interacting) through poorly understood mechanisms (188,189). There, TERRA serves numerous important roles, including promoting homology-directed repair of telomeres (190) and telomere length maintenance (181), and serving as a binding scaffold for protein factors, such as TRF2 (191) and several DNA repair factors (192,193). Loss of TERRA has been shown to cause telomere structural aberrations (191) and promote the erosion of telomeric sequences (194,195), indicating TERRA plays essential roles in telomere maintenance.

Figure 6: DNA tightrope study of TRF1/TRF2 and the interaction between TRF1 and cohesin SA1.

Figure 6:

(A) Diagram of shelterin complex and T-loop (32). (B) Schematic of telomere DNA substrate [54]. (C) Representative fluorescence images of QD-TRF1 and QD-TRF2 on T270 tightropes. Their regular spacing shows specific binding of TRF1 and TRF2 to telomere regions at even intervals (79). (D) Representative kymographs for QD-labeled TRF1 and TRF2 on T270 DNA tightropes (79). (E) Representative kymograph showing QD-SA1 pauses while moving across telomere sequences (53). (F and G) Representative tightrope kymographs (F) and analysis of diffusion range (G) of QD-TRF1 (green) and QD-SA1 (red) on telomeric T270 DNA and genomic DNA (G-DNA) (53).

Additionally, numerous DNA repair proteins play important roles in telomere maintenance (170,171,196,197). While understanding of the DNA binding activities of individual telomere proteins is expanding (180,198,199), many holes remain in our understanding of the interplay between telomere proteins and DNA repair pathways. In particular, it is not well understood how telomere-binding proteins assemble on telomeric sequences to promote higher-order structures, like T-loops, and telomere compaction that regulate DNA repair protein access at telomeres (170,200). The complex array of interactions between shelterin components, DNA repair proteins, telomeric DNA, and TERRA are extensive, making it difficult to tease apart the functional roles and collaborative activities of each of the different components in telomere maintenance. As such, while ensemble biochemistry assays have revealed essential elements of telomere biology (170,201), they are limited in their abilities to capture the molecular mechanisms by which shelterin proteins associate with telomeric sequences and how the shelterin complex regulates the formation and maintenance of higher-order telomere structures. Further, while cell-based assays are essential approaches for studying the downstream effects of protein knockouts (173,191,202,203), T-loop formation, and protein localization to telomeres (188,204), teasing apart the contributions from the different constituents of shelterin with these methods is challenging. Some shelterin proteins bind DNA collaboratively, so that eliminating one protein may simultaneously remove several of its binding partners from DNA. In comparison, AFM and the DNA tightrope assay enable the binding dynamics and structures of shelterin’s constituents on telomeric sequences to be directly (and individually) studied in vitro.

To study telomere binding proteins using SMI, we utilize a DNA substrate (T270) that was initially designed with 270 repeats of the telomeric sequence by the de Lange lab (205). We linearized it to position the telomeric region at a desired location (Fig. 6B). This substrate can be used directly for AFM imaging or can be ligated in tandem for the DNA tightrope assay.

TRF1 and TRF2 homodimerize through their TRFH domain and recruit other shelterin components to telomeres (56,206). However, it was poorly understood how these proteins initially find telomeric sequences in the genome and subsequently recruit other shelterin and DNA repair proteins (170). A previous in cellulo study using fluorescence recovery after photobleaching suggested TRF1 and TRF2 bind telomeric sequences dynamically (207). We used the DNA tightrope assay to characterize the binding dynamics of individual TRF1 and TRF2 proteins on telomeric (Fig. 6B) and non-telomeric DNA tightropes. On T270 DNA tightropes, static TRF1 and TRF2 show spacing corresponding to the positions of telomere repeat sequences (Fig. 6C) (79). Interestingly, mobile TRF1 on telomeric regions displays a much narrower diffusion range than TRF2 (Fig. 6D, Video S3). Taken together, the DNA tightrope assay revealed that TRF1 and TRF2 use distinct search mechanisms to find telomere sequences, in which TRF1 binds DNA largely through 3D diffusion while TRF2 performs extensive 1D search on DNA through a sliding mechanism. We propose that individual TRF proteins use 1D sliding on DNA to find protein partners and assemble the shelterin complex, which in turn stabilizes the interactions on telomeric DNA (79).

It is worth noting that recent sequencing of the T270 plasmid using the Nanopore technology (Plasmidsaurus) revealed degeneracy in the telomeric sequences, which shows roughly 153 TTAGGG repeats and 126 TTGGGG repeats. While this new information should not affect the interpretation of results from previous studies using the T270 plasmid, it does raise the awareness that telomeric sequences are highly prone to mutations during replication inside bacteria. To compare across different studies, we need to monitor and report the telomeric sequences when using plasmid-based DNA substrates.

5.4. Higher-order telomere structures

To study how TRF2 promotes higher-order DNA structures, we first directly probed TRF2’s DNA compaction ability using DREEM imaging. We observed T270 DNA passing through small TRF2 complexes (208) and budding out from the edge of large TRF2 complexes (Fig. 4C), suggesting DNA is compacted within the complex. Additionally, DREEM imaging of TRF2’s homodimerization domain (TRFH) on telomeric DNA revealed that TRFH wraps DNA around itself (Fig. 5D) (171). These findings suggest that TRF2 condenses telomeric DNA by wrapping it around individual TRFH domains inside TRF2 and by corralling it into protein-free loops that protrude from multi-TRF2 complexes.

In cellulo, most TRF1 and TRF2 homodimers are complexed with TIN2 (209). While TIN2 itself cannot bind DNA, it is known as the linchpin of the shelterin complex as it bridges TRF1 to TRF2, stabilizing their interaction on telomere sequences and with other protein components of shelterin (Fig. 6A) (170,209212). This stabilization of TRF1 and TRF2 on telomeres is essential to prevent triggering a DNA damage response (212). However, these very attributes that make TIN2 essential for telomere biology also complicate its study: because it interacts with several shelterin components, it is difficult to characterize all its functional roles with ensemble biochemistry approaches. Because TIN2 facilitates TRF1/TRF2 association with telomeres, the knockdown of TIN2 in cells also removes TRF1/TRF2 from DNA, confounding the interpretation of observed downstream effects.

Using both short (S) and long (L) TIN2 isomers, DNA tightrope studies showed that TRF2 is required for the recruitment of TIN2 to telomere sequences in vitro (213). We next sought to determine how TIN2 affects TRF2 DNA binding. Previous research using biochemical assays (214) and AFM (214,215) suggest that TRF1 and TRF2 compact short telomeric DNA substrates. One technical barrier in the single-molecule study of DNA compaction by the telomere system is the availability of long native telomeric DNA (>2 kb) with high purity. To overcome this technical barrier, we optimized the DNA target capture technology, the Region-Specific Extraction (RSE) system, for purifying longer telomeric DNA (Fig. 7A). We first digested genomic DNA purified from mouse liver tissues using a restriction enzyme cocktail, followed by the hybridization of a complementary capture primer to the 3’ overhang of the telomere DNA, extension of the primer using PCR reactions containing biotinylated-dNTP, and purification using streptavidin-coated magnetic beads (213). PCR-extended primers contain multiple biotinylated nucleotides that provide high stability and accessibility to streptavidin-coated magnetic beads. From 125 μg of mouse genomic DNA, the RSE protocol generated ~0.18 ug of purified mouse telomeric DNA with lengths equivalent to λ DNA (~48 kbp), and its purity was validated using a dot blot assay (213). We characterized the effect of TIN2 on TRF2-mediated DNA compaction using the surface-anchored and flow-stretched telomeric DNA purified using the RSE method. We measured the change in end-to-end length of YOYO1-labeled long mouse telomeric DNA as an indicator of DNA compaction and found that TIN2L and TIN2S isomers significantly enhanced TRF2-mediated telomeric DNA compaction (Fig. 7B&C) (213). This result was corroborated by observations from AFM imaging.

Figure 7: TIN2 enhances TRF2-mediated DNA compaction, bridging of single-stranded telomeric RNA to telomere sequences, and T-loop formation.

Figure 7:

(A) The RSE method for purifying long telomeric DNA from mouse liver (213). (B and C) QD-TIN2 (purple) enhances TRF2-mediated compaction of YOYO1-labeled long telomere DNA (tDNA) purified using the RSE method. Black arrows mark the DNA anchoring proint. White arrows mark reversal in buffer flow direction to validate the absence of nonspecific interactions between DNA and surface (213). (D) Schematic of the DNA tightrope assay for monitoring the bridging of single-stranded telomeric RNA (TERRA) onto DNA tightropes. (E) example images (top) and kymographs (bottom) of QD-labeled telomeric RNA on DNA tightropes in the presence of TRF2 or TRF2-TIN2S (213). (F) Representative AFM images showing TRF2 and TRF2-TIN2L mediated T-loops on linear T270 DNA. A QD in the top image marks the non-telomeric end (213).

TRF2 directly binds to TERRA (191,216). To investigate if TIN2 promotes the recruitment of TERRA to telomeric DNA in trans, we used a modification of the tightrope assay. We labeled biotinylated ssDNA or ssRNA (TERRA) molecules containing twelve TTAGGG or UUAGGG repeats, respectively, with QDs (Fig. 7D) and introduced them to a flow cell containing ligated T270 DNA tightropes. Without proteins, no signal from telomeric ssDNA or ssRNA accumulated on the tightropes, indicating any bridging we might observe is protein-dependent. TRF2 successfully recruits TERRA molecules to the tightropes at spacing consistent with the distance between telomeric regions (Fig. 7E) (213); the QD signal remains at the tightropes for the extent of the 2-minute recording window, indicating the interaction is highly stable. In the presence of TIN2, the intensity of QD recruited to T270 tightropes increases significantly, indicating TIN2 induces cluster formation at telomere sequences by enhancing TRF2-mediated bridging of telomeric ssDNA and ssRNA. The bridging of telomeric ssRNA to telomeric duplex DNA mediated by TRF2 and TIN2 is consistent with a model in which TERRA serves as a scaffold for shelterin protein binding. These results also provide direct experimental evidence that TERRA expressed from one chromosome can act in trans at other chromosomal ends.

Given the collaborative activities of TRF2 and TIN2 on telomeric DNA that we were able to observe using AFM imaging, we next sought to investigate if TIN2 modulates TRF2-mediated T-loop formation. Importantly, previous research has demonstrated that TRF2 alone is sufficient for T-loop formation (198), but has relied on TIN2-knockouts that are also known to destabilize TRF2 at telomeres. To characterize the effect of TIN2 on TRF2-dependent T-loop formation, we generated a model T-loop substrate by incubating linear T270 DNA with an exonuclease, producing a G-rich 3’ overhang (213). We attached a QD via a biotin-tagged DNA primer to the opposite non-telomeric side to distinguish the telomeric end from the non-telomeric one (Fig. 7F top panel). Using AFM imaging in air, we observed TRF2 alone promotes T-loop formation at the 3’ overhang in approximately 10% of DNA molecules (213). In the presence of TIN2, however, the percent of DNA molecules with T-loops increases in a TIN2 concentration-dependent manner (213). We observed large multi-protein complexes of TRF2 and TIN2 form at the base of the T-loops (Fig. 7F bottom panel), which grew larger with increasing TIN2 concentrations. No T-loop formation was observed on control T270 DNA not treated with an exonuclease, indicating this collaborative TRF2-TIN2-mediated activity requires a 3’ overhang.

Importantly, in a separate study of TRF1 and TIN2 using AFM imaging and the DNA tightrope assay, we observed that TIN2 also facilitates TRF1-mediated telomeric DNA compaction and DNA-DNA bridging (56). These observations support a model that TIN2 serves as an architectural protein at telomeres by promoting TRF1 and TRF2-mediated higher-order DNA structures, including DNA compaction, bridging of TERRA to duplex telomeric DNA, and T-loop formation.

5.5. The interplay between cohesin SA1 and telomere binding protein TRF1

Cohesion between the telomere regions of sister chromatids is an important prerequisite to genome stability, as it promotes proper chromosome segregation (217). However, the mechanism by which this occurs at telomeres was not well understood. Previous research has shown that cohesin’s SA1 subunit, TRF1, and TIN2 interact (203). Sister telomere association is largely dependent on SA1, and not the cohesin ring subunits (SMC1, SMC3, and RAD21) (218). Given the importance of SA1 for telomere cohesion and its interaction with shelterin proteins TRF1 and TIN2, we sought to determine how these proteins collaboratively maintain telomere DNA-DNA bridging, which is a key activity during sister telomere cohesion. Having discovered using AFM imaging in air that SA1 directly binds telomere repeat sequences in vitro (53), we used the tightrope assay to characterize how SA1 finds telomere sequences and if TRF1 influences this process. As anticipated, we observed QD-labeled SA1 binds to tightropes of ligated T270 DNA with spacing corresponding to the positions of telomeric sequences. Some QD-SA1 proteins bound telomere sequences statically, while others remain mobile, covering distances up to 8.55 μm along the tightropes (Fig. 6E&G), a distance that would include several telomere sequence repeats. A subpopulation of mobile QD-SA1 molecules alternates between fast and slow diffusion, a behavior known as bimodal diffusion; periods of slow diffusion correspond to the positions of telomere sequences (Fig. 6E and Video S4). This result suggests that SA1 alone diffuses along DNA containing telomeric and non-telomeric sequences and pauses as it passes over telomere sequences. We next repeated the tightrope assay with SA1 and TRF1 differentially conjugated to red and green quantum dots, respectively. Under these conditions, a subpopulation of QD bound to the tightropes was dual-colored, indicating QD-TRF1 and QD-SA1 are bound to the same telomere repeat sequence (Fig. 6F). When complexed with TRF1 in this way, a greater percentage of SA1 binding events are static than for QD-SA1 alone and bound with spacing corresponding to the telomere sequence spacing. Furthermore, the SA1-TRF1 complexes that remained mobile possess a much narrower diffusion range of approximately 0.5 μm (Fig. 6G); this confined distance is less than the length of one (TTAGGG)270 region on the tightrope. On non-telomeric control DNA, most mobile SA1-TRF1 complexes are not confined to such a narrow diffusion range and display much higher diffusion constants and alpha factors than on telomeric DNA. In summary, evidence from the tightrope assay revealed that TRF1 interacts with SA1 to enhance stable SA1 association with telomere sequences. Furthermore, AFM imaging revealed that SA1 enhances TRF1-mediated telomeric DNA-DNA bridging (53). Similar to what we observed on telomeric DNA, we speculate a longer dwell time of SA1 at AT-rich sequences (in its recognition mode) could stabilize the cohesin complex for carrying out unique functions at these regions across the genome. We propose a model in which SA1 and SA2 are the “DNA sequence and structure guide” for the cohesin complex, which directs its loading at specific sequences/structures along the genome.

6. Outlook

Synergistic with high-resolution structure determination using cryo-EM, HS-AFM will continue to shed unprecedented insight into the dynamic of multi-protein DNA repair complexes. While AFM resolution is limited by the radius of the tip, localization AFM (LAFM) (219) and simulation AFM (220) further improve image resolution for biological samples. By applying localization image reconstruction algorithms, simulation AFM and automated fitting to experimental images, these analytical approaches thus reveal yet finer details of protein-DNA interactions. Another promising development in AFM technology is nanoendoscopy-AFM, which inserts a nanoprobe into living cells and images their interior while preserving cellular viability (221). Molecular resolution imaging of intracellular or even intranuclear dynamics of DNA repair proteins in live cells would contribute fascinating research to the DNA repair field. Furthermore, the recently developed Single-Molecule Analysis of DNA-binding proteins from Nuclear Extracts (SMADNE) promises to expand the capacity of the DNA tightrope assay to probing of native protein-DNA interactions in the context of physiologically-relevant nuclear proteins (222). Furthermore, laser-trap based SMADNE enabled studies of protein-DNA interactions under variable tension. For example, using SMADNE, it was demonstrated that increasing tension on nicked DNA leads to more binding events by PARP1 (222). With new generations of high-speed and high-sensitivity scientific CMOS cameras, brighter and more stable fluorophores will break the temporal resolution bottleneck of the fluorescence imaging-based DNA tightrope assay. In summary, SMI techniques, including AFM and the tightrope assay, continue to evolve and enhance spatial and temporal resolutions of protein-DNA complexes, which will drive the new frontiers in the DNA repair research field.

Supplementary Material

1
2
Download video file (7.2MB, mp4)
3
Download video file (4MB, avi)
4
Download video file (7.6MB, wmv)
5
Download video file (9.5MB, mp4)

Highlights.

  • High-speed AFM and the DNA tightrope assays provide novel findings regarding the structure and dynamics of protein-DNA interactions.

  • Unique substrates containing specific DNA sequences and structures are generated to emulate DNA repair or replication intermediates.

  • Higher-order telomeric DNA structures, such as DNA compaction and T-loop, are formed through multi-protein interaction networks.

  • Twinkle helicase self-loads onto DNA through domain protrusion, and mitochondrial SSB wraps DNA.

  • R-loops recruit cohesin and activate PARP1.

Acknowledgements

We thank Ingrid Tessmer for her thoughtful comments. This work was supported by the National Institutes of Health [R01GM123246 and P30 ES025128 Pilot Project Grants through the Center for Human Health and the Environment at NCSU to H.W., and NIEHS training grant T32ES007046 to E.M.I.], and the Goodnight Doctoral Fellowship to E.M.I..

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Conflict of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Author Statement

All data related to cited figures will be shared upon request (corresponding author).

References

  • 1.Ciccia A, and Elledge SJ (2010) The DNA Damage Response: Making It Safe to Play with Knives. Mol Cell 40, 179–204 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Friedberg EC (2008) A brief history of the DNA repair field. Cell Res 18, 3–7 [DOI] [PubMed] [Google Scholar]
  • 3.Howard MJ, and Wilson SH (2018) DNA scanning by base excision repair enzymes and implications for pathway coordination. DNA Repair 71, 101–107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Kunkel TA, and Erie DA (2005) DNA mismatch repair. Annu Rev Biochem 74, 681–710 [DOI] [PubMed] [Google Scholar]
  • 5.Iyer RR, Pluciennik A, Burdett V, and Modrich PL (2006) DNA mismatch repair: functions and mechanisms. Chem Rev 106, 302–323 [DOI] [PubMed] [Google Scholar]
  • 6.van Houten B, Croteau DL, DellaVecchia MJ, Wang H, and Kisker C (2005) ‘Close-fitting sleeves’: DNA damage recognition by the UvrABC nuclease system. Mutat Res 577, 92–117 [DOI] [PubMed] [Google Scholar]
  • 7.Liao Y, Schroeder JW, Gao B, Simmons LA, and Biteen JS (2015) Single-molecule motions and interactions in live cells reveal target search dynamics in mismatch repair. PNAS 112, E6898–E6906 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Henrikus SS, Henry C, McGrath AE, Jergic S, McDonald JP, Hellmich Y, Bruckbauer ST, Ritger ML, Cherry ME, Wood EA, Pham PT, Goodman MF, Woodgate R, Cox MM, van Oijen AM, Ghodke H, and Robinson A (2020) Single-molecule live-cell imaging reveals RecB-dependent function of DNA polymerase IV in double strand break repair. Nucleic Acids Res 48, 8490–8508 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Ho HN, van Oijen AM, and Ghodke H (2020) Single-molecule imaging reveals molecular coupling between transcription and DNA repair machinery in live cells [DOI] [PMC free article] [PubMed]
  • 10.Ghodke H, Ho HN, and van Oijen AM (2020) Single-molecule live-cell imaging visualizes parallel pathways of prokaryotic nucleotide excision repair. Nat Commun 11, 1477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Opresko PL, Otterlei M, Graakjaer J, Bruheim P, Dawut L, Kolvraa S, May A, Seidman MM, and Bohr VA (2004) The Werner syndrome helicase and exonuclease cooperate to resolve telomeric D loops in a manner regulated by TRF1 and TRF2. Mol Cell 14, 763–774 [DOI] [PubMed] [Google Scholar]
  • 12.Gassman NR, and Wilson SH (2015) Micro-irradiation tools to visualize base excision repair and single-strand break repair. DNA Repair 31, 52–63 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Liu Z, Gong Z, Cao Y, Ding YH, Dong MQ, Lu YB, Zhang WP, and Tang C (2018) Characterizing Protein Dynamics with Integrative Use of Bulk and Single-Molecule Techniques. Biochemistry 57, 305–313 [DOI] [PubMed] [Google Scholar]
  • 14.Bianco PR (2022) Insight into the biochemical mechanism of DNA helicases provided by bulk-phase and single-molecule assays Methods 204, 348–360 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Deguchi T, Iwanski MK, Schentarra EM, Heidebrecht C, Schmidt L, Heck J, Weihs T, Schnorrenberg S, Hoess P, Liu S, Chevyreva V, Noh KM, Kapitein L, and Ries J (2023) Direct observation of motor protein stepping in living cells using MINFLUX. Science 379, 1010–1015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Wolff JO, Scheiderer L, Engelhardt T, Engelhardt J, Matthias J, and Hell SW (2023) MINFLUX dissects the unimpeded walking of kinesin-1. Science 379, 1004–1010 [DOI] [PubMed] [Google Scholar]
  • 17.Xue C, Daley JM, Xue X, Steinfeld J, Kwon Y, Sung P, and Greene EC (2019) Single-molecule visualization of human BLM helicase as it acts upon double- and single-stranded DNA substrates. Nucleic Acids Res 47, 11225–11237 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Wu WQ, Hou XM, Zhang B, Fosse P, Rene B, Mauffret O, Li M, Dou SX, and Xi XG (2017) Single-molecule studies reveal reciprocating of WRN helicase core along ssDNA during DNA unwinding. Sci Rep 7, 43954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Sukhanova MV, Hamon L, Kutuzov MM, Joshi V, Abrakhi S, Dobra I, Curmi PA, Pastre D, and Lavrik OI (2019) A Single-Molecule Atomic Force Microscopy Study of PARP1 and PARP2 Recognition of Base Excision Repair DNA Intermediates. J Mol Biol 431, 2655–2673 [DOI] [PubMed] [Google Scholar]
  • 20.Davidson IF, Goetz D, Zaczek MP, Molodtsov MI, Huis In ‘t Veld PJ, Weissmann F, Litos G, Cisneros DA, Ocampo-Hafalla M, Ladurner R, Uhlmann F, Vaziri A, and Peters JM (2016) Rapid movement and transcriptional re-localization of human cohesin on DNA. EMBO J 35, 2671–2685 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Kim Y, Shi Z, Zhang H, Finkelstein IJ, and Yu H (2019) Human cohesin compacts DNA by loop extrusion. Science 366, 1345–1349 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Hwang H, Opresko P, and Myong S (2014) Single-molecule real-time detection of telomerase extension activity. Sci Rep 4, 6391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Gauer JW, LeBlanc S, Hao P, Qiu R, Case BC, Sakato M, Hingorani MM, Erie DA, and Weninger KR (2016) Single-Molecule FRET to Measure Conformational Dynamics of DNA Mismatch Repair Proteins. Methods Enzymol 581, 285–315 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Liu L, Kong M, Gassman NR, Freudenthal BD, Prasad R, Zhen S, Watkins SC, Wilson SH, and Van Houten B (2017) PARP1 changes from three-dimensional DNA damage searching to one-dimensional diffusion after auto-PARylation or in the presence of APE1. Nucleic Acids Res 45, 12834–12847 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Tyagi S, VanDelinder V, Banterle N, Fuertes G, Milles S, Agez M, and Lemke EA (2014) Continuous throughput and long-term observation of single-molecule FRET without immobilization. Nat Methods 11, 297–300 [DOI] [PubMed] [Google Scholar]
  • 26.Banigan EJ, van den Berg AA, Brandao HB, Marko JF, and Mirny LA (2020) Chromosome organization by one-sided and two-sided loop extrusion. Elife 9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Higashi TL, Pobegalov G, Tang M, Molodtsov MI, and Uhlmann F (2021) A Brownian ratchet model for DNA loop extrusion by the cohesin complex. Elife 10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Yang Y, Wang H, and Erie DA (2003) Quantitative characterization of biomolecular assemblies and interactions using atomic force microscopy. Methods 29, 175–187 [DOI] [PubMed] [Google Scholar]
  • 29.LeBlanc S, Wilkins H, Li Z, Kaur P, Wang H, and Erie DA (2017) Using Atomic Force Microscopy to Characterize the Conformational Properties of Proteins and Protein–DNA Complexes That Carry Out DNA Repair. Methods Enzymol 592, 187–212 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Wang H, Yang Y, and Erie DA (2007) Characterization of Protein–Protein Interactions Using Atomic Force Microscopy. in Protein Interactions: Biophysical Approaches for the Study of Complex Reversible Systems
  • 31.Kaur P, Pan H, Longley MJ, Copeland WC, and Wang H (2021) Using Atomic Force Microscopy to Study the Real Time Dynamics of DNA Unwinding by Mitochondrial Twinkle Helicase. Bio-protocol 11, 4139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Lin J, Kaur P, Countryman P, Opresko PL, and Wang H (2014) Unraveling secrets of telomeres: One molecule at a time. DNA Repair 20, 142–153 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Kong M, Liu L, Chen X, Driscoll KI, Mao P, Bohm S, Kad NM, Watkins SC, Bernstein KA, Wyrick JJ, Min JH, and van Houten B (2016) Single-Molecule Imaging Reveals that Rad4 Employs a Dynamic DNA Damage Recognition Process. Mol Cell 64, 376–387 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Rill N, Mukhortava A, Lorenz S, and Tessmer I (2020) Alkyltransferase-like protein clusters scan DNA rapidlyover long distances and recruit NER to alkyl-DNA lesions. PNAS 117, 9318–9328 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Gross J, Wirth N, and Tessmer I (2017) Atomic Force Microscopy Investigations of DNA Lesion Recognition in Nucleotide Excision Repair. J. Vis. Exp 123 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Beckwitt EC, Kong M, and van Houten B (2018) Studying protein-DNA interactions using atomic force microscopy. Semin Cell Dev Biol 73, 220–230 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Bradford KC, Wilkins H, Hao P, Li ZM, Wang B, Burke D, Wu D, Smith AE, Spaller L, Du C, Gauer JW, Chan E, Hsieh P, Weninger K, and Erie DA (2020) Dynamic human MutSα–MutLα complexes compact mismatched DNA. Biochemistry 117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Sacho EJ, Kadyrov FA, Modrich P, Kunkel TA, and Erie DA (2008) Direct Visualization of Asymmetric Adenine Nucleotide-Induced Conformational Changes in MutLa. Mol Cell 29, 112–121 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Tse ECM, Zwang TJ, and Barton JK (2017) The Oxidation State of [4Fe4S] Clusters Modulates the DNA-Binding Affinity of DNA Repair Proteins. J Am Chem Soc 139, 12784–12792 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Nagpal A, Raja S, and van Houten B (2022) The role of UV-DDB in processing 8-oxoguanine during base excision repair. Biochem Soc Trans 50, 1481–1488 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Bangalore DM, Heil HS, Mehringer CF, Hirsch L, Hemmen K, Heinze KG, and Tessmer I (2020) Automated AFM analysis of DNA bending reveals initial lesion sensing strategies of DNA glycosylases. Sci Rep 10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Gupta R, Unciuleac MC, Shuman S, and Glickman MS (2017) Homologous recombination mediated by the mycobacterial AdnAB helicase without end resection by the AdnAB nucleases. Nucleic Acids Res 45, 762–774 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Moreno-Herrero F, de Jager M, Dekker NH, Kanaar R, Wyman C, and Dekker C (2005) Mesoscale conformational changes in the DNA-repair complex Rad50/Mre11/Nbs1 upon binding DNA. Nature 437, 440–443 [DOI] [PubMed] [Google Scholar]
  • 44.Ando T (2012) High-speed atomic force microscopy coming of age. Nanotechnology 23, 062001. [DOI] [PubMed] [Google Scholar]
  • 45.Finzi L, and Dunlap DD (2010) Single-molecule Approaches to Probe the Structure, Kinetics, and Thermodynamics of Nucleoprotein Complexes That Regulate Transcription. J Biol Chem 285, 18973–18978 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Janicijevic A, Ristic D, and Wyman C (2003) The molecular machines of DNA repair: scanning force microscopy analysis of their architecture. J. Microsc 212, 264–272 [DOI] [PubMed] [Google Scholar]
  • 47.Shlyakhtenko LS, Gall AA, and Lyubchenko YL (2013) Mica functionalization for imaging of DNA and protein-DNA complexes with atomic force microscopy. Methods Mol Biol 931, 295–312 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Rivetti C, Guthold M, and Bustamante C (1996) Scanning Force Microscopy of DNA Deposited onto Mica: Equilibration versus Kinetic Trapping Studied by Statistical Polymer Chain Analysis. J Mol Biol 264, 919–932 [DOI] [PubMed] [Google Scholar]
  • 49.Voigtlander B (2019) Atomic Force Microscopy. (Technology, S. N. a ed., 2 Ed., Springer Nature; Switzerland AG [Google Scholar]
  • 50.Kaur P, Longley MJ, Pan H, Wang H, and Copeland WC (2018) Single-molecule DREEM imaging reveals DNA wrapping around human mitochondrial single-stranded DNA binding protein. Nucleic Acids Res 46, 11287–11302 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Countryman P, Fan Y, Gorthi A, Pan H, Strickland J, Kaur P, Wang X, Lin J, Lei X, White C, You C, Wirth N, Tessmer I, Piehler J, Riehn R, Bishop AJR, Tao YJ, and Wang H (2018) Cohesin SA2 is a sequence-independent DNA-binding protein that recognizes DNA replication and repair intermediates. The Journal of biological chemistry 293, 1054–1069 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Pan H, Jin M, Ghadiyaram A, Kaur P, Miller HE, Ta HM, Liu M, Fan Y, Mahn C, Gorthi A, You C, Piehler J, Riehn R, Bishop AJR, Tao YJ, and Wang H (2020) Cohesin SA1 and SA2 are RNA binding proteins that localize to RNA containing regions on DNA. Nucleic acids research 48, 5639–5655 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Lin J, Countryman P, Chen H, Pan H, Fan Y, Jiang Y, Kaur P, Miao W, Gurgel G, You C, Piehler J, Kad NM, Riehn R, Opresko P, Smith S, Tao YJ, and Wang H (2016) Functional interplay between SA1 and TRF1 in telomeric DNA binding and DNA-DNA pairing. Nucleic Acids Res 44, 6363–6376 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Liu M, Movahed S, Dangi S, Pan H, Kaur P, Bilinovich SM, Faison EM, Leighton GO, Wang H, Williams DC Jr., and Riehn R (2020) DNA looping by two 5-methylcytosine-binding proteins quantifed using nanofuidic devices. Epigenetics Chromatin 13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Mawhinney MT, Liu R, Lu F, Maksimoska J, Damico K, Marmorstein R, Lieberman PM, and Urbanc B (2018) CTCF-Induced Circular DNA Complexes Observed by Atomic Force Microscopy. J Mol Biol 430, 759–776 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Pan H, Kaur P, Barnes R, Detwiler AC, Sanford SL, Liu M, Xu P, Mahn C, Tang Q, Hao P, Bhattaram D, You C, Gu X, Lu W, Piehler J, Xu G, Weninger K, Riehn R, Opresko PL, and Wang H (2021) Structure, dynamics, and regulation of TRF1-TIN2-mediated trans- and cis-interactions on telomeric DNA. J Biol Chem 297, 101080. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Wagner C, Olbrich C, Brutzer H, Salomo M, Kleinekathofer U, Keyser UF, and Kremer F (2011) DNA condensation by TmHU studied by optical tweezers, AFM and molecular dynamics simulations. J Biol Phys 37, 117–131 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Zabolotnaya E, Mela I, Williamson MJ, Bray SM, Yau SK, Papatziamou D, Edwardson JM, Robinson NP, and Henderson RM (2020) Modes of action of the archaeal Mre11/Rad50 DNA-repair complex revealed by fast-scan atomic force microscopy. PNAS 117, 14936–14974 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Hansma HG, Bezanilla M, Zenhausern F, Adrian M, and Sinsheimer RL (1993) Atomic force microscopy of DNA in aqueous solutions. Nucleic Acids Res 21, 505–512 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Kasas S, Thomson NH, Smith BL, Hansma HG, Zhu X, Guthold M, Bustamante C, Kool ET, Kashlev M, and Hansma PK (1997) Escherichia coli RNA Polymerase Activity Observed Using Atomic Force Microscopy. Biochemistry 36, 461–468 [DOI] [PubMed] [Google Scholar]
  • 61.Ohkubo T, Shiina T, Kawaguchi K, Sasaki D, Inamasu R, Yang Y, Li Z, Taninaka K, Sakaguchi M, Fujimura S, Sekiguchi H, Kuramochi M, Arai T, Tsuda S, Sasaki YC, and Mio K (2022) Visualizing Intramolecular Dynamics of Membrane Proteins. IJMS 23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Lyubchenko YL (2018) Direct AFM visualization of the nanoscale dynamics of biomolecular complexes. J. Phys. D: Appl. Phys 51 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Muller DJ, and Engel A (2007) Atomic force microscopy and spectroscopy of native membrane proteins. Nat Protoc 2, 2191–2197 [DOI] [PubMed] [Google Scholar]
  • 64.Bauer BW, Davidson IF, Canena D, Wutz G, Tang W, Litos G, Horn S, Hinterdorfer P, and Peters JM (2021) Cohesin mediates DNA loop extrusion by a “swing and clamp” mechanism. Cell 184, 5448–5464 e5422 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Tatebe H, Lim CT, Konno H, Shiozaki K, Shinohara A, Uchihashi T, and Furukohri A (2020) Rad50 zinc hook functions as a constitutive dimerization module interchangeable with SMC hinge. Nat Commun 11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Perrino AP, Miyagi A, and Scheuring S (2021) Single molecule kinetics of bacteriorhodopsin by HS-AFM. Nat Commun 12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Kodera N, Noshiro D, Dora SK, Mori T, Habchi J, Blocquel D, Gruet A, Dosnon M, Salladini E, Bignon C, Fujioka Y, Oda T, Noda NN, Sato M, Lotti M, Mizuguchi M, Longhi S, and Ando T (2021) Structural and dynamics analysis of intrinsically disordered proteins by high-speed atomic force microscopy. Nat Nanotechnol 16, 181–189 [DOI] [PubMed] [Google Scholar]
  • 68.Matin TR, Heath GR, Huysmans GHM, Boudker O, and Scheuring S (2020) Millisecond dynamics of an unlabeled amino acid transporter. Nat Commun 11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Ishino S, Yamagami T, Kitamura M, Kodera N, Mori T, Sugiyama S, Ando T, Goda N, Tenno T, Hiroaki H, and Ishino Y (2014) Multiple interactions of the intrinsically disordered region between the helicase and nuclease domains of the archaeal Hef protein. J Biol Chem 289, 21627–21639 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Ruan Y, Miyagi A, Wang X, Chami M, Boudker O, and Scheuring S (2017) Direct visualization of glutamate transporter elevatormechanism by high-speed AFM. PNAS 114, 1584–1588 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Sakiyama Y, Mazur A, Kapinos LE, and Lim RYH (2016) Spatiotemporal dynamics of the nuclear pore complex transport barrier resolved by high-speed atomic force microscopy. Nat Nanotechnol 11, 719–723 [DOI] [PubMed] [Google Scholar]
  • 72.Ogawa T, Saijo S, Shimizu N, Jiang X, and Hirokawa N (2017) Mechanism of Catalytic Microtubule Depolymerization via KIF2-Tubulin Transitional Conformation. Cell Rep 20, 2626–2638 [DOI] [PubMed] [Google Scholar]
  • 73.Kodera N, Yamamoto D, Ishikawa R, and Ando T (2010) Video imaging of walking myosin V by high-speed atomic force microscopy. Nature 468, 72–76 [DOI] [PubMed] [Google Scholar]
  • 74.Uchihashi T, Iion R, Ando T, and Noji H (2011) High-Speed Atomic Force Microscopy Reveals Rotary Catalysis of Rotorless F1-ATPase. Science 333, 755–758 [DOI] [PubMed] [Google Scholar]
  • 75.Shibata M, Nishimasu H, Kodera N, Hirano S, Ando T, Uchihashi T, and Nureki O (2017) Real-space and real-time dynamics of CRISPR-Cas9 visualized by high-speed atomic force microscopy. Nat Commun 8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Wu D, Kaur P, Li ZM, Bradford KC, Wang H, and Erie DA (2016) Visualizing the Path of DNA through Proteins Using DREEM Imaging. Mol Cell 61, 315–323 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Kad NM, Wang H, Kennedy GG, Warshaw DM, and van Houten B (2010) Collaborative Dynamic DNA Scanning by Nucleotide Excision Repair Proteins Investigated by Single-Molecule Imaging of Quantum-Dot-Labeled Proteins. Mol Cell 37, 702–713 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Tokunaga M, Imamoto N, and Sakata-Sogawa K (2008) Highly inclined thin illumination enables clear single-molecule imaging in cells. Nat Methods 5, 159–161 [DOI] [PubMed] [Google Scholar]
  • 79.Lin J, Countryman P, Buncher N, Kaur PE,L, Zhang Y, Gibson G, You C, Watkins SC, Piehler J, Opresko PL, Kad NM, and Wang H (2014) TRF1 and TRF2 use different mechanisms to find telomeric DNA but share a novel mechanism to search for protein partners at telomeres. Nucleic Acids Res 42, 2493–2504 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Vemulapalli S, Hashemi M, and Lyubchenko YL (2022) Site-Search Process for Synaptic Protein-DNA Complexes. IJMS 23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Stigler J, Camdere GO, Koshland DE, and Greene EC (2016) Single-Molecule Imaging Reveals a Collapsed Conformational State for DNA-Bound Cohesin. Cell Rep 15, 988–998 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Spenkelink LM, Lewis JS, Jergic S, Xu ZQ, Robinson A, Dixon NE, and van Oijen AM (2019) Recycling of single-stranded DNA-binding protein by the bacterial replisome 47, 4111–4123 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Marechal A, and Zou L (2015) RPA-coated single-stranded DNA as a platform for post-translational modifications in the DNA damage response. Cell Res 25, 9–23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Dickey TH, Altschuler SE, and Wuttke DS (2013) Single-Stranded DNA-Binding Proteins: Multiple Domains for Multiple Functions. Structure 21, 1074–1084 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Kowalczykowski SC (2004) An Overview of the Molecular Mechanisms of Recombinational DNA Repair. Cold Spring Harbor Perspectives in Biology 7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Marti TM, and Fleck O (2004) DNA repair nucleases. Cell Mol Life Sci 61, 336–354 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Raderschall E, Golub EI, and Haaf T (1999) Nuclear foci of mammalian recombination proteins are located at single-stranded DNA regions formed after DNA damage. PNAS 96, 1921–1926 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Ashton NW, Bolderson E, Cubeddu L, O’Byrne KJ, and Richard DJ (2013) Human single-stranded DNA binding proteins are essential for maintaining genomic stability. BMC Molecular Biol 14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Forget AL, and Kowalczykowski SC (2010) Single-molecule imaging brings Rad51 nucleoprotein filaments into focus. Trends Cell Biol 20, 269–276 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Forget AL, and Kowalczykowski SC (2010) Single-molecule imaging of DNA pairing by RecA reveals a three-dimensional homology search. Nature 482, 423–427 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Gibb B, Silverstein TD, Finkelstein IJ, and Greene EC (2012) Single-Stranded DNA Curtains for Real-Time Single-Molecule Visualization of Protein−Nucleic Acid Interactions. Anal Chem 84, 7607–7612 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Schaub JM, Soniat MM, and Finkelstein IJ (2022) Polymerase theta-helicase promotes end joining by stripping single-stranded DNA-binding proteins and bridging DNA ends. Nucleic Acids Res 50, 3911–3921 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Geng H, Du C, Chen S, Salerno V, Manfredi C, and Hsieh P (2011) In Vitro Studies of DNA Mismatch Repair Proteins. Anal Biochem 413, 179–184 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Buechner CN, and Tessmer I (2013) DNA substrate preparation for atomic force microscopy studies of protein–DNA interactions. J Mol Recognit 26, 605–617 [DOI] [PubMed] [Google Scholar]
  • 95.Kaur P (2020) Single-molecule level structural dynamics of DNA unwinding by human mitochondrial Twinkle helicase. J Biol Chem 295, 5564–5576 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Iyer DR, and Rhind N (2017) Replication fork slowing and stalling are distinct, checkpoint-independent consequences of replicating damaged DNA. PLoS Genet 13, e1006958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Merrick CJ, Jackson D, and Diffley JFX (2004) Visualization of Altered Replication Dynamics after DNA Damage in Human Cells. J Biol Chem 279, 20067–20075 [DOI] [PubMed] [Google Scholar]
  • 98.Chastain II PD, Heffernan TP, Nevis K, Lin L, Kaufmann WK, Kaufman DG, and Cordeiro-Stone M (2006) Checkpoint Regulation of Replication Dynamics in UV-Irradiated Human Cells 5, 2160–2167 [DOI] [PubMed] [Google Scholar]
  • 99.Kumar S, and Huberman JA (2008) Checkpoint-Dependent Regulation of Origin Firing and Replication Fork Movement in Response to DNA Damage in Fission Yeast. Mol Cell Biol 29, 602–611 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Qiu S, Jiang G, Cao L, and Huang J (2021) Replication Fork Reversal and Protection. Front Cell Dev Biol 9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Su F, Mukherjee S, Yang Y, Mori E, Bhattacharya S, Kobayashi J, Yannone SM, Chen DJ, and Asaithamby A (2014) Non-enzymatic Role for WRN in Preserving Nascent DNA Strands after Replication Stress. Cell Rep 9, 1387–1401 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Bhat KP, and Cortez D (2018) RPA and RAD51: fork reversal, fork protection, and genome stability. Nat Struct Mol Biol 25, 446–453 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Subramanian D, and Griffith JD (2005) p53 Monitors Replication Fork Regression by Binding to “Chickenfoot” Intermediates. Journal of Biological Chemistry 280, 42568–42572 [DOI] [PubMed] [Google Scholar]
  • 104.Costantino L, and Koshland D (2015) The Yin and Yang of R-loop biology. Curr Opin Cell Biol 34, 39–45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Niehrs C, and Luke B (2020) Regulatory R-loops as facilitators of gene expression and genome stability. Nat Rev Mol Cell Biol 21, 167–178 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Marnef A, and Legube G (2021) R-loops as Janus-faced modulators of DNA repair. Nat Cell Biol 23, 305–313 [DOI] [PubMed] [Google Scholar]
  • 107.Petermann E, Lan L, and Zou L (2022) Sources, resolution and physiological relevance of R-loops and RNA–DNA hybrids. Nat Rev Mol Cell Biol 23, 521–540 [DOI] [PubMed] [Google Scholar]
  • 108.Carrasco-Salas Y, Malapert A, Sulthana S, Molcrette B, Chazot-Franguiadakis L, Bernard P, Chedin F, Faivre-Moskalenko C, and Vanoosthuyse V (2019) The extruded non-template strand determines the architecture of R-loops. Nucleic Acids Res 47, 6783–6795 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Laspata N, Kaur P, Mersaoui SY, Muoio D, Liu ZS, Bannister MH, Nguyen HD, Curry C, Pascal JM, Poirier GG, Wang H, Masson JY, and Fouquerel E (2023) PARP1 associates with R-loops to promote their resolution and genome stability. Nucleic Acids Res 44, 6363–6373 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Anderson S, Bankier AT, Barrell BG, de Bruijn MHL, Coulson AR, Drouin J, Eperon IC, Nierlich DP, Roe BA, Sanger F, Schreier PH, Smith AJH, Staden R, and Young IG (1981) Sequence and organization of the human mitochondrial genome. Nature 290, 457–465 [DOI] [PubMed] [Google Scholar]
  • 111.Taanman JW (1999) The mitochondrial genome: structure, transcription, translation and replication. BBA - Bioenergetics 1410, 103–123 [DOI] [PubMed] [Google Scholar]
  • 112.DeBalsi KL, Hoff KE, and Copeland WC (2017) Role of the mitochondrial DNA replication machinery in mitochondrial DNA mutagenesis, aging and age-related diseases. Ageing Res Rev 33, 89–104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Ciesielski GL, Oliveira MT, and Kaguni LS (2016) Animal Mitochondrial DNA Replication. Enzymes 39, 255–292 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Ciesielski GL, Bermek O, Rosado-Ruiz FA, Hovde SL, Neitzke OJ, Griffith JD, and Kaguni LS (2015) Mitochondrial Single-stranded DNA-binding Proteins Stimulate the Activity of DNA Polymerase Y by Organization of the Template DNA. J Biol Chem 290, 28697–28707 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Plaza- G, Lemishko KM, Crespo R, Truong TQ, Kaguni LS, Cao-Garcia FJ, Ciesielski GL, and Ibarra B (2023) Mechanism of strand displacement DNA synthesis by the coordinated activities of human mitochondrial DNA polymerase and SSB. Nucleic Acids Res 51, 1750–1765 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Korhonen JA, Gaspari M, and Falkenberg M (2003) TWINKLE Has 5 3 3 DNA Helicase Activity and Is Specifically Stimulated by Mitochondrial Single-stranded DNA-binding Protein. J Biol Chem 278, 48627–48632 [DOI] [PubMed] [Google Scholar]
  • 117.Oliveira MT, and Kaguni LS (2011) Reduced stimulation of recombinant DNA polymerase γ and mitochondrial DNA (mtDNA) helicase by variants of mitochondrial single-stranded DNA-binding protein (mtSSB) correlates with defects in mtDNA replication in animal cells. J Biol Chem 286, 40649–40658 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Ruhanen H, Borrie S, Szabadkai G, Tyynismaa H, Jones AWE, Kang D, Tananman JW, and Yasukawa T (2010) Mitochondrial single-stranded DNA binding protein is required for maintenance of mitochondrial DNA and 7S DNA but is not required for mitochondrial nucleoid organisation. Biochim Biophys Acta 1803, 931–939 [DOI] [PubMed] [Google Scholar]
  • 119.Chen XJ (2013) Mechanism of homologous recombination and implications for aging-related deletions in mitochondrial DNA. Microbiol Mol Biol Rev 77, 476–496 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Sun Z, Tan HY, Bianco PR, and Lyubchenko YL (2015) Remodeling of RecG Helicase at the DNA Replication Fork by SSB Protein. Sci Rep 5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Sun Z, Hashemi M, Warren G, Bianco PR, and Lyubchenko YL (2018) Dynamics of the Interaction of RecG Protein with Stalled Replication Forks. Biochemistry 57, 1967–1976 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Wang Y, Sun Z, Bianco PR, and Lyubchenko YL (2020) Atomic force microscopy– based characterization of the interaction of PriA helicase with stalled DNA replication forks. J Biol Chem 295, 6043–6052 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Sun Z, Wang Y, Bianco PR, and Lyubchenko YL (2021) Dynamics of the PriA Helicase at Stalled DNA Replication Forks. J Phys Chem 125, 4299–4307 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Shlyakhtenko LS, Lushnikov AY, Miyagi A, and Lyubchenko YL (2012) Specificity of Binding of Single-Stranded DNA-Binding Protein to Its Target. Biochemistry 51, 1500–1509 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Hamon L, Pastre D, Dupaigne P, Le Breton C, Le Cam E, and Pietrement O (2007) High-resolution AFM imaging of single-stranded DNA-binding (SSB) protein— DNA complexes. Nucleic Acids Res 35 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Longley MJ, Humble MM, Sharief FS, and Copeland WC (2010) Disease Variants of the Human Mitochondrial DNA Helicase Encoded by C10orf2 Differentially Alter Protein Stability, Nucleotide Hydrolysis, and Helicase Activity. J Biol Chem 285, 29690–29702 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Naufer MN, Morse M, Moller GB, McIsaac J, Rouzina I, Beuning PJ, and Williams MC (2021) Multiprotein E. coli SSB–ssDNA complex shows both stable binding and rapid dissociation due to interprotein interactions. Nucleic Acids Res 49, 1532–1549 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Adkins NL, Swygert SG, Kaur P, Niu H, Grigoryev SA, Sung P, Wang H, and Peterson CL (2017) Nucleosome-like, Single-stranded DNA (ssDNA)-Histone Octamer Complexes and the Implication for DNA Double Strand Break Repair. J Biol Chem 292, 5271–5281 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Li Z, Kaur P, Lo CY, Chopra N, Smith J, Wang H, and Gao Y (2022) Structural and dynamic basis of DNA capture and translocation by mitochondrial Twinkle helicase. Nucleic acids research 50, 11965–11978 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Stanic M, and Mekhail K (2022) Integration of DNA damage responses with dynamic spatial genome organization. Trends in genetics : TIG 38, 290–304 [DOI] [PubMed] [Google Scholar]
  • 131.Szabo Q, Bantignies F, and Cavalli G (2019) Principles of genome folding into topologically associating domains. Sci Adv 5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Rao SSP, Huang SC, Glenn St Hilaire B, Engreitz JM, Perez EM, Kieffer-Kwon KR, Sanborn AL, Johnstone SE, Bascom GD, Bochkov ID, Huang X, Shamim MS, Shin J, Turner D, Ye Z, Omer AD, Robinson JT, Schlick T, Bernstein BE, Casellas R, Lander ES, and Aiden EL (2017) Cohesin Loss Eliminates All Loop Domains. Cell 171, 305–320 e324 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Schwarzer W, Abdennur N, Goloborodko A, Pekowska A, Fudenberg G, Loe-Mie Y, Fonseca NA, Huber W, Haering CH, Mirny L, and Spitz F (2017) Two independent modes of chromatin organization revealed by cohesin removal. Nature 551, 51–56 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Wutz G, Varnai C, Nagasaka K, Cisneros DA, Stocsits RR, Tang W, Schoenfelder S, Jessberger G, Muhar M, Hossain MJ, Walther N, Koch B, Kueblbeck M, Ellenberg J, Zuber J, Fraser P, and Peters JM (2017) Topologically associating domains and chromatin loops depend on cohesin and are regulated by CTCF, WAPL, and PDS5 proteins. EMBO J 36, 3573–3599 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Hirano T (2006) At the heart of the chromosome: SMC proteins in action. Nat Rev Mol Cell Biol 7, 311–322 [DOI] [PubMed] [Google Scholar]
  • 136.Canudas S, and Smith S (2009) Differential regulation of telomere and centromere cohesion by the Scc3 homologues SA1 and SA2, respectively, in human cells. J Cell Biol 187, 165–173 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Kim JS, Krasieva TB, LaMorte V, Taylor AM, and Yokomori K (2002) Specific recruitment of human cohesin to laser-induced DNA damage. J Biol Chem 277, 45149–45153 [DOI] [PubMed] [Google Scholar]
  • 138.Kong X, Ball AR Jr., Pham HX, Zeng W, Chen HY, Schmiesing JA, Kim JS, Berns M, and Yokomori K (2014) Distinct functions of human cohesin-SA1 and cohesin-SA2 in double-strand break repair. Mol Cell Biol 34, 685–698 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Meisenberg C, Pinder SI, Hopkins SR, Wooller SK, Benstead-Hume G, Pearl FMG, Jeggo PA, and Downs JA (2019) Repression of Transcription at DNA Breaks Requires Cohesin throughout Interphase and Prevents Genome Instability. Mol Cell 73, 212–223 e217 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Strom L, Karlsson C, Lindroos HB, Wedahl S, Katou Y, Shirahige K, and Sjogren C (2007) Postreplicative formation of cohesion is required for repair and induced by a single DNA break. Science 317, 242–245 [DOI] [PubMed] [Google Scholar]
  • 141.Unal E, Heidinger-Pauli JM, and Koshland D (2007) DNA double-strand breaks trigger genome-wide sister-chromatid cohesion through Eco1 (Ctf7). Science 317, 245–248 [DOI] [PubMed] [Google Scholar]
  • 142.Kim BJ, Li Y, Zhang J, Xi Y, Li Y, Yang T, Jung SY, Pan X, Chen R, Li W, Wang Y, and Qin J (2010) Genome-wide reinforcement of cohesin binding at pre-existing cohesin sites in response to ionizing radiation in human cells. J Biol Chem 285, 22784–22792 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Arnould C, Rocher V, Finoux AL, Clouaire T, Li K, Zhou F, Caron P, Mangeot PE, Ricci EP, Mourad R, Haber JE, Noordermeer D, and Legube G (2021) Loop extrusion as a mechanism for formation of DNA damage repair foci. Nature 590, 660–665 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Sjogren C, and Nasmyth K (2001) Sister chromatid cohesion is required for postreplicative double-strand break repair in Saccharomyces cerevisiae. Curr Biol 11, 991–995 [DOI] [PubMed] [Google Scholar]
  • 145.Bauerschmidt C, Arrichiello C, Burdak-Rothkamm S, Woodcock M, Hill MA, Stevens DL, and Rothkamm K (2010) Cohesin promotes the repair of ionizing radiation-induced DNA double-strand breaks in replicated chromatin. Nucleic Acids Res 38, 477–487 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Dorsett D, and Strom L (2012) The ancient and evolving roles of cohesin in gene expression and DNA repair. Curr Biol 22, R240–250 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Shi Z, Gao H, Bai XC, and Yu H (2020) Cryo-EM structure of the human cohesin-NIPBL-DNA complex. Science 368, 1454–1459 [DOI] [PubMed] [Google Scholar]
  • 148.Ouyang J, Yadav T, Zhang JM, Yang H, Rheinbay E, Guo H, Haber DA, Lan L, and Zou L (2021) RNA transcripts stimulate homologous recombination by forming DR-loops. Nature 594, 283–288 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Chen H, Yang H, Zhu X, Yadav T, Ouyang J, Truesdell SS, Tan J, Wang Y, Duan M, Wei L, Zou L, Levine AS, Vasudevan S, and Lan L (2020) m(5)C modification of mRNA serves a DNA damage code to promote homologous recombination. Nat Commun 11, 2834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Thapar R (2018) Regulation of DNA Double-Strand Break Repair by Non-Coding RNAs. Molecules 23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Lu WT, Hawley BR, Skalka GL, Baldock RA, Smith EM, Bader AS, Malewicz M, Watts FZ, Wilczynska A, and Bushell M (2018) Drosha drives the formation of DNA:RNA hybrids around DNA break sites to facilitate DNA repair. Nat Commun 9, 532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Michelini F, Pitchiaya S, Vitelli V, Sharma S, Gioia U, Pessina F, Cabrini M, Wang Y, Capozzo I, Iannelli F, Matti V, Francia S, Shivashankar GV, Walter NG, and d’Adda di Fagagna F (2017) Damage-induced lncRNAs control the DNA damage response through interaction with DDRNAs at individual double-strand breaks. Nat Cell Biol 19, 1400–1411 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Porter H, Li Y, Neguembor MV, Beltran M, Varsally W, Martin L, Cornejo MT, Pezic D, Bhamra A, Surinova S, Jenner RG, Cosma MP, and Hadjur S (2023) Cohesin-independent STAG proteins interact with RNA and R-loops and promote complex loading. Elife 12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Kruger A, Burkle A, Hauser K, and Mangerich A (2020) Real-time monitoring of PARP1-dependent PARylation by ATR-FTIR spectroscopy. Nat Commun 11, 2174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Alemasova EE, and Lavrik OI (2019) Poly(ADP-ribosyl)ation by PARP1: reaction mechanism and regulatory proteins. Nucleic Acids Res 47, 3811–3827 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Matta E, Kiribayeva A, Khassenov B, Matkarimov BT, and Ishchenko AA (2020) Insight into DNA substrate specificity of PARP1-catalysed DNA poly(ADP-ribosyl)ation. Sci Rep 10, 3699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Luo X, and Kraus WL (2012) On PAR with PARP: cellular stress signaling through poly(ADP-ribose) and PARP-1. Genes Dev 26, 417–432 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Luscher B, Ahel I, Altmeyer M, Ashworth A, Bai P, Chang P, Cohen M, Corda D, Dantzer F, Daugherty MD, Dawson TM, Dawson VL, Deindl S, Fehr AR, Feijs KLH, Filippov DV, Gagne JP, Grimaldi G, Guettler S, Hoch NC, Hottiger MO, Korn P, Kraus WL, Ladurner A, Lehtio L, Leung AKL, Lord CJ, Mangerich A, Matic I, Matthews J, Moldovan GL, Moss J, Natoli G, Nielsen ML, Niepel M, Nolte F, Pascal J, Paschal BM, Pawlowski K, Poirier GG, Smith S, Timinszky G, Wang ZQ, Yelamos J, Yu X, Zaja R, and Ziegler M (2021) ADP-ribosyltransferases, an update on function and nomenclature. The FEBS journal [DOI] [PMC free article] [PubMed]
  • 159.Azarm K, and Smith S (2020) Nuclear PARPs and genome integrity. Genes Dev 34, 285–301 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Popp O, Veith S, Fahrer J, Bohr VA, Burkle A, and Mangerich A (2013) Site-specific noncovalent interaction of the biopolymer poly(ADP-ribose) with the Werner syndrome protein regulates protein functions. ACS Chem Biol 8, 179–188 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Ray Chaudhuri A, and Nussenzweig A (2017) The multifaceted roles of PARP1 in DNA repair and chromatin remodelling. Nat Rev Mol Cell Biol 18, 610–621 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Singatulina AS, Hamon L, Sukhanova MV, Desforges B, Joshi V, Bouhss A, Lavrik OI, and Pastre D (2019) PARP-1 Activation Directs FUS to DNA Damage Sites to Form PARG-Reversible Compartments Enriched in Damaged DNA. Cell Rep 27, 1809–1821 e1805 [DOI] [PubMed] [Google Scholar]
  • 163.Martin-Hernandez K, Rodriguez-Vargas JM, Schreiber V, and Dantzer F (2017) Expanding functions of ADP-ribosylation in the maintenance of genome integrity. Semin Cell Dev Biol 63, 92–101 [DOI] [PubMed] [Google Scholar]
  • 164.Eisemann T, and Pascal JM (2020) Poly(ADP-ribose) polymerase enzymes and the maintenance of genome integrity. Cell Mol Life Sci 77, 19–33 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Pascal JM (2018) The comings and goings of PARP-1 in response to DNA damage. DNA Repair (Amst) 71, 177–182 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Sukhanova MV, Abrakhi S, Joshi V, Patre D, Kutuzov MM, Anarbaev RO, Curmi PA, Hamon L, and Lavrik OI (2015) Single molecule detection of PARP1 and PARP2 interaction with DNA strand breaks and their poly(ADP-ribosyl)ation using high-resolution AFM imaging. Nucleic Acids Res 44 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Patidar RL, Viera T, Morales JC, Singh N, Motea EA, Khandelwal M, and Fattah FJ (2020) XRN2 interactome reveals its synthetic lethal relationship with PARP1 inhibition. Sci Rep 10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Ye BJ, Kang HJ, Lee-Kwon W, Kwon HM, and Choi SY (2021) PARP1-mediated PARylation of TonEBP prevents R-loop–associated DNA damage. DNA Repair 104 [DOI] [PubMed] [Google Scholar]
  • 169.Lin WL, Chen JK, Wen X, He W, Zarceno GA, Chen Y, Chen S, Paull TT, and Liu HW (2022) DDX18 prevents R-loop-induced DNA damage and genome instability via PARP-1. Cell Rep 40 [DOI] [PubMed] [Google Scholar]
  • 170.de Lange T (2018) Shelterin-Mediated Telomere Protection. Annu Rev Genet 52, 223–247 [DOI] [PubMed] [Google Scholar]
  • 171.Benarroch D, Pisano S, Bermudez-Mendez A, Lototska L, Kaur P, Bauwens S, Djerbi N, Latrick CM, Fraisier V, Pei B, Gay A, Jaune E, Foucher K, Cherfils-Vicini J, Aeby E, Miron S, Londono-Vallejo A, Ye J, Le Du MH, Wang H, Gilson E, and Giraud-Panis MJ (2016) TRF2-Mediated Control of Telomere DNA Topology as a Mechanism for Chromosome-End Protection. Mol Cell 61, 274–286 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.di Fagagna F, Reaper PM, Clay-Farrace L, Fiegler H, Carr P, von Zglinicki T, Saretzki G, Carter NP, and Jackson SP (2003) A DNA damage checkpoint response in telomere-initiated senescence. Nature 426, 194–198 [DOI] [PubMed] [Google Scholar]
  • 173.Cesare AJ, Hayashi MT, Crabbe L, and Karlseder J (2013) The telomere deprotection response is functionally distinct from the genomic DNA damage response. Mol Cell 51, 141–155 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Timashev LA, Babcock H, Zhuang X, and de Lange T (2017) The DDR at telomeres lacking intact shelterin does not require substantial chromatin decompaction. Genes Dev 31, 578–589 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Arnoult N, and Karlseder J (2015) Complex interactions between the DNA-damage response and mammalian telomeres. Nat Struct Mol Biol 22, 859–866 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Cesare AJ (2014) Mitosis, double strand break repair, and telomeres: a view from the end: how telomeres and the DNA damage response cooperate during mitosis to maintain genome stability. Bioessays 36, 1054–1061 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Hayashi MT, Cesare AJ, Fitzpatrick JA, Lazzerini-Denchi E, and Karlseder J (2012) A telomere-dependent DNA damage checkpoint induced by prolonged mitotic arrest. Nat Struct Mol Biol 19, 387–394 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Erdel F, Kratz K, Willcox S, Griffith JD, Greene EC, and de Lange T (2017) Telomere Recognition and Assembly Mechanism of Mammalian Shelterin. Cell Rep 18, 41–53 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Jim CJ, and Cech TR (2021) Shaping human telomeres: from shelterin and CST complexes to telomeric chromatin organization. Nat Rev Mol Cell Biol 22, 283–298 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Moye AL, Porter KC, Cohen SB, Phan T, Zyner KG, Sasaki N, Lovrecz GO, Beck JL, and Bryan TM (2015) Telomeric G-quadruplexes are a substrate and site of localization for human telomerase. Nat Commun 6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Takahama K, Takada A, Tada S, Shimizu M, Sayama K, Kurokawa R, and Oyoshi T (2013) Regulation of telomere length by G-quadruplex telomere DNA-and TERRA-binding protein TLS/FUS. Chem Biol 20, 341–350 [DOI] [PubMed] [Google Scholar]
  • 182.Biffi G, Tannahill D, and Balasubramanian S (2012) An intramolecular G-quadruplex structure is required for binding of telomeric repeat-containing RNA to the telomeric protein TRF2. J Am Chem Soc 134, 11974–11976 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Bryan TM (2020) G-Quadruplexes at Telomeres: Friend or Foe? Molecules 25 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Tomaska L, Nosek J, Kar A, Willcox S, and Griffith JD (2019) A New View of the T-Loop Junction: Implications for Self-Primed Telomere Extension, Expansion of Disease-Related Nucleotide Repeat Blocks, and Telomere Evolution. Front Genet 10, 792. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Van Ly D, Low RRJ, Frolich S, Bartolec TK, Kafer GR, Pickett HA, Gaus K, and Cesare AJ (2018) Telomere Loop Dynamics in Chromosome End Protection. Mol Cell 71, 510–525 e516 [DOI] [PubMed] [Google Scholar]
  • 186.Doksani Y, Wu JY, de Lange T, and Zhuang X (2013) Super-resolution fluorescence imaging of telomeres reveals TRF2-dependent T-loop formation. Cell 155, 345–356 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Yoshimura SH, Maruyama H, Ishikawa F, Ohki R, and Takeyasu K (2004) Molecular mechanisms of DNA end-loop formation by TRF2. Genes to Cells 9, 205–218 [DOI] [PubMed] [Google Scholar]
  • 188.Azzalin CM, Reichenbach P, Khoriauli L, Giulotto E, and Lingner J (2007) Telomeric repeat containing RNA and RNA surveillance factors at mammalian chromosome ends. Science 318, 798–801 [DOI] [PubMed] [Google Scholar]
  • 189.Schoeftner S, and Blasco MA (2008) Developmentally regulated transcription of mammalian telomeres by DNA-dependent RNA polymerase II. Nat Cell Biol 10, 228–236 [DOI] [PubMed] [Google Scholar]
  • 190.Graf M, Bonetti D, Lockhart A, Serhal K, Kellner V, Maicher A, Jolivet P, Teixeira MT, and Luke B (2017) Telomere Length Determines TERRA and R-Loop Regulation through the Cell Cycle. Cell 170, 72–85 e14 [DOI] [PubMed] [Google Scholar]
  • 191.Deng Z, Norseen J, Wiedmer A, Riethman H, and Lieberman PM (2009) TERRA RNA binding to TRF2 facilitates heterochromatin formation and ORC recruitment at telomeres. Mol Cell 35, 403–413 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Vohhodina J, Goehring LJ, Liu B, Kong Q, Botchkarev VV Jr., Huynh M, Liu Z, Abderazzaq FO, Clark AP, Ficarro SB, Marto JA, Hatchi E, and Livingston DM (2021) BRCA1 binds TERRA RNA and suppresses R-Loop-based telomeric DNA damage. Nat Commun 12, 3542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Feretzaki M, Pospisilova M, Valador Fernandes R, Lunardi T, Krejci L, and Lingner J (2020) RAD51-dependent recruitment of TERRA lncRNA to telomeres through R-loops. Nature 587, 303–308 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Montero JJ, Lopez de Silanes I, Grana O, and Blasco MA (2016) Telomeric RNAs are essential to maintain telomeres. Nat Commun 7, 12534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Arora R, Lee Y, Wischnewski H, Brun CM, Schwarz T, and Azzalin CM (2014) RNaseH1 regulates TERRA-telomeric DNA hybrids and telomere maintenance in ALT tumour cells. Nat Commun 5, 5220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Fouquerel E, Parikh D, and Opresko P (2016) DNA damage processing at telomeres: The ends justify the means. DNA Repair 44, 159–168 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Fouquerel E, and Opresko P (2017) Convergence of The Nobel Fields of Telomere Biology and DNA Repair. Photochem. Photobiol 93, 229–237 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Timashev LA, and de Lange T (2020) Characterization of t-loop formation by TRF2. Nucleus 11, 164–177 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Hwang H, Buncher N, Opresko PL, and Myong S (2012) POT1-TPP1 regulates telomeric overhang structural dynamics. Structure 20, 1872–1880 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Tan J, and Lan L (2020) The DNA secondary structures at telomeres and genome instability. Cell Biosci 10, 47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Ghilain C, Gilson E, and Giraud-Panis MJ (2021) Multifunctionality of the Telomere-Capping Shelterin Complex Explained by Variations in Its Protein Composition. Cells 10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Bandaria JN, Qin P, Berk V, Chu S, and Yildiz A (2016) Shelterin Protects Chromosome Ends by Compacting Telomeric Chromatin. Cell 164, 735–746 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Canudas S, Houghtaling BR, Kim JY, Dynek JN, Chang WG, and Smith S (2007) Protein requirements for sister telomere association in human cells. EMBO J 26, 4867–4878 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Broccoli D, Chong L, Oelmann S, Fernald AA, Marziliano N, van Steensel B, Kipling D, Le Beau MM, and de Lange T (1997) Comparison of the human and mouse genes encoding the telomeric protein, TRF1: chromosomal localization, expression and conserved protein domains. Hum Mol Genet 6, 69–76 [DOI] [PubMed] [Google Scholar]
  • 205.Zhong Z, Shiue L, Kaplan S, and de Lange T (1992) A Mammalian Factor That Binds Telomeric TTAGGG Repeats In Vitro. Mol cell biol 12, 4834–4843 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Broccoli D, Smogorzewska A, Chong L, and de Lange T (1997) Human telomeres contain two distinct Myb-related proteins, TRF1 and TRF2. Nat Genet 17, 231–235 [DOI] [PubMed] [Google Scholar]
  • 207.Mattern KA, Swiggers SJJ, Nigg AL, Lowenberg B, Houtsmuller AB, and Zijlmans JMJM (2004) Dynamics of Protein Binding to Telomeres in Living Cells: Implications for Telomere Structure and Function. Mol Cell Biol 24, 5587–5594 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Kaur P, Wu D, Lin J, Countryman P, Bradford KC, Erie DA, Riehn R, Opresko PL, and Wang H (2016) Enhanced electrostatic force microscopy reveals higher-order DNA looping mediated by the telomeric protein TRF2. Sci Rep 6, 20513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Frescas D, and de Lange T (2014) TRF2-tethered TIN2 can mediate telomere protection by TPP1/POT1. Mol Cell Biol 34, 1349–1362 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Ye JZ, Donigian JR, van Overbeek M, Loayza D, Luo Y, Krutchinsky AN, Chait BT, and de Lange T (2004) TIN2 binds TRF1 and TRF2 simultaneously and stabilizes the TRF2 complex on telomeres. J Biol Chem 279, 47264–47271 [DOI] [PubMed] [Google Scholar]
  • 211.Kim SH, Beausejour C, Davalos AR, Kaminker P, Heo SJ, and Campisi J (2004) TIN2 mediates functions of TRF2 at human telomeres. J Biol Chem 279, 43799–43804 [DOI] [PubMed] [Google Scholar]
  • 212.Takai KK, Kibe T, Donigian JR, Frescas D, and de Lange T (2011) Telomere protection by TPP1/POT1 requires tethering to TIN2. Mol Cell 44, 647–659 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Kaur P, Barnes R, Pan H, Detwiler AC, Liu M, Mahn C, Hall J, Messenger Z, You C, Piehler J, Smart RC, Riehn R, Opresko P, and Wang H (2021) TIN2 is an architectural protein that facilitates TRF2-mediated trans- and cis-interactions on telomeric DNA. Nucleic Acids Res 49, 13000–13018 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Poulet A, Pisano S, Faivre-Moskalenko C, Pei B, Tauran Y, Haftek-Terreau Z, Brunet F, Le Bihan YV, Ledu MH, Montel F, Hugo N, Amiard S, Argoul F, Chaboud A, Gilson E, and Giraud-Panis MJ (2012) The N-terminal domains of TRF1 and TRF2 regulate their ability to condense telomeric DNA. Nucleic Acids Res 40, 2566–2576 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Pisano S, Leoni D, Galati A, Rhodes D, Savino M, and Cacchione S (2010) The human telomeric protein hTRF1 induces telomere-specific nucleosome mobility. Nucleic Acids Res 38, 2247–2255 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Lee YW, Arora R, Wischnewski H, and Azzalin CM (2018) TRF1 participates in chromosome end protection by averting TRF2-dependent telomeric R loops. Nat Struct Mol Biol 25, 147–153 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Markrantoni V, and Marston AL (2018) Cohesin and chromosome segregation. Curr Biol 28, 688–693 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Bisht KK, Daniloski Z, and Smith S (2013) SA1 binds directly to DNA through its unique AT-hook to promote sister chromatid cohesion at telomeres. J Cell Sci 126, 3493–3503 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Heath GR, Kots E, Robertson JL, Lansky S, Khelashvili G, Weinstein H, and Scheuring S (2021) Localization atomic force microscopy. Nature 594, 385–390 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Amyot R, Marchesi A, Franz CM, Casuso I, and Flechsig H (2022) Simulation atomic force microscopy for atomic reconstruction of biomolecular structures from resolution-limited experimental images. PLos Comput. Biol [DOI] [PMC free article] [PubMed]
  • 221.Penedo M, Miyazawa K, Okano N, Furusho H, Ichikawa T, Alam MS, Miyata K, Nakamura C, and Fukuma T (2021) Visualizing intracellular nanostructures of living cells by nanoendoscopy-AFM. Sci Adv 7, eabj4990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Schaich MA, Schnable BL, Kumar N, Roginskaya V, Jakielski RC, Urban R, Zhong Z, Kad NM, and Van Houten B (2023) Single-molecule analysis of DNA-binding proteins from nuclear extracts (SMADNE). Nucleic Acids Res 10.1093/nar/gkad095 [DOI] [PMC free article] [PubMed]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1
2
Download video file (7.2MB, mp4)
3
Download video file (4MB, avi)
4
Download video file (7.6MB, wmv)
5
Download video file (9.5MB, mp4)

RESOURCES